Original Articles
 

By Prof. Kurt Heininger
Corresponding Author Prof. Kurt Heininger
Department of Neurology, Heinrich Heine University Duesseldorf, - Germany
Submitting Author Prof. Kurt Heininger
REPRODUCTION

Sexual reproduction, oxidative stress, mutagenesis, Baldwin effect, epigenetics, evolvability, robustness, phenotypic plasticity, Cambrian explosion, stochasticity

Heininger K. The mutagenesis-selection-cascade theory of sexual reproduction. WebmedCentral REPRODUCTION 2013;4(9):WMC004367
doi: 10.9754/journal.wmc.2013.004367

This is an open-access article distributed under the terms of the Creative Commons Attribution License(CC-BY), which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.
No
Submitted on: 02 Sep 2013 08:55:13 AM GMT
Published on: 02 Sep 2013 10:35:25 AM GMT

Abstract


Sexual reproduction is common in eukaryotic organisms from yeasts to humans. However, the question as to why sexual reproduction is so ubiquitous is a conundrum in evolutionary biology. In theory, sexual populations suffer a twofold cost compared to asexual populations. A multitude of theories attempts to explain this conundrum, but none gained wider acceptance. In this comprehensive work with more than 8,500 references I present compelling evidence that it takes a joint holist and reductionist approach to solve this “queen of problems in evolutionary biology”.

In a world of limited resources, the twofold cost of sexual reproduction may only be relevant in organisms that live far below the carrying capacity of their habitat. Near the habitat’s carrying capacity, however, quality (fitness) of offspring is much more important than quantity and the theoretical twofold cost of sexual reproduction has no evolutionary meaning.

Environments never remain static. They continuously undergo changes that alter the fitness landscapes, displacing populations towards suboptimal fitness regions. Sexual reproduction has its evolutionary roots in a microbial stress response. Environmentally challenged microorganisms take up foreign DNA to either use it as template for DNA repair or increase their genetic repertoire. Thus organisms that are poorly adapted to their current environmental conditions “bet-hedge” (as part of a variety of stress-related responses) hoping to boost their fitness. What evolved as reactive process to environmental challenge was genetically accommodated in higher taxa as cybernetic and proactive evolutionary engine. A multitude of both direct and circumstantial evidences reveals the metabolic/oxidative stress, particularly of males, under which higher taxa generate the genetic variation of gametes (e.g. explaining mammalian testicular extra-abdominal descent or lower-temperature comfort zone of male fishes).

Stochastic mutagenesis and epimutagenesis are fraught by the substantially higher risk that mutations are rather deleterious than beneficial. Therefore, the evolutionary fate of mutators is inextricably linked to their effective population size: only with a large population size chances are substantial that some individuals may create beneficial mutations that provide a fitness advantage. Under environmental stress, microorganisms take advantage of their large population sizes, create increased numbers of mutations and let natural selection select the mutants with the best fit to the actual environment. The evolutionary success of sexual reproduction in multicellular organisms was dependent on solving the “mutator-population size dilemma”. In multicellular organisms, the large investment of scarce resources into mature organisms limits population sizes. Exposing small populations of mutated organisms to natural selection would mean a large investment into possibly poorly viable organisms. Evolution selected for an alternative approach: to invest as little as possible into small, mutated units and to bet-hedge with large gamete populations. The huge energetic investment into this “waste” production leads to gonadal functional hypoxia in higher taxa and is associated with increased oxidative stress and genetic instability. During gametogenesis, oxidative stress, mutagenesis and epimutagenesis increase along the phylogenetic axis, particularly in the smaller gametes, the sperm, resulting in male-driven mutagenesis. In addition to the legendary twofold cost of sex this manifold “costs of gamete overproduction” (e.g. the human testicular output is approx. 2,000 spermatozoa per second), if it were non-adaptive, should definitely deter any organism from investing into sexual reproduction.

Gamete overproduction requires a gamete bottleneck: a stochastic bottleneck inevitably would drive Muller’s ratchet and result in strong mutational meltdown and, together with their low number of offspring, this would have detrimental effects on population fitness in higher taxa. Only a selection-associated gamete bottleneck ensures long-term viability of populations. Conversely, the selection-associated gamete bottleneck allows to infer the (epi)mutagenic generation of gamete variation. Thus sexual reproduction selects the “pearls among the pebbles” before natural selection may possibly frustrate a larger investment into more mature organisms. Using a variety of stressors and competitive regimes as selection principle, e.g. oxidative stress and competition for limited trophic factors, a casacade of selective steps at the level of (primordial) germ cells, differentiated gametes, fertilization, embryos, offspring and nonrandom mating select from this extreme abundance the most viable, competitive and stress resistant gametes and offspring. These selection principles are increasing mutational and environmental robustness and evolvability/adaptation to life in general, rather than to any particular mode of life and environment. The pervasive germ-soma conflict is the Red Queen that drives the coevolutionary dynamics underlying the maintenance of sexual reproduction. Along the phylogenetic axis, the mutagenic spectrum and the population size of gametes that is exposed to sexual selection cascades were increased and the population size of offspring that is exposed to natural selection was reduced. As a result of this evolutionary process, particularly male gametes represent quasi-species whose quasi-species structure, however, remains cryptic due to their gamete bottleneck. Oxidative stress as final common pathways of stress responses is the joint mediator of mutagenesis, epimutagenesis, canalization and gamete quality control by gamete competition. Thus both evolvability and robustness are phenotypes of the same fundamental processes, mediating “educated guess”-based innovation, plasticity and phenotype conservation in the face of environmental and genetic variation.

Ecological resource availability and stress intensity define the distribution of sexual and asexual reproduction. In resource-limited, moderately stressful habitats, the pre-selected genetic variation provided by sexual mutagenesis-selection cascades determines the short- and long-term benefit of sexual reproduction. The flip side of oxidative stress-driven mutagenesis-selection-cascades is that sexual reproduction, particularly male reproduction, became sensitive to additional environmental stressors and, particularly when exposed to extreme environmental challenges, operates near or above a threshold of ‘‘error catastrophe’’. Thus, in larger animals geographic parthenogenesis is the phenotype of impaired male gametogenesis under intense environmental stress. On the other hand, a favorable ratio of habitat resource availability and minute individual biomass allows asexually reproducing bdelloid rotifers, Darwinulid ostracods, and orbatid mites a microorganism-like lifestyle with large population sizes and species-specific solutions with regard to their balance between evolvability and robustness in response to environmental challenges. In cyclical parthenogenesis, ecological factors that modulate the resource-stress balance determine the switch between asexual and sexual reproduction. Modular organisms (e.g. plants, benthic aquatic invertebrates) do not fall under Weismann’s doctrine (the separation of germline and soma), and in addition to creating genetic variation by shuffling their genes during sexual reproduction are able to transmit heritable somatic mutations that are acquired and selected during the organism’s lifetime. In addition, polyploidy, long-term seed/egg banks, and phenotypic plasticity help these organisms to face the manifold challenges of variable environments. This capacity makes them flexible to switch habitat- and disturbance-dependently between asexual and sexual reproduction.

I can show that the predictions of the genetic theories of sexual reproduction (Fisher-Muller model, Muller’s ratchet, Kondrashov’s hatchet) are simplistic, incompatible with evolutionary dynamics, and based on fallacious post hoc, ergo propter hoc argumentation. In addition, it is argued that the DNA repair theory and parasite-host conflict theory (aka Red Queen theory), focussing on partial aspects of the resource-stress ecological network are insufficient to explain the evolutionary success of sexual reproduction.

That stochasticity is a property of evolutionary mechanisms like mutation, recombination and drift has been recognized earlier but I argue that stochasticity is, in addition to selection, the second pillar and organizing principle on which evolution is resting. Bet-hedging is the evolutionary risk-spreading response to stochasticity. Sexual reproduction is the ultimate bet-hedging enterprise and its evolutionary success the selective signature of stochastic environments. The identification of stochasticity/uncertainty as input variable into the cybernetic machine of evolution has far-reaching implications for a multitude of evolutionary enigmas such as the levels of selection enigma, the selection-genetic variation enigma, the adaptationism controversy, the mystery of low heritability and the Darwinian-Lamarckian type evolution controversy. Since sexual reproduction is so pervasive in nature and involved in the vast majority of genetic transmissions, deciphering its underlying mechanisms provides, as a corollary, insights into a variety of evolutionary conundrums such as the viability paradox, the molecular evolution–population size conundrum, and latitudinal biodiversity gradients. Moreover, sexual reproduction and the germ-soma conflict are identified as the processes causing punctuated evolution (e.g. the Cambrian explosion), and a mixed genetic/non-genetic type heritability. Finally, Francis Crick’s central dogma of molecular biology is extended taking into account the feedback of the environment to the genome via the phenotype using the noncoded signals energy (ATP), Ca2+, and redox homeostasis (reactive oxygen and nitrogen species). This feedback cycle reveals that environmental modulation of the epi/genome is subject to the dualism of stochasticity (resulting in variation) and selection vindicating the Darwinian evolutionary principles.

Table of contents

1. Introduction
2. An ecological definition of stress
3. Reproduction in a world of limited resources
  3.1 Germline-derived signals limit the soma’s reproductive potential
4. The phylogenetic roots of sexual reproduction
  4.1 Oxidative and nitrosative stress, DNA repair and mutagenesis
  4.2 Stress and microbial transformation
  4.3 Stress and multicellular reproductive behavior
5. Mutagenesis and fitness
  5.1 Are mutations genetic “accidents”?
  5.2 May mutagenesis be adaptive?
6. Developmental and reproductive biology hold the clue
  6.1 A primer on the evolutionary roots of multicellular development
  6.2 A primer on reproductive biology
7. Sex: Proactive mutagenesis….
  7.1 Sexual reproduction: domesticating the fire
  7.2 The oxidative stress of gametogenesis
    7.2.1 Functional hypoxia
    7.2.2 Metabolic and replicative stress
    7.2.3 Oxidative stress
    7.2.4 Hypoxia-inducible factors
    7.2.5 Cytokines and nuclear factor-kappaB
    7.2.6 Heat shock response
    7.2.7 Stress and germ granules
    7.2.8 DNA damage and repair
    7.2.9 Mitochondrial ROS, uncoupling and aquaporin-8
    7.2.10 Membrane lipid unsaturation
  7.3 Male-driven mutagenesis
    7.3.1 Chromatin remodeling in elongating spermatids
    7.3.2 The gametogenesis-cancerogenesis connection             
8. Sex: ….and selection of the “pearls among the pebbles”
  8.1 Selection: the pervasive phenomenon in evolution             
    8.1.1 Cell competition and Myc
    8.1.2 Germ cell competition
  8.2 Germ cell selection
    8.2.1 Case study: paternal age effect disorders
  8.3 Selective mitochondrial bottleneck              
    8.3.1 Mitochondrial homoplasmy
    8.3.2 Follicle atresia
  8.4 Gametic selection
  8.5 Fertilization selection
  8.6 Embryo and offspring selection
  8.7 Sexual selection
9. Coevolutionary dynamics of the germ-soma conflict
10. The redox regulation of gametogenesis and development: mutagenesis, epigenesis, canalization and gamete quality control
  10.1 Mutagenesis
  10.2 Recombination
  10.3 Epimutagenesis
    10.3.1 DNA methylation
    10.3.2 Histone modifications
    10.3.3 Noncoding RNA
    10.3.4 Epi/genetic reprogramming
  10.4 Canalization
  10.5 Apoptosis
  10.6 Transposable elements and epigenetics
    10.6.1 Role of microRNAs in gametogenetic DNA instability
11. Sexual mutagenesis-selection cascades (SMSC): there is more than meets the eye
  11.1 Excursion: RNA virus quasispecies
  11.2 SMSC, bottlenecks and natural selection
12. Random trial or educated guess?
  12.1 Stress-induced mutagenesis
  12.2 Non-random recombination
  12.3 Mutability of simple sequence repeats
  12.4 Activity of transposable elements              
  12.5 Phenotypic plasticity, genetic assimilation and accommodation
    12.5.1 Assimilation and accommodation
  12.6 Sexual reproduction
13. Sexual reproduction: evolvability and robustness
  13.1 Evolvability
    13.1.1 The ecological pattern of evolvability
    13.1.2 The tempo of evolution
  13.2 Robustness
14. Stress and sex: a double-edged relationship             
  14.1 Male reproduction is more susceptible to a variety of stressors
  14.2 Hormonal modulation of oxidative stress as evolutionary tuning knobs of reproductive activity and genetic variation
    14.2.1 Glucocorticoids
    14.2.2 Thyroid hormones
    14.2.3 Melatonin
15. Asexual reproduction
  15.1 Non-sexual ways to increase evolvability
    15.1.1 Large population size
    15.1.2 Polyploidy
    15.1.3 Automixis and mitotic recombination
    15.1.4 Phenotypic plasticity
    15.1.5 Seed/egg banks and dormancy as bet-hedging strategies
  15.2 Asexual reproduction in sessile organisms
  15.3 Asexual reproduction in mobile animals
    15.3.1 The “evolutionary scandals”: microscopic organisms in r-selected habitats
    15.3.2 Cyclical parthenogenesis
    15.3.3 Geographical parthenogenesis: when males are too stressed to reproduce
16. Germ granules and transgenerational epigenetic information transfer
  16.1 Germ granules are germ cell markers
  16.2 Transgenerational epigenetic information transfer
  16.3 Are germ granules the vehicles of transgenerational epigenetic information?
17. Stochasticity and selection, the organizing principles of evolution
  17.1 The cybernetics of evolution
  17.2 Bet-hedging as response to stochasticity
  17.3 Stochasticity and selection: dualism in evolution
18. Earlier theories that attempted to explain why most organisms reproduce sexually
  18.1 Fisher-Muller-model, Muller’s ratchet and Kondrashov’s hatchet
    18.1.1 Flawed static concepts               
    18.1.2 Flawed teleological concepts
  18.2 DNA repair
  18.3 Host-parasite coevolution: the Queen is dead, long live the Queen
19. Evolutionary enigmas and controversies
  19.1 The selection-genetic variation enigma
  19.2 Levels of selection controversy
  19.3 The selectionist-neutralist controversy
  19.4 The adaptationism controversy
  19.5 Paradox of viability
  19.6 Mystery of low heritability
  19.7 The Cambrian explosion
  19.8 The Darwinian-Lamarckian evolution controversy
20. The resource-stress dimensions and ecological window of sexual/asexual reproduction
  20.1 The resource-stress dimensions
  20.2 The ecological window of asexual and sexual reproduction
21. Extension of Crick’s central dogma of molecular biology
22. Conclusions
23. Abbreviations
24. References

Introduction


Sex is the queen of problems in evolutionary biology. Perhaps no other natural phenomenon has aroused so much interest; certainly none has sowed so much confusion.

Graham Bell, 1982

It is […] surprising that finding a convincing explanation for the evolutionary success of sex has proven to be one of the most difficult challenges for modern evolutionary biologists.

Nick Colegrave, 2012

Summary

Sexual reproduction is common in eukaryote organisms from yeasts to humans. But the near ubiquitous presence of sex has proven to be notoriously hard to rationalize in terms of an evolutionary cost–benefit perspective. Traditionally it has been argued that the evolutionary cost of sex is at least twofold given that sexual females would need a twofold amount of offspring (with a 1:1 sex ratio) to keep up with reproduction of asexual females. On the other hand, there is as of now no simple explanation for the benefit(s) of sex. Hypotheses for the maintenance of sex have been grouped into two main categories: (i) ecological and (ii) genetic theories. No single theory has obtained unanimous support and the rationale for the origin and maintenance of sex remains unresolved.

At its heart, evolutionary theory has an economic algorithm, a cost–benefit calculation: if a trait is beneficial for fitness, it will be selected for, when it is costly, it will be selected against. Reproduction is arguably the most fundamental activity of an organism. Reproduction is the sine qua non for the propagation of species and continuation of life. The vast majority of eukaryotic organisms reproduce sexually — at least occasionally. Among named animal species, only 0.1% are considered to be exclusively asexual (White, 1978; Vrijenhoek, 1998) and less than 0.1% of vertebrate species (some 90 species) (Dawley, 1989; Vrijenhoek et al., 1989; Vrijenhoek, 1989). Eighty vertebrate taxa from 14 families of fish, amphibians and reptiles have been found to reproduce asexually and are usually closely related to sexually reproducing species (Alves et al., 2001; Janko et al., 2007) with many examples demonstrating coexistence of unisexual and bisexual populations (Moore and McKay, 1971; Schlupp, 2005; 2010). As of 2003, over 900 species of insects were known to be asexual or have asexual forms (Normark, 2003). Asexual reproduction is mainly confined to small twigs in the phylogenetic tree suggesting that asexual lineages have a higher extinction rate than sexual lineages, possibly due to mutational meltdown (Lynch et al., 1993, Vrijenhoek, 1998; Schwander and Crespi, 2009). While the ability to reproduce asexually is more widespread in plants, less than 1% of the approximately 250,000 angiosperm species are thought to be substantially asexual (Asker and Jerling, 1992; Whitton et al., 2008). Sex is argued to be advantageous because it generates variability by allowing independent assortment of genetic material through recombination (Weismann, 1889; Muller, 1932). But the near ubiquitous presence of sex has proven to be notoriously hard to rationalize in terms of evolutionary cost–benefit calculations (Maderspacher, 2011). There is as of now no simple, single-cause explanation for the benefit(s) of sex (Birky, 2009; Otto, 2009). On the other side of the equation, the costs of sex are easily spelled out. According to classical theory, reiterated in almost every publication on the evolutionary role of sexual reproduction, the evolutionary cost of sex is at least twofold (Maynard Smith, 1978a). Sexual populations consist of two genders, one (female) which can produce offspring and the other (male) which cannot. However, in an asexual population, any individual can produce its own offspring. Sexual females would need a twofold amount of offspring (given a 1:1 sex ratio) to keep up with reproduction of asexual females. Thus, asexuals should rapidly outgrow sexuals, which they usually don’t. The second most obvious benefit is the twofold fitness advantage of an asexual female (Williams, 1975; Schwander and Crespi, 2009). For the same number of offspring produced, each asexual female transmits twice as many genes to each offspring as compared to a female producing offspring sexually, the “costs of meiosis” (Williams, 1975).

A more comprehensive inventory over all groups of organisms showed five potential costs of sex (Lewis, 1987; Joshi and Moody 1998; Hörandl, 2009): (1) cellular–mechanical costs of meiosis and syngamy (the time needed for meiosis, syngamy and karyogamy, which confers an inactive stage and a delay of synthetic processes); (2) recombination (breakup of favorable combinations of alleles); (3) fertilization (risk exposure and density dependence of locating a mating partner); (4) cost of males and ‘genome dilution’ sensu Lewis, 1987 (a sexual parent transmits only 50% of its genes to the offspring); and (5) costs of sexual selection. In fact, sexual selection has generally been viewed as reducing population mean fitness (Haldane, 1932; Lande, 1980; Kirkpatrick, 1982; Grafen, 1990; Kirkpatrick and Ryan, 1991; Price et al., 1993; Tanaka, 1996; McLain et al., 1999; Gavrilets et al., 2001; Houle and Kondrashov, 2001). Whereas (1) and (2) are costs of meiosis, (3) – (5) are costs of outcrossing.

Agrawal (2002) pointed to a key difference in the equilibrium genetic load (the reduction in fitness of a population of interest relative to a population composed solely of the most-fit genotype) in sexual and asexual populations under constant and fitness-dependent mutation rates. If the mutation rate is constant, then sexual and asexual populations are expected to have the same genetic load at mutation-selection equilibrium. In contrast, if the mutation rate is fitness-dependent, the expected fitness in an asexual population at equilibrium will be that of the most-fit genotype, while in a sexual population, the equilibrium fitness will be less than that of the most-fit genotype, perhaps much less. Thus surprisingly, the “cost of sex” may in fact be much greater than 2-fold (Baer, 2008). In chapter 8.1, I will introduce another cost of sexual reproduction, the overproduction of a huge “waste” of gametes. In the face of such legendary costs, we might expect sexual reproduction to be rare. Given the costs of sex and the widespread potential for asexual reproduction, why do so many species reproduce sexually? The question as to why sexual reproduction is so pervasive is a conundrum that has been described as “the outstanding puzzle in evolutionary biology” (Williams, 1975; Maynard Smith, 1978a; Bell, 1982; Stearns, 1987a; Otto, 2009).

Like Birdsell and Wills (2003), I follow the example of Margulis and Sagan (1986) that defined sex in the broadest sense to include any natural process that combines genes from more than a single source in an individual cell. This definition of sex thus includes any horizontal transmission of genetic materials in natural environments and includes conventional sex in eukaryotes that involves cell fusion and meiosis; genetic exchange among viral particles; and prokaryotic sex such as transformation, transduction, and conjugation. To avoid confusion, recombination is defined as breakage and reunion of genetic materials in genomes (DNA or RNA) and includes both crossing-over and independent assortment. Reproduction refers to processes that replace or increase the number of cells or organisms in populations (Xu, 2004a).

The main difficulty with the evolution of recombination is that parents who survive to reproduce have a genotype that works under local environmental conditions. Tinkering with their offsprings’ genotypes by recombination risks making the situation worse, with little hope of making it better (Otto and Michalakis, 1998). This idea underlies the Reduction Principle: in a constant environment without mutation, only modifier alleles that reduce recombination can invade a randomly mating population at equilibrium (Feldman, 1972; Charlesworth, 1976; Barton, 1995; Feldman et al., 1997; Otto and Michalakis, 1998). Lower recombination rates evolve whenever there are genetic associations among loci (linkage disequilibria), whether these are positive or negative (Feldman, 1972; Charlesworth, 1976; Barton, 1995; Feldman et al., 1997). Hypotheses for the maintenance of sex can be grouped into three main categories (Kondrashov, 1993; West et al., 1999; Birdsell and Wills, 2003; Hörandl, 2009; Otto, 2009): (1) The ecological theories posit sex as an effective way to create genetic variation in the offspring, which allows for a faster adaptation to environmental variability (for example, Burt, 2000). (2) The genetic theories consider sex as a restoration mechanism for damage of DNA strands, or mutational or epimutational change of the genome; here the main arguments are that meiosis would provide a tool for repair of double-strand breaks (e.g., Bernstein et al., 1988; 1989; Michod, 1995) or would restore DNA methylation patterns (for example, Holliday, 1984). Moreover, sex, allowing for recombination and segregation, is thought to be advantageous because beneficial mutations that arise in different individuals can be united in the same genome and therefore offspring with above average fitness can be produced (Fisher, 1930; Muller, 1932). Recombination would also avoid a long-term accumulation of disadvantageous mutations; an asexual lineage cannot produce offspring with a lower mutational load than any previous generation (Muller’s ratchet) (Felsenstein, 1974). In addition, recombination can break up negative epistasis and thus allows for a more efficient purging of deleterious mutations (Kondrashov, 1988; 1993). The latter two advantages are predicted to strongly outweigh the reproductive disadvantage of sexual reproduction. On the other hand, Muller’s ratchet and Kondrashov’s hatchet are predicted to drive unisexual lineages into extinction within 104 to 105 generations (Lynch and Gabriel, 1990). (3) A third group of hypotheses regards sex as a consequence of phylogenetic fixation, that is, a feature inherited from ancestors that organisms cannot get rid of. This model does not seek an advantage of sexual reproduction per se, but regards it as an ‘imperative relic’ inherited from eukaryotic ancestry (Margulis and Sagan, 1986). The latter theory implies that sex is maladaptive but for some evolutionary constraints cannot be lost. A variety of taxa that can reproduce both sexually and asexually (see chapter 15.3.2) clearly refute this theory. Since I will show that sex is adaptive I will not delve into this theory that, moreover, has not found much support. Many modern authors argue for pluralistic approaches and combinations of ecological- and genetic-based models (Howard and Lively, 1994; 2002; West et al., 1999; Pound et al., 2004; Ben-Ami and Heller, 2005; Cooper et al., 2005; Bruvo et al., 2007; Hörandl et al., 2008; Hörandl, 2009; Neiman and Koskella, 2009; Park AW et al., 2010). However, no single theory has obtained unanimous support and the evolution and maintenance of sex remain unresolved (Birky, 2009; Otto, 2009; Colegrave, 2012; Hartfield and Keightley, 2012). Hypotheses concerning the phylogenetic history and age of asexual taxa have been especially controversial (Judson and Normark, 1996; Little and Hebert, 1996), due to the paucity of reliable phylogenetic data for clades with sexual and asexual species and the difficulty of inferring dates of speciation events. Ancient asexual taxa may yield critical clues to understanding the causes of the evolution of sex, as aspects of their genetics and ecology presumably differ from shorter-lived asexuals in ways that indicate the selective advantages of asexual and sexual systems (Judson and Normark 1996).

Recently, I published a paper explaining the ecological-evolutionary causation of aging (Heininger, 2012). I could show that aging/death and reproduction are co-selected. I urgently advise the reader of this paper to make himself familiar with some of the concepts in this paper, e.g. the limited resources paradigm, the germ-soma conflict and its Red Queen coevolutionary dynamics.

Referring to “The heritability hang-up” Feldman and Lewontin (1975) wrote: “There is a vast loss of information in going from a complex machine to a few descriptive parameters. Therefore, there is immense indeterminacy in trying to infer the structure of the machine from those few descriptive variables, themselves subject to error. It is rather like trying to infer the structure of a clock by listening to it tick and watching the hands”. The hypothetico-deductive theories start from assumed laws or premises taken from population genetics and make inferences concerning the operation of sexual reproduction.

According to Bell (1982), robust evaluation of alternative hypotheses for the evolution of sex and asexuality requires integration of phylogenetic, biogeographic, ecological and genetic data. Thus, my approach is to open the clock to observe the gearwheels in the clockwork and infer general principles or rules from these specific facts. I look at the triggering event(s) and the processes that evolution “devised” to reach a goal. With this approach I try to reconstruct the evolutionary rationale behind the phenomenon starting from unicellular organisms and tracing the phylogenetic path along which patterns unfold. Most importantly, this approach is based solely on empirical and experimental observations (documented by more than 8,500 references) instead of assumed laws or premises as is inherent to approaches based on population genetics (see chapter 18.1). My approach has provided compelling evidence for the co-selection of reproduction and death and the programming of aging (Heininger, 2012).

In the following, I will elaborate a comprehensive theory explaining the evolutionary roots and mechanisms underlying the evolutionary success of sexual reproduction. The complexity of the causation and regulation of sexual reproduction is reflected by the argumentation and hence the paper is no easy reading. I have tried to assist the reader by providing short summaries at the beginning of the main chapters. This assistance may abet a rather saltatory style of reading with intermittent full text reading depending on the reader's focus of interest. Taking the potentially saltatory style of reading and the complexity of the topic into account there is some inbuilt redundancy of argumentation both to keep track and stress the consistency and consilience of data.

2. An ecological definition of stress


Knowledge is the object of our inquiry, and men do not think they know a thing till they have grasped the ‘why’ of it (which is to grasp its primary cause).

Aristotle

Summary

An ecological definition of stress should make a distinction between a stressor (external factor), stress (an internal state brought about by a stressor), and stress response (a cascade of internal and external changes triggered by stress). Stress is here defined as an environmental condition to which organisms are poorly adapted and that reduces Darwinian fitness. What is perceived as stressor depends on the evolutionary and ecological history of an organism, a change in the usual environmental conditions for any given life form.

The daily routines of animals and humans alike include nutritional inputs to maintain normal activities and to anticipate additional requirements (e.g., breeding, migrating, acclimating to cold and heat, etc.) during the day–night cycle and the seasons. These homeostatic mechanisms, including functional and structural changes in brain and body, allow the individual to maintain physiological and behavioral stability despite fluctuating environmental conditions. (Throughout this work, when not specified otherwise, environmental conditions will refer both to abiotic and biotic conditions.) It is more than an empty tradition when dealing with stress to engage the problem of definition (Greenberg and Wingfield, 1987). The stress concept means too many things to too many workers to do otherwise, and clearly definition can set the research agenda. In common usage, stress usually refers to an event or succession of events that cause a response, often in the form of “distress” but also, in some cases, referring to a challenge that leads to a feeling of exhilaration, as in “good” stress. But, the term “stress” is full of ambiguities. It is often used to mean the event (stressor) or, sometimes, the response (stress response). Furthermore, it is frequently used in the negative sense of “distress”, and sometimes it is used to describe a chronic state of imbalance in the response to stress (McEwen and Wingfield, 2003; Romero et al., 2009; Koolhaas et al., 2011). The definition of Hoffmann and Parsons (1991) is widely used that defined stress as an ‘environmental factor causing a change in a biological system, which is potentially injurious‘. An ecological definition of stress should make a distinction between a stressor (external factor), stress (an internal state brought about by a stressor), and stress response (a cascade of internal and external changes triggered by stress) (Parker et al., 1999; van Straalen, 2003). Stress is here defined as an environmental condition to which organisms are poorly adapted and that reduces Darwinian fitness (Sibly and Calow, 1989; Zhivotovsky, 1997). Optimal stress responses are those that maximize Darwinian fitness (Sibly and Calow, 1989). Stressful local ecological processes to which organisms are by definition poorly adapted, possibly associated with colonizations and introductions, include exposure to: (1) a novel host, infectious agent, or limited food resource; (2) new, variable or unpredictable biophysical environments; (3) a new predator community; and (4) a new coexisting competitor (Reznick and Ghalambor, 2001). Stress, in any form, exerts strong (co)evolutionary pressures on all organisms. To survive, any organism must develop tolerance, resistance or avoidance mechanisms. Tolerance allows the organism to withstand the challenge unharmed. Resistance involves active countermeasures, while avoidance prevents exposure to the stressors. In addition to ecological stressors, internal stresses, such as spontaneous gene mutations or aberrant cell division might cause adverse effects on metabolic or genetic regulation.

Selye (1936) was the first to recognize the relative uniformity of the general stress response (GSR) (that he dubbed “general adaptation syndrome”) to diverse stressors. The evolutionarily highly conserved GSR has been studied extensively in yeast, animals and plants (Kültz, 2003; 2005; Fujita et al., 2006). At the cellular level, the GSR results in oxidative and nitrosative stress as the final common pathway and a response pattern including heat shock protein expression (Lindquist, 1986; Sanchez et al., 1992; Finkel and Holbrook, 2000; Mikkelsen and Wardman, 2003; Sørensen et al., 2003). Most importantly, stress is not only an attribute of the stressor (the environmental component), but also an attribute of the stressed (the biological component). What is perceived as stressor depends on the evolutionary and ecological history of an organism, a change in the usual environmental conditions for any given life form. A certain environment may be claimed as stressful only if considered with respect to both a given population and the environment in which the population has evolved (Zhivotovsky, 1997; Bijlsma and Loeschcke, 2005). It follows that while a specific condition (e.g., a temperature of 65°C) may be stressful (or even lethal) to a certain species that normally lives at 37°C, it will be optimal for growth to a thermophilic organism (Conway de Macario and Macario, 2000). As an extreme example, the deep-sea barophilic hyperthermophile Thermococcus barophilus obviously experiences stress when grown under atmospheric pressure (Marteinsson et al., 1999). Similarly, abundant resources are stressful for human populations that have been selected for their thrifty genotype (Neel, 1962; Editorial (1989); Hales and Barker, 1992; Allen and Cheer, 1996; Fernández-Sánchez et al., 2011).

3. Reproduction in a world of limited resources


Of the ecological forces that shape evolution of life histories, competition seems particularly important. When competition is absent, those genotypes which reproduce more rapidly and prolifically are favored; when resources are limiting the advantage goes instead to those best able to compete.

Taylor and Condra (1980)

Summary

Traditional concepts with their “all else being equal” scenarios emphasize the twofold costs of sex. The implicit assumption of these scenarios is the unlimited availability of resources. A resource-rich, noncompetitive, r-environment selects for traits that enhance population growth rate, including early maturity, small body size, high reproductive effort, and high fecundity. However, most habitats are resource-limited. Resource-limited, competitive, K-environments select for traits that enhance persistence of individuals, including delayed maturity, large body size, high investment in individual maintenance at the cost of low reproductive effort, low fecundity with a large investment in each offspring, and longer lifespan. In resource-limited habitats quality of offspring is more important than quantity and individuals that invest into more competitive individuals are selectively favored. Thus, in resource-limited habitats, offspring quantity-quality trade-offs are observed in sexually reproducing plants and animals. Underscoring the self-containment of reproduction, reproductive activity is self-limited by negative feedback loops in sexually reproductive organisms.

Dobzhansky notoriously said in 1964: “Nothing in biology makes sense except in the light of evolution.” This was supplanted half a century later by Grant and Grant’s (2008): “Nothing in evolutionary biology makes sense except in the light of ecology.” Pelletier et al. (2009) followed with “Nothing in evolution or ecology makes sense except in the light of the other,” and this sentiment is pretty much where we are today (Schoener, 2011). Kokko and Lopez-Sepulcre (2007) call this “ecogenetic feedback”: “If density influences everyone’s reproductive prospects to the same extent, one has merely restated the ecological concept of density dependence. But if density variation has a differential effect on individual fitness depending on…phenotype, we have a feedback loop. In this loop, individual behavior or life history, influenced by genes, has an effect on population dynamics…and the resulting change in population dynamics in turn…[may] differentially favour…[certain] genotypes…in the population….”.

According to Mallet (2010), the earliest ecologists to investigate competition were motivated by an interest in Darwin's 'struggle for existence,' or natural selection (Gause 1934; Scudo and Ziegler 1978), but later ecologists focused mainly on population densities of competing species. In contrast, the originators of population genetic theory, Ronald A. Fisher, J.B.S. Haldane and Sewall Wright, based their theories on population ecology, but generalized selection almost exclusively in terms of gene frequencies within species. As a result, textbooks today treat population ecology and evolution by natural selection as almost entirely separate topics. This separation is reasonable if gene frequencies and population density do not interact. However, interaction is likely: selection is caused by differences in fertility and survival, the same parameters that affect population density (Mallet, 2010).

Competition among species closely resembles natural selection among genotypes, so it ought to be possible to build theories of population genetics equivalent to ecological competition. However, efforts to unify ecology and evolution have been frustrated because theories of population growth and competition conflict with common sense about evolution. In particular, ecological theory seems to demand that genotypes with the highest carrying capacities K will be fitter, while intrinsic growth rate r does not affect the outcome (MacArthur, 1962; Roughgarden, 1971); meanwhile classical evolutionary theory employs r as the basis of fitness (Fisher, 1930).

Traditionally, hypotheses for the evolution and maintenance of sex have been grouped into ecological and genetic theories (West et al., 1999; Birdsell and Wills, 2003; Hörandl, 2009; Otto, 2009; Hartfield and Keightley, 2012), reflecting the scientific divide between ecology and population genetic theory. Many previous theories and studies of the evolution of sex have implicitly assumed infinite populations that have access to limitless resources (Ackerman et al., 2010). Limited resources, however, are a pervasive selective force (Heininger, 2012). Classical concepts with their “all else being equal” scenarios emphasize the twofold costs of sex (Maynard Smith, 1971a; 1989; Jokela et al., 1997; Peck et al., 1999), but, ignoring the selective force of limited resources, are flawed from their basic assumptions. It is assumed that asexuals would produce twice as many offspring capable of reproducing than the sexual individuals, and would soon dominate the population. Tacit implication of this assumption (only realized by man-made culture conditions) is the inexhaustibility of resources. Indeed, this may happen when scientists introduce asexual mutants into cultures of sexual organisms such as yeast, algae or some rotifers (Hayden, 2008). But even with unlimited resources this assumption is not always correct. Sexual and asexual Potamopyrgus snails had an almost identical reproductive output (Jokela et al., 1997). Lamb and Willey (1979) reported lower hatching rates in parthenogenetic Drosophila. In reptiles, a study comparing sexual and parthenogenetic Aspidocelis lizards found no difference in fecundity between the two reproductive forms (Congdon et al., 1978), while asexual geckos (Heteronotia) have a 30% smaller size-corrected fecundity as compared to several sexual congeners (Kearney and Shine, 2005).

In this chapter, I will gauge how realistic this assumption is in the wild. Intriguingly, this resource allocation argument (Uyenoyama, 1984) has been rarely discussed in the context of ecological resource availability.

Since Darwin, it has been one of the central tenets of evolutionary theory that organisms maximize their fitness including their reproductive output (Darwin 1859; Dawkins 1976; Dennett 1995; Grafen, 1999). However, organismal fitness as measured by reproductive success is not relatively constant from generation to generation but is density- and time-scale-dependent (Goodnight et al., 2008). The short-term reproductive success of a genotype does not necessarily correspond to its long-term success (Rauch et al., 2002). There are evolutionary scenarios that may favor some kind of reproductive self-restraint. Self-limitation of reproduction attenuates the oscillations predicted by the Lotka-Volterra equation, stabilizes populations and prevents their extinction (Mitteldorf et al., 2002; Rauch et al., 2002; 2003; Killingback et al., 2006; Mitteldorf, 2006; 2010; Goodnight et al., 2008). The feedback between exploitation and extinction is seen to have the effect originally proposed as the ‘‘prudent predator’’ concept (Slobodkin, 1961). Modern concepts advocate optimization of reproduction instead of maximization as “sound” evolutionary strategy (de Magalhães and Church, 2005; Grafen, 2006). In fact, a multitude of genes operate to limit the reproductive potential of organisms (Heininger, 2012). Gonadal hormones both advance and limit reproductive activity (see chapter 3.1).

The concept of density-dependent natural selection was proposed in the original r/K-selection model (MacArthur and Wilson, 1967, Pianka, 1970; Boyce, 1984). The terms refer to the suggestion that r-selection will increase r (the instantaneous growth rate), whereas K-selection will increase K (the equilibrium size, usually called the carrying capacity) (Armstrong and Gilpin, 1977). This classification has been of enormous heuristic value in ecology, although it is of only limited value in describing nature (Begon and Mortimer, 1986; Rogers, 1992). (Dawkins [1982] observed that “ecologists enjoy a curious love/hate relationship with the r/K concept, often pretending to disapprove of it while finding it indispensable”). The central idea of r- and K-selection is that populations living in environments imposing high density-independent mortality (r-strategists) will be selectively favored to invest into population growth while populations living in environments imposing high density-dependent regulation (K-strategists) will be selectively favored to invest into more competitive individuals (Gadgil and Solbrig, 1972). Even under nearly ideal conditions in the laboratory, the maximum value of r which can be attained is greatest for small organisms and declines with increasing size (Blueweiss et al., 1978; Bell and Koufopanou, 1991). Lotka-Volterra competition seems to demand that species or genotypes with the highest carrying capacities K will be fitter. The intrinsic growth rate r does not affect the outcome of competition. At the same time, classical evolutionary theory uses r as a measure of fitness (Fisher, 1930; Mallet, 2010). MacArthur (1962) provided genetic models of how density-dependent selection should act and concluded that in equilibrium populations, "the carrying capacity, K, replaces fitness (r) as the agent controlling the action of natural selection". A resource-rich, noncompetitive, r-environment selects for traits that enhance population growth rate, including early maturity, small body size, high reproductive effort, and high fecundity. Conversely, resource-limited, competitive, K-environments select for traits that enhance persistence of individuals, including delayed maturity, large body size, high investment in individual maintenance at the cost of low reproductive effort, low fecundity with a large investment in each offspring, and longer lifespan. These alternative constellations of life-history traits became known as life-history strategies (Pianka 1970; 1974a; Reznick et al., 2002).

The basic premise of density-dependent selection theory is that genotypic fitness is a function of population density (Joshi et al., 2001). Thus, for populations at low densities well below carrying capacity, natural selection favors genotypes with the highest value of r. In populations at carrying capacity, genotypes with the highest value of K are favored. Recent theoretical studies have shown that the presence of males can incur a considerably less than twofold cost on population growth for sexual populations at density-dependent carrying capacity (Doncaster et al., 2000; Kerszberg, 2000; Pound et al., 2002; 2004; Tagg et al., 2005a; b). Small advantages in competition for the sexual population are sufficient to halt the invasion of asexual mutants. The asexual competitors then exert a weaker inhibitory effect on the carrying capacity of the sexual population than they exert on their own carrying capacity through intraspecific competition. The stable outcome is coexistence on a depleted resource base, both locally and regionally (Doncaster et al., 2000; 2003). Under these ecological conditions the sexual population eventually may drive out the asexual competitor by virtue of the longer-term benefits to its inherent genetic variation (Pound et al., 2004). This is a general treatment of ideas present in earlier models of ‘sib-competition’ (the ‘Tangled Bank’ of Bell, 1982; Koella, 1988) and niche differentiation (the ‘Frozen Niche Variation Hypothesis’ of Vrijenhoek, 1979). The recent theory differs from those models by calibrating the competition between sexual and asexual types against competition within each type, using the conceptual framework of classical Lotka–Volterra dynamics (Pound et al., 2002; Tagg et al., 2005a; b). Adler and Levins (1994) predicted that viability selection (= offspring fitness) would become more important than fecundity selection as the level of intraspecific competition increased (e.g. in K-selected populations), and a number of subsequent studies (Mappes et al., 2008) have provided empirical evidence in support of this prediction. Obligate parthenogens may have a higher mortality than their sexual counterparts (Roth, 1974; Corley and Moore, 1999; Kramer and Templeton, 2001). Reduced fitness in terms of fertility and/or offspring survival of asexual populations in comparison to sexual populations (Lamb and Willey 1979; Lynch, 1984; Corley and Moore, 1999; Weinzierl et al., 1999) has been observed.

The concept of density-dependent selection affecting the evolution of demographic traits was apparent to Salisbury (1942), Dobzhansky (1950), and Fisher (1958). According to Bolnick (2004), density and frequency dependence of fitness results in a dynamic landscape: a fitness ‘‘sphagnum bog’’ (Rosenzweig, 1978). A key driver of frequency-dependent fitness is intraspecific competition (Milinski and Parker, 1991; Doebeli and Dieckmann, 2000). The fitness landscape shifts between stabilizing and directional selection at low density to disruptive selection at high density (Svanbäck and Persson, 2009). If the fitness of each phenotype depends on its frequency in the population — that is, if fitness is frequency dependent — then the fitness landscape will be dynamic and the mean phenotype will be kept in a fitness valley, allowing for persistent disruptive selection. Strong competition between similar phenotypes (e.g. parents and offspring or individuals in a clonal population) can disproportionately affect the most abundant (mean) phenotype even though it may be adapted to the most abundant resource. Rarer consumer phenotypes may have fewer resources available, but also have fewer competitors with which to share those resources, so their overall fitness is relatively high. In K-selected environments (MacArthur, 1962), frequency- and density-dependent selection regimes may be highly dynamic thereby favoring sexual reproduction mode. In competitive coevolution, evolution proceeds towards a so-called evolutionary branching point, where selection becomes disruptive and splits the population into two strategies. Coevolution of these strategies eventually leads to the extinction of one of them (Kisdi et al., 2001). Disruptive selection due to frequency-dependent fitness may not only be the causal agent in the evolution of ecological variation and speciation but may equally have been operative in the selection of postreproductive aging and death (Heininger, 2012).

Accordingly, in a broad range of animal and plant taxa differences in yearly survival affect fitness disproportionately more than differences in yearly fecundity, even in many exponentially growing populations, reinforcing the notion that it may be more important to measure survival than fecundity as a single surrogate of population growth (Pfister 1998; Crone, 2001; Burns et al., 2010). Taken together, the twofold costs of reproduction may have to be paid in r-selected habitats, but in K-selected organisms fitness of the progeny plays a larger role than reproductive output. Many mammals, birds, fishes and insects are found living at densities at the carrying capacity of their environments (Sibly et al., 2005; Brook and Bradshaw, 2006). Populations of vertebrate and invertebrate taxa are in general regulated by the production of adult individuals being a decreasing function of population density (Klomp, 1964; Tanner, 1966; Harrison, 1995; Myers et al., 1995; Kuang et al., 2003; Sibly et al., 2005; Bassar et al., 2010) which explains the relative stability of animal populations that do not increase at rates their fertility would allow.

Reviewing parthenogenetic reproduction in the animal kingdom and in the protists, Bell (1982, 1988a) concluded that the best predictor of the initiation of sexual reproduction in species with intermittent mixes is scarcity of resources. For instance, when supplied with sufficient nutrients, yeast cells reproduce vegetatively (asexually), but if starved they undergo the process of meiosis, producing four spores that are functionally equivalent to the gametes involved in sexual reproduction. These spores then fuse in pairs to produce the next generation of vegetative cells (Berry, 1982; Hoekstra, 2005). Under various food levels, the cyclical parthenogenetic rotifer Brachionus plicatilis flexibly changes its reproductive patterns and lifespans. In a food-rich environment, B. plicatilis produces approximately 30 offspring during its lifespan of approximately 10 days. In contrast, with limited reources it suppresses active reproduction and produces less than 10 offspring, while surviving for nearly a month (Yoshinaga et al., 2000; 2003). Offspring quality (starvation resistance) also increases when B. plicatilis reproduces in a food poor environment (Yoshinaga et al., 2001). There are several documented cases of Brachionus strains that have permanently lost the ability to reproduce sexually (Boraas, 1983; Buchner, 1987; Bennett and Boraas, 1988; Fussmann et al., 2003; Stelzer, 2008; 2011; Stelzer et al., 2010). Boraas (1983) found that newly established cultures of Brachionus calycifloru collected from the field produced 40% mictic (sexual) females when induced. After 2–3 months in a chemostat, i.e. with unlimited resources, that percentage was reduced to 0 in similarly inducing environments. This work has been confirmed (Bennet and Boraas, 1989; Fussmann et al., 2003), establishing the costs of sex under unlimited resources (Stelzer, 2011). Similar resource-dependent mechanisms appear to operate within sexually reproducing populations. Population density and fecundity are inversely correlated in a multitude of populations (Clarke, 1955; Christian, 1961; French et al., 1965; Tanner, 1966; Dahlgren, 1979; Clark and Feldman, 1981; Reznick and Endler, 1982; Albon et al., 1983; Andersen and Linnell, 2000; Bonenfant et al., 2009). MacArthur and Wilson (1967) predicted that increased intraspecific competition will select for increased investment of resources in somatic growth, a decreased rate of investment in reproductive tissues, and a tendency to produce fewer, larger and hence more competitive offspring. Supporting this concept, in populations in resource-limited habitats, offspring quantity-quality trade-offs are observed in sexually reproducing plants and animals (Salisbury, 1942; Lack, 1947a; MacArthur and Wilson, 1967; Harper et al., 1970; Smith and Fretwell, 1974; Perrins and Moss, 1975; Gustafsson and Sutherland, 1988; Dijkstra et al., 1990; Hardy et al., 1992; Koskela, 1998; Badyaev and Ghalambor, 2001; Blanckenhorn and Heyland, 2004; Mappes and Koskela, 2004; Charnov and Ernest, 2006; Hagen et al., 2006; Gillespie et al., 2008; Walker et al., 2008; Meij et al., 2009) and also in asexually reproducing organisms (Marshall et al., 2006). Similarly, evolution favors r-selected reproductive strategies at lower population densities and K-selected strategies at the carrying capacity (Sinervo et al., 2000). A test of concept is available within the field conditions experienced by Daphnia pulex. Sexuals might suffer the cost of males in the r-selected environment at the beginning of the season, when resource competition is low; accordingly, reproduction is asexual. But when conditions deteriorate as the population approaches carrying capacity, sexuals seem to be better competitors in spite of male production (Wolinska and Lively, 2008). Ecological studies in aquatic meiobenthic communities including ostracods and copepods revealed that dominance is dependent on tolerance to environmental conditions in the broadest sense, including competitive abilities, and not on the rate of increase, hence not with fertility or generation time (Heip, 1977). In 21 species of woody plants and 45 herbaceous perennials progression to later stage classes, e.g. recruitment and growth, was more important than fecundity in almost all species (Silvertown et al., 1993). In laboratory microcosms where resources declined with time, density of parthenogenetic oribatid mite taxa suffered more from resource limitation than sexual species (Domes et al., 2007). Conversely, sexual processes such as outcrossing are abandoned in favor of inbreeding and parthenogenesis if resources are in ample supply (Hamilton, 1967; 1993; Bell 1982; Knowlton and Jackson, 1993; Scheu and Drossel, 2007; Song et al., 2012). In uniform, resource-rich, agricultural environments parthenogenetic reproduction might be favored over sexual reproduction in pest invertebrates (Corrie et al. 2002, Hoffmann et al., 2008).

Taken together, resources are the limiting factor to population growth. It should not come as a surprise that limited resources also can annihilate the theoretical twofold cost of sex and shape the ecological balance of sexual/asexual reproduction. In populations at or near their habitat carrying capacity, quality rather than quantity, offspring fitness rather than parent fertility is evolutionarily favored. That sexually reproducing organisms that allegedly already pay a twofold cost, further reduce their reproductive output (increasing their putative competitive disadvantage) to produce fitter progeny in more competitive environments argues in favor of the quantity-quality trade-off in reproductive strategies. The twofold cost of sexual reproduction, if at all, may only be realized under r-selected, resource-rich ecological conditions (see chapter 15).

3.1 Germline-derived signals limit the soma’s reproductive potential

The legendary “twofold cost of sex” (Maynard Smith, 1978a) is a logical consequence of classical evolutionary thinking that evolution selects for organisms that maximize their fitness and reproductive output (Darwin 1859; Dawkins 1976; Giesel, 1976; Dennett 1995; Grafen, 1999). As Finch and Holmes (2010) put it: “..under most circumstances selection is expected to eliminate germline mutations that decrease reproductive success or are otherwise detrimental to genetic fitness”. Within this conceptual framework it is counter-intuitive that reproductive activity should be self-limited in sexually reproductive organisms. In fact, the female reproductive aging process is dominated by a gradual decrease in both the quantity and the quality of the oocytes residing within the follicles present in the ovarian cortex (te Velde and Pearson, 2002; Broekmans et al., 2009). Oxidative stress, at least in part mediated by gonadal hormones, is a hallmark of gametogenesis and a variety of other sexual reproduction-related events (Riley and Behrman, 1991; Heininger, 2001; Agarwal et al., 2005; Felty et al., 2005; Fujii et al., 2005; Nedelcu, 2005; Metcalfe and Alonso-Alvarez, 2010; Shkolnik et al., 2011; Rizzo et al., 2012). Oxidative stress is involved in granulosa cell estrogen and progesterone production (LaPolt and Hong, 1995; Yacobi et al., 2007) and estrogen-mediated oocyte maturation (Tarin et al., 1998; Behrman et al., 2001) but, on the other hand, appears to contribute to ovarian senescence (Tarin et al., 1998; Behrman et al., 2001; Miyamoto et al., 2010). Ovulation induces ROS generation in the ovaries that is an essential preovulatory signaling event for proper ovulation (Miyazaki et al., 1991; Shkolnik et al., 2011). Repeated exposure of stored oocytes to ROS at each ovulation results in oxidative damages, measured by 8-hydroxydeoxyguanosine levels in oocytes of the follicle pool (Chao et al., 2005; Miyamoto et al., 2010; Lim and Luderer, 2011), declining antioxidant defenses (Lim and Luderer, 2011), oocyte quality, and size of the ovarian follicle pool (Imai et al., 2012). Proapoptotic Bax is a key component of mitochondrial oxidative stress-mediated apoptotic pathways (Le Bras et al., 2005; Orrenius et al., 2007; Circu and Aw, 2010). Ovarian lifespan can be extended by selectively disrupting Bax function (Perez et al., 1999) suggesting a causal role of oxidative stress in ovarian aging (Miyamoto et al., 2010). The decrease in follicle numbers because of aging reinforces aberrant hormonal regulation as a result of impaired negative feedback mechanisms of the hypothalamic-pituitary-ovarian axis. Decreasing numbers of follicles in the ovaries result in decreasing concentrations of circulating estrogens and inhibins during aging (Broekmans et al., 2009).

In testicular Leydig cells, steroidogenesis is associated with aging of the steroidogenic capacity. In male reproductive senescence, Leydig cells, the testicular cells responsible for testosterone production, become steroidogenically hypofunctional (Zirkin et al., 1997). Oxidative stress, although indispensable for spermatogenesis (Chainy et al., 1997), may play a causal role (Myers and Abney, 1988; Peltola et al., 1996; Zirkin et al., 1997; Chen et al., 2001; Desai et al., 2009) as evidenced by the prevention of Leydig cell aging following the suppression of Leydig cell steroidogenesis (Chen and Zirkin, 1999; Zirkin and Chen, 2000) and culturing Leydig cells with vitamin E, or administering vitamin E to rats (Chen et al., 2005). Cyclooxygenase-2, whose gene expression is up-regulated by many pathological and stress-related factors, including reactive oxygen (Turini and DuBois, 2002) is a putative mediator of the age-related decline in testosterone biosynthesis (Wang et al., 2005). Thus, testosterone-induced oxidative stress in the testes appears to advance the aging of the male reproductive organ (Luo L et al., 2006; Desai et al., 2009).

Reproductive activity not only has progeroid effects at the gonadal level but also degenerates the hypothalamic-pituitary axes. Both pro-oxidant and anti-oxidant effects of estrogens might involve different estrogen receptors that can have either genomic or non-genomic action to manifest further hormonal response (Kumar et al., 2010). The progeroid actions of estrogens appear to depend on neuronal cell types, ratio of different types of estrogen receptors present in a particular cell and context specificity of the estrogen hormone responses (Nilsen, 2008; Kumar et al., 2010). Specifically, estrogens, although neuroprotective at short term (Heininger, 1999), induce degenerative changes in cholinergic basal forebrain neurons upon chronic replacement (Gibbs, 1997). The neurotrophic/neuroinhibitory dualism of estrogens has also been described in hypothalamic slices and cell cultures (Bueno and Pfaff, 1976; Rasmussen et al., 1990); the neurotrophic actions presumably operate in cells regulating reproductive behavior while aging-pacemaker neurons may be targets of the toxic actions. Aging is associated with a disruption of hypothalamic catecholaminergic networks which engender the aging of the somatotropic, thyrotropic and gonadotropic axes (Meites, 1990; Wise et al., 1997). Evidence indicates that estrogens contribute to the derangement of hypothalamic catecholaminergic rhythmicity and function (Wise et al., 1997; Legan and Callahan, 1999). Moreover, gonadal steroids are degenerative and cytotoxic in a variety of hypothalamic nuclei (Schipper et al., 1981; Brawer et al., 1983; 1993; Yang et al., 1993), incite loss of arcuate nucleus synapses (Leedom et al., 1994), actuate the oxidative stress-mediated degeneration of beta-endorphin neurons in the arcuate nucleus (Brawer et al., 1993; Desjardins et al., 1995) and elicit neuronal and glial stress reactions (Day et al., 1993; Seifer et al., 1994; Mydlarski et al., 1995; Krebs et al., 1999). As a result, estrogens exert a self-limiting feedback control of reproductive activity by degenerating the HPG axis. The degeneration of the HPG axis by gonadal hormones not only promotes reproductive aging but has also systemic progeroid effects. Prepubertally ovariectomized mice that received young transplanted ovaries at 11 months showed a 40% increase in life expectancy, relative to intact controls. The 11-month-old recipient females resumed estrus and continued to cycle for several months past the normal point of reproductive senescence (Cargill et al., 2003). Estrogens and other ovarian factor(s) may also play a role in aging-related pituitary changes (Pasqualini et al., 1986; Telford et al., 1987). In a negative feedback loop, the aging-related activation of hypothalamic proinflammatory pathways inhibits the hypothalamic relase of gonadotropin-releasing hormone and accelerates systemic aging (Zhang G et al., 2013). With the loss of negative ovarian feedback at the hypothalamic-pituitary level (MacNaughton et al., 1992; Burger et al., 1995; Landgren et al., 2004), follicle-stimulating hormone levels are elevated (Burger et al., 1995; Klein et al., 1996) which, in turn, accelerates the loss of follicles (McTavish et al., 2007). On the other hand, the increased length of male reproductive phase may, at least in part, depend on the protective action of testosterone against the neurotoxic action of estrogen on distinct hypothalamic neuron populations (Bloch and Gorski, 1988; Yang et al., 1993). As result of these actions, the reproductive phase is self-limited due to the multiple detrimental effects of gonadal steroids and reproductive activity-related oxidative stress on gonadal functioning and the HPG axis

4. The phylogenetic roots of sexual reproduction


In short, it seems certain — and here we must simply grit our teeth and swallow — that it is impossible to understand the function of sex without both a firm grasp of the rival hypotheses, their assumptions, logical development and consequences, and also a reasonably comprehensive knowledge of the nature, occurrence and correlates of sexual systems in nature.

Graham Bell, 1982

Summary

Cellular oxidative metabolism generates reactive oxygen species (ROS). ROS have both beneficial and deleterious effects. Living organisms are heavily dependent on the signaling properties of ROS but, on the other hand, have developed specific mechanisms to contain the production and destructive effects of ROS. Oxidative stress is a state in which the production of ROS exceeds the capacity of the antioxidant defense systems in cells and tissues. Oxidative DNA lesions, if not repaired, possibly lead to mutagenesis and cell death. Environmental stress can alter the balance between stability/repair and mutagenesis. During microbial stress, mutation rates can be increased. Many bacterial species are capable of exchanging genetic material by three parasexual mechanisms: conjugation, transduction, and transformation as a means both to template-guide recombinational DNA repair or increase the genetic repertoire. The heritable incorporation of genetic information is a powerful mechanism of horizontal gene transfer. Microorganisms that are poorly adapted to the current environmental conditions gamble with their genome hoping to boost their fitness. Environmental stress is known to initiate sexual reproduction in a broad range of multicellular eukaryotic species that normally undergo asexual reproduction.

4.1 Oxidative and nitrosative stress, DNA repair and mutagenesis

The appearance of aerobic forms of life was an important step in evolutionary history, since oxygen consumption leads to the production of ten-fold more energy from glucose than does anaerobic metabolism. However, this process imposes constraints on cell viability, because of the generation of highly reactive intermediates, collectively referred to as reactive oxygen species (ROS) during respiration. For the complete reduction of an O2 molecule, a transfer of four electrons is required to produce e.g. two molecules of water or equivalent compounds. The reduction of oxygen by less than four electrons generates unstable ROS. Intracellular ROS originate from multiple sites (Dikalov et al., 2007; Zinkevich and Gutterman, 2011), including the mitochondrial respiratory chain (Skulachev, 1996a; Ježek and Hlavatá, 2005), cytochrome P450 oxygenases (Fleming et al., 2001), nicotinamide adenine dinucleotide (NADPH) oxidase complexes, which include the Nox family (Altenhöfer et al., 2012), xanthine oxidase (McNally et al., 2003), lipid peroxidases (Zhang R et al., 2002), cyclooxygenases (Rowe et al., 1983; Boldyrev et al., 1999), and uncoupled nitric oxide (NO) synthase (NOS) (Vasquez-Vivar et al., 1998). Most of these enzymes act in the mitochondria, which is the main source of oxidative stress. There is a functional connectivity between intracellular sites of ROS production. This cross talk has been termed ROS-induced ROS release and is supported by a variety of observations showing that ROS-induced ROS release is a common mechanism for ROS amplification and regional ROS generation (Zinkevich and Gutterman, 2011). For instance, a reciprocal activation of ROS-induced ROS release between mitochondria and NADPH oxidase can produce a feed-forward acceleration of ROS generation (Kimura et al., 2005; Lee SB et al., 2006; Hawkins BJ et al., 2007; Wenzel et al., 2008). NADPH oxidases are the only known enzyme family with the sole function to produce ROS (Altenhöfer et al., 2012) and induce the release of ROS produced by other enzymatic systems (Brandes and Kreuzer, 2005; Griendling, 2004) like xanthine oxidase (McNally et al., 2003). Approx. 90% of ROS arise in the mitochondria as byproducts of aerobic metabolism. Complex I (NADH dehydrogenase) and particularly complex III (cytochrome bc1) are major sources of ROS (Boveris, 1977; Finkel and Holbrook, 2000; Ježek and Hlavatá, 2005). During mitochondrial respiration, 0.12 to 2% (in some reports even up to 5%) of the oxygen undergoes single electron transfer in vitro generating the superoxide anion radical (O2•-) (Boveris and Chance, 1973; Boveris, 1977; Chance et al., 1979; Turrens, 2003; Harper et al., 2004; Orrenius et al., 2007; Murphy, 2009). However, these values cannot be extrapolated to the in vivo situation where mitochondrial O2•- production will be much lower (Murphy, 2009). This molecule shows limited reactivity but is converted to hydrogen peroxide (H2O2) by superoxide dismutase. H2O2 is not a free radical but in the presence of transition metals, such as iron and copper, is reduced to highly reactive hydroxyl radicals (OH) (Loft and Poulsen, 1996). These ROS are potent oxidants of lipids, proteins, and nucleic acids (Halliwell and Gutteridge, 1984; Meneghini, 1988; Buettner, 1993). Living organisms have developed specific mechanisms to prevent the production and effects of ROS (Halliwell and Gutteridge, 1999). Oxidative stress is a state in which the production of ROS exceeds the capacity of the antioxidant defense system in cells and tissues (Sies, 1985; 1986; 1991; Davies, 1995; Finkel and Holbrook, 2000; Mittler, 2002; Slupphaug et al., 2003; Monaghan et al., 2009; Lushchak, 2011). The ROS-dependent inactivation of antioxidant enzymes can be expected to aggravate the dysregulation (Kono and Fridovich, 1982; Tabatabaie and Floyd, 1994; Kwon et al., 1998; Lee SM et al., 2001; Macmillan-Crow and Cruthirds, 2001; Alvarez et al., 2004). Nitric oxide (NO) is chemically rather unreactive toward most bio-organic compounds but unstable in the presence of O2 and ROS and will rapidly autoxidize to yield a variety of nitrogen oxide intermediates, some of which are potent nitrosating agents (Marletta, 1988; Williams, 1988; Bartsh et al., 1989; Huie and Padmaja, 1993; Packer et al., 1996; Szabo et al., 2007) that will produce mutagenic and carcinogenic nitrosamines (Wink et al., 1991; Laval and Wink, 1994; Wink and Laval, 1994; Tamir et al., 1996; Burney et al., 1999; Grisham et al., 2000). Moreover, mitochondria can act as a “redox signaling box,” converting an NO signal into an ROS signal (Brookes and Darley-Usmar, 2002).

Initially, scientific interest was almost exclusively focused on investigating the destructive potential of oxidative stress. Increasingly, however, protein modification by oxidation/reduction is emerging as a critical mechanism for the modulation of protein activities, and the protective and creative actions of ROS signal transduction and their key role in cellular homeostasis have been recognized (Remacle et al., 1995; Lander, 1997; Finkel, 1998; Gamaley and Klyubin, 1999; Rhee, 1999; Thannickal and Fanburg, 2000; Heininger, 2001; Dröge, 2002; Apel and Hirt, 2004; Forman et al., 2004; 2010; Pitzschke et al., 2006; Rhee, 2006; Veal et al., 2007; Janssen-Heininger et al., 2008; Lane, 2011; Shapiguzov et al., 2012; Wong, 2012; Xie and Roy, 2012; Zarse et al., 2012). Because of these opposing and dose-dependent functions, the cellular level of ROS is likely to be subjected to tight regulation via processes involved in production, distribution and removal (Bienert et al., 2006). Compared with other ROS, H2O2 is a relatively long-lived molecule (1 ms) that is able to diffuse freely across cell membranes (Chance et al., 1979; Gutteridge, 1994; Bienert et al. 2006; but see Miller et al., 2010), making it the most prevalent ROS form (Mathai and Sitaramam, 1994). Due to their relatively low reagibility and long-range effects, H2O2 and superoxide qualify as messengers via (i) reversible modification of target protein molecules and (ii) changes of intracellular redox state (Finkel and Holbrook, 2000; Thannickal and Fanburg, 2000; Sauer et al., 2001; Neill et al., 2002; Van Breusegem and Dat, 2006; Bartosz, 2009). However, due to high concentrations of mitochondrial SOD, the intramitochondrial concentrations of O2•- are maintained at very low steady-state levels (Tyler, 1975) and mitochondria-generated O2•- is unlikely to escape into the cytoplasm. Serving various functions, cellular organelles maintain redox control unique to each compartment and may also influence membrane permeability to H2O2 (Makino et al., 2004; Hansen et al., 2006; Malinouski et al., 2011). Plasma membrane was also shown to change H2O2 permeability under certain conditions (Antunes and Cadenas, 2000; Seaver and Imlay, 2001; Sousa-Lopes et al., 2004; Jang et al., 2012). Redox-sensitive cysteine residues are present in the interaction domains of many proteins. Enzymes that possess a highly reactive cysteine in either the catalytic centre or a regulatory site play a key role in the transmission of ROS signals (Barford, 2004; Forman et al., 2004). There are examples in all of the major categories of transcription factors, including basic region, leucine zipper, helix-loop-helix, and zinc finger (Webster et al., 2001). Proteins vulnerable to oxidation that can be reversed by thiol donors such as glutathione, glutaredoxin and thioredoxin, include transcription factors (such as the nuclear factor κ-B, activator-protein 1, hypoxia inducible factor, p53 and the p21 Ras family of protooncogenes), protein tyrosine kinase (PTKs), such as Src family kinases and some receptor tyrosine kinases (RTKs) and finally protein tyrosine phosphatases (PTPs), such as PTP1B, Low Molecular Weight-PTP, Src-homology-2 domain-PTP and PTEN (Lee et al., 1998; Chiarugi et al., 2001; Mahadev et al., 2001; Meng et al., 2002; 2004; Kwon et al., 2004; Chiarugi and Fiaschi, 2007). The activation of redox-regulated transcription factors such as NFκ-B and HSF1 may be selective for the type of oxidant, H2O2 being effective while superoxide is not (Schreck et al., 1992; Jacquier-Sarlin et al., 1995). Tyrosine phosphorylation/dephosphorylation in response to extracellular stimuli has been characterized most thoroughly (Cho et al., 2004; Fujii and Tsunoda, 2011). However, whereas both superoxide and hydrogen peroxide have low oxidative toxicity themselves and do not interact at all with DNA bases (Halliwell and Aruoma, 1991; Dizdaroglu, 1993), they are believed to elicit their toxicity to DNA by conversion to hydroxyl radicals mediated by transition metal ions (e.g. iron and copper) through Haber-Weiss and Fenton reactions (von Sonntag, 1987; Steenken, 1989). Superoxide and hydrogen peroxide are both components of the net Haber-Weiss reaction by oxidation of superoxide anion radical that readily produces hydroxyl radicals.

O2•- + H2O2 OH + OH- + O2

The main source of biological hydroxyl radicals is the metal-catalyzed breakdown of hydrogen peroxide by the Fenton reaction, one-half of the net Haber-Weiss reaction:

Fe2+ + H2O2 Fe3+ + OH + OH-

Both Cu+ and Fe2+ are capable of reacting with H2O2 to form OH in vitro. However, Fe2+ is evidently the coreactant in vivo, since cell-permeable iron chelators protect DNA from exogenous H2O2 (Imlay and Linn, 1988). The hydroxyl radical is highly reactive and is responsible for most of the damage incurred to biological macromolecules by ROS. Because of their short half life (less than 1 ns), hydroxyl radicals attack molecules very close to their site of formation (von Sonntag, 1987a; Halliwell and Gutteridge, 1999; Valko et al., 2004) and their very high reactivity makes them virtually impossible to scavenge (Halliwell and Gutteridge, 1999; Halliwell, 2007).

OH + DNA → H2O + damage

It has been demonstrated that the thiol groups of the cysteines involved in the zinc-finger motif as part of the metal-binding, DNA-intercalating fingers (Kadonaga et al., 1987; Desjarlais and Berg, 1992) confer gene regulation by ROS (Cimino et al., 1997). These zinc finger structures are potentially very sensitive redox targets in DNA binding motifs of many DNA transcription factors and DNA repair enzymes (Berg, 1992; Rhodes and Klug, 1993; Wu et al., 1996; Webster et al., 2001; Wilcox et al., 2001). Specifically, when redox iron displaces zinc ions in zinc-finger structures of DNA repair enzymes, it generates more aggressive free radicals.

Over 100 oxidative DNA adducts have been identified thus far; however, whether each of these adducts are produced in measurable amounts in vivo to be biologically relevant is unclear (von Sonntag, 1987b; Dizdaroglu, 1992; Demple and Harrison, 1994; Cadetet al., 2003). Among the oxidative DNA lesions, one of the major classes of DNA damage leads to at least 20 modifications in purine and pyrimidine bases, together with oligonucleotide strand breaks, DNA-protein cross-links and abasic sites, possibly leading to mutagenesis and cell death (Aitken et al., 1998a; Lopes et al., 1998; Halliwell and Gutteridge, 1999; Kemal Duru et al., 2000; Dizdaroglu et al., 2002; Sawyer et al., 2003; Sedelnikova et al., 2010). For instance, the number of oxidative lesions to DNA per cell per day, even without excessive exposure to DNA damaging agents, is estimated to be about 100,000 in rats and greater than 20,000 in humans (Fraga et al., 1990; Ames et al., 1993; Lindahl, 1993; Lindahl and Barnes, 2000; De Bont and van Larebeke, 2004) and is thought to contribute to carcinogenesis, aging and tissue damage following ischemia (Jackson and Loeb, 2001; Ma et al., 2008; Klaunig et al., 2010; Ralph et al., 2010; Sedelnikova et al., 2010). An analysis of the mutation spectrum suggests that there are at least four different pathways by which oxidative DNA damage leads to mutations (Feig et al., 1994): (i) chemical modification of nucleotide moieties in DNA causing an alteration in their hydrogen bonding or "coding" specificity; (ii) damage-mediated exacerbation of polymerase-specific hot spots; (iii) damage-induced conformational change in the DNA template that prevents accurate replication by DNA polymerases (Hsu et al., 2004); (iv) induction of a DNA polymerase conformation that is error-prone. To overcome the potential problems to the functionality of DNA, essentially all organisms use a battery of DNA repair enzymes that identify the damaged site, and remove it with restoration of the original DNA sequence (Friedberg et al., 2006). Several DNA repair pathways are involved in maintaining cell genomic stability; these include direct repair (DR), base-excision repair (BER), nucleotide-excision repair (NER), mismatch repair (MMR), DNA double-strand break repair (DSBR), and post replication repair (PRR). More than 150 proteins are involved (Miles and Sancar, 1989; Luo M et al., 2010). Eukaryotic cells repair DNA double-strand breaks (DSBs) by at least two pathways, homologous recombination (HR) and non-homologous end-joining (NHEJ) (Shinohara and Ogawa, 1995). The best-studied transcriptional response to DNA damage is the bacterial SOS response (Friedberg et al., 1995; Walker, 1996; Storz and Hengge-Aronis, 2000; van der Veen et al., 2010). The SOS response involves two kinds of repair systems (Shapiro, 2011):

• A precise repair process that removes the UV-damaged DNA and does not introduce mutations. This “error-free” process operates very much like the mismatch repair proofreading system, except that the sensor protein recognizes the characteristic chemistry of UV damage in DNA rather than helix distortions. It is called excision repair because the result of damage sensing, as in mismatch correction, leads to excision of a section of the damaged DNA strand (Smith, 1978; Van Houten, 1990; Petit and Sancar, 1999).

• A mutagenic repair process that involves the synthesis of specialized “error-prone” DNA polymerases, which can replicate DNA that carries unrepaired damage. Without these specialized polymerases, mutations do not occur in response to DNA damage. Instead, the cells or molecules that cannot remove or replicate past the damage are simply doomed and produce no mutant progeny (Pham et al., 2001; Goodman, 2002; Wagner et al., 2002; Broyde et al., 2008).

A low concentration of H2O2 results in SOS gene induction in wild-type cells (Imlay and Linn, 1987; Goerlich et al., 1989). The regulon orchestrates cellular survival responses to a variety of stressors, mediates growth arrest, mutagenesis and resistance to DNA damage (Heininger, 2001; Aertsen and Michiels, 2005). Environmental stress can alter the balance between stability/repair and mutagenesis/exploration/bet-hedging (Foster, 2005). Recombination frequency is markedly elevated by cellular stressors such as drugs, ionizing radiation, parasites, and heat shock, known activators of oxidative stress (Lebel et al., 1993; Lucht et al., 2002; Winn et al., 2003; Defoort et al., 2006). Stress-induced mutation mechanisms differ from those that produce classical spontaneous mutations, which occur with a definable relationship to cell generations in proliferating cells (Luria and Delbrück, 1943; Lea and Coulson, 1949). In an E. coli model, stationary-phase sigma factor RpoS (which encodes the stress response σS transcription factor)-controlled switch from high-fidelity to mutagenic double-strand break repair activates stress-induced mutagenesis during stationary phase/starvation (Ponder et al., 2005). B. subtilis has similar stationary phase/starvation-dependent mutagenic programs (Robleto et al., 2007). Mutants that lack components of the SOS system are more sensitive to a variety of damaging agents (Imlay and Linn, 1987; Yonei et al., 1987; Escarceller et al., 1994; van der Veen et al., 2010). Higher levels of oxidative stress obviously favor mutagenesis (Greenberg and Demple, 1988; DeRose and Claycamp, 1991; McBride et al., 1991; Escarceller et al., 1994; Kato T et al., 1994; Blanco et al., 1995; Touati et al., 1995; Urios et al., 1995; Wang and Humayun, 1996; MacPhee, 1999; Ruiz-Laguna et al., 2000; Bjedov et al., 2003). Importantly, the repair/mutagenesis balance is modulated by interactions with heat shock proteins (Liu and Tessman, 1990a; b; Donnelly and Walker, 1989; 1992; Petit et al., 1994) which, in dependence of the cellular energy level may provide the regulatory feedback with cellular metabolic/oxidative homeostasis and its derangement during stress (MacPhee, 1985; 1994, 1996; Seetharam and Seidman, 1992; Keszenman et al., 2000). Oxidative stress is the final common pathway of responses to a multitude of biotic and abiotic stressors, including psychosocial stress (Lindquist, 1986; Sanchez et al., 1992; Finkel and Holbrook, 2000; Heininger, 2001; Mittler, 2002; Mikkelsen and Wardman, 2003; Sørensen et al., 2003; Apel and Hirt, 2004; Ardanaz and Pagano, 2006; Rollo, 2007; Miller et al., 2008; Slos and Stoks, 2008; Jaspers and Kangasjärvi, 2010; Steinberg, 2012; Choudhury et al., 2013).

4.2 Stress and microbial transformation

Bacteria multiply asexually by binary fission. Hence, bacteria are essentially clonal organisms. Under stressful conditions, mechanisms that increase genetic variation can bestow a selective advantage. Bacteria have several stress responses that provide ways in which mutation rates can be increased (Bjedov et al., 2003; Hastings et al., 2004; Miller, 2005; Galhardo et al., 2007). These include the SOS response, the general stress response, the heat-shock response, and the stringent response, all of which impact the regulation of error-prone polymerases. Two classes of genes are known to accelerate mutation and/or recombination rates in bacterial populations: stress-inducible wild-type genes, usually part of the SOS regulon, and genes whose functional loss, or downregulation, increases the rate of genetic variability (mutator and/or hyper-rec mutants) (Radman et al., 2000). Stress-induced mutation appears to be a process by which cells can respond to selective pressure specifically by producing mutations (Foster, 2005; Miller, 2005). Many bacterial species are capable of exchanging genetic material by three parasexual mechanisms: conjugation, transduction, and transformation. Parasexuality is any process in which more than one parent participates, without meiosis and fertilization, which gives as a result a new cell (Heitman, 2006). The induction of genetic competence, the ability of cells to bind to and to take up exogenous DNA for food (Redfield, 1993b), evolved as a strategy used by bacteria to increase their genetic repertoire. Bacterial transformation, in which cells take up and recombine free strands of DNA, is a general response of bacteria to stress (Storz and Hengge-Aronis, 2000; Claverys et al., 2006; Singh et al., 2008). The heritable incorporation of this genetic information is a powerful mechanism of horizontal gene transfer. Genetic transformation is a widespread both natural and inducible feature of bacteria (Carlson et al., 1983; Dubnau 1991; Palmen et al., 1993; Lorenz and Wackernagel, 1994; Ogunseitan, 1995; Baur et al., 1996; Tortosa and Dubnau, 1999; Birdsell and Wills, 2003; Claverys et al., 2006; Singh et al., 2008). Competence for genetic transformation is under control of the SOS regulon, the bacterial master stress regulator (Yasbin et al., 1991; 1992; Cheo et al., 1993; Storz and Hengge-Aronis, 2000; Erill et al., 2007; Chiang and Schellhorn, 2010). As shown in computer simulations, bacterial transformation, which has been dubbed ‘sex with dead cells’ (Redfield, 1988), can increase the fitness of the recombinant progeny at equilibrium although the DNA taken up may be from cells killed by selection against mutations and may be of inferior genetic quality (Redfield, 1988; Hoelzer and Michod, 1991). The competence-regulatory apparatus senses and interprets environmental information and via an elaborate signal transduction system passes this information to the competence-specific transcriptional machinery. Remarkably, these regulatory pathways play also a role in the expression of other post-exponential phenomena such as motility, sporulation, autolysis (Weinrauch et al., 1990, Dubnau, 1991; Grossman, 1995; Solomon and Grossman, 1996; Azoulay-Dupuis et al., 1998), stress resistance (Yasbin et al., 1991; 1992), recombination (Lovett et al., 1989) and “adaptive” mutagenesis (Radicella et al., 1995). Many stress responses are only pursued by discrete subpopulations of cells in a given bacterial community and are regulated as mutually exclusive events coordinated by integrated molecular switches and threshold phenomena (Msadek, 1999; Heininger, 2001; Tenaillon et al., 2001; Booth, 2002; Aertsen and Michiels, 2005; Foster, 2005; 2007; Fehér et al., 2012). However, conflict between these adaptive strategies may arise (Tenaillon et al., 2000; 2001). For example, stationary phase adaptive responses engage a highly interconnected network of signal transduction pathways which regulate motility (to exploit other carbon sources, flight), sporulation (differentiation to highly resistant, metabolically dormant spores), antibiotic production (cytocide), cannibalism, competence for DNA uptake (mating responses), stress resistance and mutagenesis (Love et al., 1985; Yasbin et al., 1991; 1992; Nunoshiba, 1996; Loomis et al., 1998; Perego, 1998; Msadek, 1999; Heininger, 2012). A stress-induced, fratricide-associated predatory mechanism may dramatically increase the efficiency of lateral gene transfer in Streptococcus pneumoniae and related commensal species (Prudhomme et al., 2006; Claverys and Håvarstein, 2007; Johnsborg et al., 2008; Johnsborg and Håvarstein, 2009). In a stressed B. subtilis colony only a minority of approximately 10-20% cells becomes competent (Somma and Polsinelli, 1970; Dubnau, 1982). Importantly, in subsequent starvation episodes, organisms carrying stress-induced genomic rearrangements exhibited a fitness advantage relative to the parental strain (Coyle and Kroll, 2008).

Sexual reproduction and eukaryotes probably evolved together (Cavalier-Smith, 2002), about 2.0–3.5 billion years ago (Miyamoto and Fitch, 1996; Gu, 1997). The frequency of sexual reproduction generally depends on the condition of an individual, called fitness-associated sex, a pattern found broadly across both facultatively sexual prokaryotes and eukaryotes (Bell, 1982; Hadany and Beker, 2003b; Hadany and Otto, 2007; 2009). Thus, organisms that are poorly adapted to the current environmental conditions, as part of a variety of stress-related responses, gamble with their genome hoping to boost their fitness (Dubnau and Losick, 2006; Veening et al., 2008a). Mutational alteration of genotype enables a sampling of the ‘sequence space’ of the organism and increases the chances of natural selection acting on better adapted alleles (Sundin and Weigand, 2007). Individuals that are starved or deprived of other vital resources undergo sex in a wide variety of microorganisms, including bacteria (Dubnau, 1991; Redfield, 1888; 1993b; Jarmer et al., 2002; Foster, 2005), yeast (Kassir et al., 1988; Mochizuki and Yamamoto, 1992; Mai and Breeden, 2000; Abdullah and Borts, 2001; Davis and Smith, 2001; van Werven and Amon, 2011), and Chlamydomonas reinhardtii (Harris EH, 1989; Quarmby, 1994; Goodenough et al., 2007). In a variety of unicellular eukaryotic organisms e.g. Dictyostelium, asexual and mating behavior are alternative responses to metabolic stress (Saga and Yanagisawa, 1982; Cornillon et al., 1994). Evidence suggests, however, that mating responses are restricted to less harsh environmental conditions. Sexual reproduction may also be induced by crowding as social stressor (Harvell and Grosberg, 1988; Harris EH, 1989; Gilbert, 2003). In V. carteri a brief exposure to elevated temperatures generates egg-bearing sexual daughters (Kirk and Kirk, 1986). The same set of genes triggered by the sexual pheromone was also inducible in V. carteri by wounding (Amon et al., 1998). Mating behavior is favored by unfavorable and fluctuating environmental conditions (Frank and Swingland, 1988, Hoffman and Parsons, 1991; Nelson, 1996; Robson et al., 1999; Becks and Agrawal, 2010; 2012; Schoustra et al., 2010). In soil microfungi, sex is more common under stressful environmental conditions associated with drought severity and high salinity (Grishkan et al., 2003).

Cells with DNA damage have also been shown to undergo sex in viruses (Bernstein, 1987), bacteria (Wojciechowski et al., 1989), and yeast (Bernstein and Johns, 1989). In plants and benthic marine animals there is relatively greater allocation to sexual than to vegetative reproduction at high density (competitive stress) (Harvell and Grosberg, 1988; van Kleunen et al., 2001; van Kleunen and Fischer, 2003). The external signals may include pheromones that act as quorum-sensing signals and induce competence under a wide variety of conditions (Morrison and Baker, 1979; Zhang et al., 1993; Gomer, 1994; Hwang et al., 1994; Morrison, 1997, Tortosa and Dubnau 1999). Intriguingly, these pheromones and their transduction pathways may function as sexual isolation mechanism (Tortosa and Dubnau, 1999). These findings link the microbial mating behavior with population density as representing the cost of finding a mating partner (Bernstein et al., 1985a; b). The pheromone-induced conjugational transfer of plasmids has an immense importance for the acquisition of virulence and antibiotic resistence (Wirth, 1994, Dunny et al., 1995, Zatyka and Thomas, 1998). The finding that plasmids (Goodman et al., 1993) or chromosomal DNA from other bacterial species can also be taken up and integrated, supports the adaptive concept but is not compatible with the concept that the ingested DNA may serve as template for recombinational repair (Lorenz and Wackernagel, 1994; Nielsen, 1998). For instance, ‘adaptive’ mutation was substantially mitigated when conjugational behavior was inhibited in starving bacteria (Radicella et al., 1995).

Fungi and yeast, facultatively sexual/asexual eukaryotes, exhibit sexual reproduction under various stressors such as nutritional and nitrogen starvation and following oxidative stress (Bernstein and Johns, 1989; Mochizuki and Yamamoto, 1992; Nelson, 1996). This mating behavior may also be induced by pheromones (Konopka and Fields, 1992; Kurjan, 1992; Bardwell et al., 1994; Spellig et al., 1994) which also require stress factors to induce their synthesis (Moore and Edman, 1993; Jee and Ko, 1998). Remarkably, pheromones induce growth arrest in G1 (Kurjan, 1992; Bardwell et al., 1994; Oehlen and Cross, 1994) the starting point of another stress response, differentiation (Heininger, 2001; 2012). Entry into stationary phase, differentiation, stress and conjugation responses are regulated in yeast by the same signaling pathways and transcription factors (Ammerer, 1994; Takeda et al., 1995; Wu and McLeod, 1995; Kato et al., 1996; Shiozaki and Russell, 1996; Madhani and Fink, 1998; Widmann et al., 1999). These pathways also implicate the expression of hsp (Danjoh and Fujiyama, 1999). That this phylogenetically ancient mechanism is highly conserved throughout eukaryotes is indicated by a striking conservation from yeasts to humans of a gene controlling stress-induced sexual development in yeast (Okazaki et al., 1998). The spontaneous mutation rate for some, though not all, mutations in cells of Saccharomyces cerevisiae undergoing meiosis is 3 to 30 times that of cells reproducing mitotically (Magni and von Borstell, 1962; Magni, 1963; Lawrence, 1982).

Another feature of conjugational behavior is the formation of polyploid giant cells which can be formed by algae, yeasts and other protozoans under various stress conditions (Rink and Partke, 1975; Urushihara, 1992; Mares et al., 1993) again facultatively mediated by pheromones (Urushihara, 1992). Even bacteria may be able to exhibit this type of conjugational behavior in stationary phase (Akerlund et al., 1995).

Three non-mutually exclusive models can account for the evolution of DNA uptake systems (Mehr and Seifert, 1998; Dubnau, 1999):

• DNA for genetic diversity and innovation — the acquisition of potential useful genetic information, such as novel metabolic functions, virulence traits or antibiotic resistance (Dubnau, 1999; Ochman et al., 2000).

• DNA repair — environmental DNA from closely related bacteria might serve as templates for the repair of DNA damage (Bernstein et al., 1981; 1985a; b; Michod et al., 1988; Solomon and Grossman, 1996).

• DNA as food — DNA can be used as a source of carbon, nitrogen and phosphorous (Redfield, 1993b; 2001; Finkel and Kolter, 2001; Palchevskiy and Finkel, 2006).

4.3 Stress and multicellular reproductive behavior

During metazoan phylogeny many aspects of reproductive physiology including the biosynthesis, structure and function of steroid hormones show a remarkable degree of conservation arguing for the maintenance of shared origins and rationales (Dawson, 1998). In simple metazoans, the impact of environmental challenges on the preference of sexual reproduction can still be observed (Curtis, 1902; Hyman, 1939; Burnett and Diehl, 1963; Bell, 1982; Bell and Wolfe, 1985; Harvell and Grosberg, 1988; Schierwater and Hadrys, 1998). Volvocine algae switch from their principal, asexual mode of reproduction to a sexual mode when life-threatening conditions approach such as heat shock (Kirk and Kirk, 1986; Kirk, 1998; Harris et al., 2009), indicating drying up of the pond. The sexual cycle produces heavily walled, dormant zygotes (zygospores) that can resist tough conditions like drought, heat, and cold for a long period of time (Hallmann, 2011). Individuals that are starved also undergo sex in cladocera (Stuart et al., 1931; D’Abramo, 1980) and daphnia (together with inductive photoperiod and crowding) (Kleiven et al., 1992). Competitive stress also favors sexual reproduction in Daphnia and a rotifer (Kleiven et al., 1992; Berg et al., 2001; Gilbert, 2004). Sex in the nematode, Strongyloides ratti, is induced in response to a host immune response (Gemmill et al., 1997; West et al., 2001). Plants and sessile aquatic animals often exhibit labile sex expression but turn to sexual reproduction under moderate environmental challenges (Harvell and Grosberg, 1988; Kimmerer, 1991; Romme et al., 1997; Korpelainen, 1998; van Kleunen et al., 2001). Likewise, biotic and abiotic stress increase plant homologous recombination (Filkowski et al., 2004; Molinier et al., 2006; Boyko et al., 2010).

Environmental stress is known to initiate sexual reproduction in a broad range of species that normally undergo asexual reproduction (Bell, 1982; Harris PD, 1989; Dubnau, 1991; Kleiven et al., 1992; Gemmill et al., 1997; Dacks and Roger, 1999; Mai and Breeden, 2000). Some enchytraeid species (pot worms) reproduce asexually by fragmentation and subsequent regeneration (Bell 1959; Christensen 1959; Bouguenec and Giani 1989; Dózsa-Farkas, 1995; Schmelz et al., 2000; Yoshida-Noro and Tochinai, 2010) but reproduce sexually under environmental stressors such as starvation (Christensen 1959; Dózsa-Farkas 1996; Myohara et al., 1999; Tadokoro et al., 2006). Likewise, facultative outcrossing may enhance the adaptive response of highly selfing populations in the face of environmental and genetic stress (Morran et al., 2009a; b).

In multicellular unitary organisms with their increasing independence from environmental changes, a special variation of the environmental challenge leitmotif is the change to and maintenance of sexual reproduction in hosts to counter the pressure of parasitic coevolution (the Red Queen hypothesis) (Hamilton, 1980; Hamilton and Zuk, 1982; Hamilton et al., 1990; Howard and Lively, 1994; Møller and Saino, 1994, Able, 1996; John, 1997; Ooi and Yahara, 1999; Martins, 2000; Salathé et al., 2008; Lively, 2010). The coevolutionary models have been extended to interactions with other species in the biotic space (Otto and Nuismer, 2004). Climatic change has been suggested to be a comodulating factor to the host-parasite coevolution (Atkinson, 1991). The same appears to be true for parasite reproduction. A parasitic nematode reproduces sexually in immune-competent hosts but may propagate clonally in the less challenging environment of immune-deficient hosts (Gemmill et al., 1997). In fact, random mating is the by far favored reproduction type of a pathogenic fungus in a natural immune-competent environment (Chen and McDonald, 1996). As further support of the coevolution concept, genetic diversity of viruses is substantially higher in sexual than asexual hosts (Ooi and Yahara, 1999). The modulation of sexual reproduction appears to be even exploited as strategy in the parasite-host arm race. Bacteria induce parthenogenesis in infected female wasps (Huigens et al., 2000) thus putatively reducing the hosts’ evolutionary adaptability and immune competence. Both circumstantial evidence (Clark, 1996) and theoretical calculations (Cui et al., 2000) suggest that aging may contribute to the maintenance of sexual reproduction. Conversely, asymmetric, both asexual and sexual, reproduction are coselected with senescence and death (Heininger, 2002; 2012).

Oxidative stress is the final common pathway of responses to a variety of biotic and abiotic stressors (Lindquist, 1986; Sanchez et al., 1992; Finkel and Holbrook, 2000; Heininger, 2001; Mittler, 2002; Mikkelsen and Wardman, 2003; Sørensen et al., 2003; Apel and Hirt, 2004; Ardanaz and Pagano, 2006; Rollo, 2007; Miller et al., 2008; Slos and Stoks, 2008; Jaspers and Kangasjärvi, 2010; Steinberg, 2012; Choudhury et al., 2013). And in this capacity, oxidative stress is a general inducer of sexual reproductive activity (Bernstein and Johns, 1989; Heininger, 2001; Nedelcu and Michod, 2003; Nedelcu et al., 2004; Nedelcu, 2005; McInnis et al., 2006). In fungi, deletion or mutagenesis of NADPH oxidases that are used deliberately to produce ROS, specifically block differentiation of sexual fruit bodies, without affecting asexual development (Lara-Ortíz et al., 2003; Malagnac et al., 2004; Takemoto et al., 2007).

5. Mutagenesis and fitness


Summary

The Modern Synthesis posited that (1) mutations occur independently of the environment, (2) mutations are due to replication errors, and (3) mutation rates are constant. Since the vast majority of new mutations are likely to be neutral or deleterious for fitness, organisms should evolve as low mutation rates as possible. Contrary to this prediction, there are situations in nature, e.g. in stressful conditions that by definition are associated with impaired fitness, in which being a mutator confers a selective advantage (condition-dependent mutagenesis). The mutation rate of an organism is subject to heritable variation, and therefore evolves. Mathematical models suggested that mutation rates adapt up (or down) as the evolutionary demands for novelty in variable environments (genetic innovation) or memory in stable environments (genetic conservation) increase. Nature selected for those organisms with a mutation rate that compromises between adaptability and adaptedness. Stressful conditions can lead to a genetically controlled increase in bacterial and eukaryote mutagenesis that is genome-wide. The evolutionary fate of mutators, however, is inextricably linked to their effective population size: only with a large population size chances are substantial that some individuals may create beneficial mutations that confer a competitive advantage to cells, allowing them to take over the population. Based on the nearly universal inverse relationship between body mass/size and population size and density, increased multicellularity is generally associated with a reduction in effective population size, and this in turn reduces the efficiency of natural selection. Given the deleterious effects of most mutations, theory predicts that increasing the mutation rate should generate progressively larger reductions in fitness, assuming that the mutations accumulate in the genome and that the effects of the mutations are additive or act synergistically. However, efficiency of artificial selection can be enhanced due to the increased genetic variation induced by mutagenesis. An efficient selection regime that is able to pick the “pearl(s)” of beneficial or at least neutral mutations out of the “pebbles” of deleterious ones may fundamentally change the course of evolution.

5.1 Are mutations genetic “accidents”?

In this paper, mutagenesis and mutation, when not specified otherwise, will refer in the broadest sense to any permanent, and consequently heritable, change of DNA sequence including transposable elements activity and recombination. Homologous recombination by itself does not generate sequence polymorphisms. However, recombination does prevent the reduction in variability caused by selective sweeps and sequential bottlenecks, thus increasing the polymorphism in a population (Suerbaum et al., 1998).

The Modern Synthesis posited that (1) mutations occur independently of the environment, (2) mutations are due to replication errors, and (3) mutation rates are constant (Lenski and Mittler, 1993; Brisson, 2003). Mutation is a double-edged sword. On one side, it is the ultimate source of genetic variation and the basis of evolution, the material upon which natural selection works (Punnett, 1911, p. 139; Mayr, 2001). Yet, the vast majority of new mutations are likely to be neutral or deleterious for fitness (Bridges, 1919; Fisher, 1930; 1958; Falconer and Mackay, 1996; Kibota and Lynch, 1996; Elena and Moya, 1999; Lynch et al., 1999; Wloch et al., 2001a; Keightley and Lynch, 2003; Sanjuán et al., 2004a; Eyre-Walker and Keightley, 2007; Sawyer et al., 2007; Soskine and Tawfik, 2010; King, 2012a). For example, for Escherichia coli K-12, the rate of deleterious mutations per genome per replication is, at least, 2–8×10-4 (Kibota and Lynch, 1996; Boe et al., 2000), while that of beneficial mutations is, at least, 2×10-9 (Imhof and Schlotterer, 2001). As Calvin Bridges, one of the pioneers of Drosophila genetics, put it (1919): “Any organism as it now exists must be regarded as a very complex physicochemical machine with delicate adjustments of part to part. Any haphazard change made in this mechanism would almost certainly result in a decrease of efficiency… Only an extremely small proportion of mutations may be expected to improve a part or the interrelation of parts in such a way that the fitness of the whole organism for its available environments is increased.” It is generally assumed that beneficial effects are ~1,000-fold less common that neutral and deleterious ones (Miralles et al., 1999; Keightley and Lynch, 2003; Orr, 2003), although large differences exist between taxa (Eyre-Walker and Keightley, 2007). Applying an extended McDonald-Kreitman test (McDonald and Kreitman, 1991; Charlesworth, 1994; Fay et al., 2001; Smith and Eyre-Walker, 2002; Loewe and Charlesworth, 2006) to polymorphism data from D. melanogaster, the fraction of deleterious newly arising mutations was estimated to be ~94% at amino acid sites, ~81% in untranslated regions, ~56% in introns, and ~61% in intergenic regions (Andolfatto, 2005). Any variant that has increased mutation rates is expected to have reduced fitness due to increased production of deleterious mutations. Sturtevant (1937) asked: “why does the mutation rate not become reduced to zero?” And gave the reply: “In short, mutations are accidents, and accidents will happen”. Without mutation, evolution would not be possible, and life itself could never have arisen in the first place. Thus, as Baer (2008) noted “deleterious mutations are the price living organisms pay for the ability to evolve”. Naively one may think that organisms should evolve as low mutation rates as possible. It has been concluded that “natural selection of mutation rates has only one possible direction, that of reducing the frequency of mutation to zero” (Williams, 1966a). Thus, there should be a strong selective pressure to eliminate mutation altogether, due to the high probability that any particular mutation will have deleterious effects. Accordingly, theory indicates that under most conditions, selection puts a premium on the faithful maintenance and transmission of genetic information and is expected to favor alleles that reduce the mutation rate (Karlin and McGregor, 1974; Feldman and Liberman, 1986; Kondrashov, 1995; Sniegowski et al., 2000; Sniegowski, 2004). In fact, DNA replication can have a remarkable fidelity, estimated to produce 10-9–10-11 mutations/nucleotide, achieved by multiple mechanisms of error avoidance and correction (Kunkel, 2004). Eigen (1992) argued that replication error rates established themselves near an error-threshold where the best conditions for evolution exist. However, analyzing the genomes of a bacterium, a yeast, and a nematode, Ackermann and Chao (2006) found that codons are used to encode proteins in a way that avoids the emergence of mononucleotide repeats, suggesting that selection for genetic stability may be more important than selection for the generation of variation. In well-adapted populations in stable environments the rate of mutation will evolve towards lower values (Leigh, 1973; Karlin and McGregor, 1974; Liberman and Feldman, 1986; Drake, 1991; Kunkel, 2004) until further improvements in replication fidelity and DNA repair become too costly (André and Godelle, 2006) reflecting the combined metabolic and temporal costs of perfection in replication and transcription fidelity (Kimura, 1967; Sniegowski et al., 2000). Another theory suggests that if the mutation rate is reduced to a very low level, a point will eventually be reached at which the small advantage of any further reduction is overwhelmed by the power of drift (Lynch, 2011).

Measurements of mutation rates, however, show that organisms have copying fidelities much lower than what could be expected from this assumption (Drake, 1993; 1999; Drake et al., 1998). A genotype with a null mutation rate would be sentenced to extinction because of its inability to respond to environmental perturbations. Dobzhansky (1950), in a seminal statement on adaptation to diverse environments, wrote ‘Changeable environments put the highest premium on versatility rather than on perfection in adaptation’. In changing environments, the potential benefit of generating a few adaptive variants may outweigh the cost of many deleterious mutations (Muller, 1928; Sturtevant, 1937; Radman, 1999). Also for host–parasite interactions, it has been shown that mean fitness is optimized by high or non-zero mutation rates, in a manner similar to that found with fluctuating abiotic environments (Nee, 1989; Sasaki, 1994; Haraguchi and Sasaki, 1996; M'Gonigle et al., 2009). Mathematical analyses have shown that evolution yields optimal, positive mutation rates under certain conditions (Kimura, 1967; Leigh, 1970; 1973; Eschel, 1973a; b; Gillespie, 1981; 1991a; Holsinger and Feldman, 1983; Liberman and Feldman, 1986; Bedau and Packard, 2003; Clune et al., 2008; Dees and Bahar, 2010). In fact, contrary to Williams’ intuitive prediction, strains having high mutation rates (mutators) are not rare in natural bacterial populations (Giraud et al. 2001a; Denamur and Matic, 2006). They have been found in populations of Escherichia coli (LeClerc et al., 1996; Matic et al., 1997; Denamur et al., 2002; Baquero et al., 2004), Salmonella enterica (LeClerc et al., 1996), Neisseria meningitides (Richardson et al., 2002), Haemophilus influenza (Watson et al., 2004), Staphylococcus aureus (Prunier et al., 2003), Helicobacter pylori (Björkholm et al., 2001), Streptococcus pneumoniae (del Campo et al., 2005) and Pseudomonas aeruginosa (Oliver et al., 2000; Ciofu et al., 2005; Maciá et al., 2005), with frequencies ranging from 0.1% to above 60%. Experimental (Mao et al., 1997; LeClerc et al., 1998) and theoretical (Boe et al., 2000) studies indicate that the frequency of mutators observed among natural isolates is much higher than expected from mutation/selection equilibrium, which suggests that there are situations in nature where being a mutator confers a selective advantage (Giraud et al. 2001a; Tanaka et al., 2003; Denamur and Matic, 2006; Ram and Hadany, 2012). Enrichment of spontaneously originated mutators in microbial populations undergoing adaptation to a new environment has been shown in laboratory experiments (Sniegowski et al., 1997; Miller et al., 1999; Notley-McRobb et al., 2002a; b). In long-term experiments with E. coli, 3 of 12 populations spontaneously evolved into mutators in 10,000 generations (Sniegowski et al., 1997) and a fourth by 20,000 generations (Cooper and Lenski, 2000). Also antimutator strains, with a significantly lower mutation rate, can be found (Fijalkowska et al., 1993). This suggests that mutation rates are adjustable and subject to selection (Taddei et al., 1997; Sniegowski et al., 2000; De Visser, 2002).

5.2 May mutagenesis be adaptive?

Environments never remain static. They continuously undergo changes that alter the fitness landscapes, displacing populations towards suboptimal fitness regions, where the amount of mutations with positive effects increases. These poorly-adapted populations could benefit from having higher than standard mutation rates (Stich et al., 2010). The variation of mutation effects with fitness, together with the fact that error rates can be easily modified as a consequence of mutations producing genotypes with variable capacity to cause errors, suggest that mutation rates are a character subjected to the action of natural selection (Drake et al., 1998; Sniegowski et al., 2000). Selection of mutation rates was studied by a variety of models (Taddei et al., 1997; Orr, 2000; Andre and Godelle, 2006; Sniegowski et al., 2000; Gerrish et al., 2007; Heo et al., 2009). The mutation rate of an organism is subject to heritable variation, and therefore evolves (Weber, 1996; Kirschner and Gerhart, 1998; Radman et al., 1999; Metzgar and Wills, 2000; Sniegowski et al., 2000; Earl and Deem, 2004; Aharoni et al., 2005; Pigliucci, 2008; Wagner, 2008a).

Fisher’s fundamental theorem of natural selection (1930) states that the rate of increase in fitness of a population at any time is proportional to its genetic variance in fitness at that time. If there is no variation in survival and reproduction or if this variation has no genetic basis, then the composition of a population will not evolve over time. The theorem has been confirmed experimentally in animals and plants. The degree of fitness evolved is greater when the initial amount of genetic variability is larger (Strickberger, 1965; Ayala, 1968a; Vrijenhoek, 1994; Reed and Frankham, 2003; Leimu et al., 2006). Populations with genetic variability increased by hybridization adapt to the environment at a faster rate than parental, genetically less variable populations (Strickberger, 1965; Ayala, 1965; 1968 a; b). Adaptation theory also assigns a central role to mutation (Orr, 2000a; Grenfell et al., 2004). Adaptation by natural selection occurs through the spread and substitution of mutations that improve the performance of an organism and its reproductive success in its environment. An important focus of evolution experiments using microorganisms has been to investigate the dynamics of this process (Elena and Lenski, 2003). It seems that nature selected for those organisms with a mutation rate that compromises between adaptability and adaptedness (Radman et al., 2000). Mathematical models suggested that mutation rates adapt up (or down) as the evolutionary demands for novelty in variable environments (genetic innovation) or memory in stable environments (genetic conservation) increase (Bedau and Packard, 2003; Buchanan et al., 2004; Clune et al., 2008; Dees and Bahar, 2010). Similarly, the optimal genomic mutation rate was found to depend only on the environmental change and its severity (Nilsson and Snoad, 2002; Ancliff and Park, 2009). Even in stable abiotic environments, relatively high mutation rates may be observed for traits subject to cyclical frequency-dependent population dynamics (Allen and Scholes Rosenbloom, 2012). Studies in bacteriophage T4 (Drake et al., 1969; Schapper, 1998), E. coli (Schapper, 1998) and Drosophila (Nöthel, 1987) have demonstrated that mutation rates can be driven lower than (or can increase above) wild type values under various external pressures, and can be restored to wild type when control conditions are re-established (Sniegowski et al., 1997).

Mutation rate variability is a key theme in current discussions of the need for an extended evolutionary synthesis (Pigliucci, 2007; Pennisi, 2008) under the name of ‘‘the evolution of evolvability’’, the tantalizingly recursive possibility that the ability of organisms to evolve is itself a trait, or spectrum of traits, under evolutionary control (Bedau and Packard, 2003; Earl and Deem, 2004; Bell, 2005; Pigliucci, 2008; Dees and Bahar, 2010). Genes which may cause elevated rates of mutation (`mutator genes’) have been identified in many species (Demerec, 1937; Neel, 1942; Ives, 1950; Smith, 1992; Goho and Bell, 2000). Mutator bacterial strains are more antibiotic resistant than non-mutator strains, and thus have a clear selective advantage, potentially leading to an increase in the overall mutation rate within a population (Maciá et al., 2005; Denamur and Matic, 2006). In 115 replicate populations to characterize the genetic response of E. coli to high temperature (42.2°C), during 2000 generations one of the lines had evolved a mutator phenotype, many more mutations, and a higher fitness (Tenaillon et al., 2012). Overall, stressful conditions can lead to a genetically controlled increase in bacterial and eukaryote mutagenesis that is genome-wide (Taddei et al., 1995; Torkelson et al. 1997; Rosche and Foster, 1999; Goho and Bell, 2000; Shaver et al., 2002; Bjedov et al., 2003; Tenaillon et al., 2004; Saint-Ruf and Matic, 2006; Foster, 2007). These stresses include nutritional deprivation, DNA damage, temperature shift, or exposure to antibiotics. All of these global stress responses include functions that can increase genetic variation. In particular, up-regulation and activation of error-prone DNA polymerases, down-regulation of error-correcting enzymes, and movement of mobile genetic elements are common features of several stress responses. The result is that under a variety of stressful conditions, bacteria are induced for genetic change (Foster, 2007). Transformation is only one of several bacterial stress responses that advance genetic variation (Msadek, 1999; Merlin et al., 2000; Storz and Hengge-Aronis, 2000). Several authors have proposed that some form of stress-induced increase in the mutation rate might be favored by natural selection (Echols, 1981; Wills, 1984; McDonald, 1987). According to Thoday’s fitness definition (1953) “increase of fitness,..., must be brought about largely by changes which increase either genetic stability or variability without bringing about corresponding decrease in the other component. A progressive change is thus one that increases the sum of these components”. Microorganisms achieve this goal by balancing robustness and evolvability (Lenski et al., 2006).

At the end of 31 days of cocultivation for 320 generations of 69 E. coli mutants with polymerase fidelities varying by more than 6 orders of magnitude, all surviving strains were moderate mutators with 10 to 47 lower fidelity than the wild type, whereas antimutators and extreme mutators had been outcompeted (Loh et al., 2007; Loeb et al., 2008). Under specific conditions, selective pressure favors mutator strains of bacteria over nonmutator strains in both natural (Gross and Siegel, 1981; Leclerc et al., 1996; 1998; Matic et al., 1997; Oliver et al., 2000; Bjorkholm et al., 2001; Denamur et al., 2002; Giraud et al., 2002; Richardson et al., 2002; Blázquez, 2003; Prunier et al., 2003; Watson et al., 2004; del Campo et al., 2005; Labat et al., 2005; Denamur and Matic, 2006) and experimental populations (Cox and Gibson, 1974; Trobner and Piechocki, 1981; Chao and Cox, 1983; Chao et al., 1983; Mao et al., 1997; Sniegowski et al., 1997; Giraud et al., 2001a; Notley-McRobb et al., 2002; Shaver et al., 2002; Thompson et al., 2006; Pal et al., 2007). These situations are fulfilled under a variety of stressors (Gonzalez et al., 2008; Ram and Hadany, 2012). A generic thermodynamical analysis of genetic information storage suggests that an elevated stress level (i) reduces an organism’s ability to deflect mutagenic influence and/or (ii) restricts or redistributes metabolic resources available to mutation suppression in an unstressed situation and/or (iii) impairs the means to utilize these resources. Vice versa, any influence exerting one or several of the aforementioned effects can be understood as stress (Hilbert, 2011). As the environment at the ecological extremes is stressful by definition, an increase in mutation rates as a stress response can play a fundamental role in adaptation to new conditions (Gostincar et al., 2010). This response has been seen in a variety of organisms, in the laboratory as well as in wild strains of, for example, Escherichia coli (Rosenberg et al., 1998; Bjedov et al., 2003; Hastings et al., 2004), S. cerevisiae (Heidenreich et al., 2003) and Caenorhabditis elegans (Rosenberg and Hastings, 2004). Mutators have an advantage against phages in the coevolutionary arms race (Pal et al., 2007) and play a role in the emergence of antibiotic-resistant bacteria (Chopra et al., 2003; Woodford and Ellington, 2007). Freezing can increase the mutation rates of mtDNA in S. cerevisiae (Stoycheva et al., 2007a), and activate retrotransposons (Stamenova et al., 2008). Frequencies of new spontaneous mutations of Sordaria fimicola, Penicillium lanosum and Aspergillus niger are clearly related to whether they had been isolated from a region of high or low microclimatic stress, being much lower in strains from the north-facing, less stressful "European" slope of “Evolution Canyon I”, compared with strains from the south-facing "African" slope, which is a much more stressful environment (Lamb et al., 1998; 2008). Similarly, isolates of the ericoid mycorrhizal fungus Oidiodendron maius from stressed sites showed a significantly higher polymorphism in the Cu,Zn superoxide dismutase (Cu,ZnSOD) promoter region, suggesting that environmental stress may increase the rate of mutations in specific regions of the Sod1 locus (Vallino et al., 2011). Mutator strains were found to have a fitness advantage over wild-type strains, which stemmed from the fact that the mutator generated more beneficial mutations. The advantage of the mutator strain was shown to be frequency dependent, so that below a threshold level — ~1/10,000 of the population — the mutator went extinct, because then the probability of a beneficial mutation arising in the wild-type population was greater than in the much smaller mutator population. Above this level, however, the mutator was able to out-compete the wild-type because beneficial mutations had a greater chance of arising in the mutator strain and their selective advantage caused the linked mutator to hitchhike to high frequency (Gibson et al., 1970; Chao and Cox, 1983; Taddei et al., 1997; Rainey, 1999; Tenaillon et al., 1999; Elena and Lenski, 2003; Denamur and Matic, 2006). Mutational heritability in the asexual bacteria Pseudomonas fluorescens was estimated as 1 x 10-3 per generation in simple, single-substrate environments and 3 x 10-3 per generation in complex, four-substrate environments. Populations selected in complex environments evolved into genetically diverse communities that exhibited greater metabolic differentiation from other genotypes in their own population than to genotypes evolving in other populations, presumably as a result of resource competition. In populations selected in simple environments, little genetic diversity evolved, and genotypes shared very similar phenotypes (Barrett and Bell, 2006). Evolution in complex environments results in a genetically diverse population of overlapping generalists, each of which is adapted to a certain range of substrates but not all (Barrett et al., 2005).

Since too high a mutation rate has obvious negative consequences (Fisher, 1930; Crow, 1997b; Giraud et al., 2001a; Gerrish et al., 2012), there is evidence that evolution may have established brakes that ensure a “speed limit” to mutagenesis and select against high mutation rates when fitness gains no longer counterbalance the fitness loss due to continuous generation of deleterious mutations (Gerrish and Lenski, 1998; de Visser et al., 1999; Rainey, 1999; Sniegowski et al., 2000; Denamur and Matic, 2006; Pybus et al., 2007). High mutation rates help organisms to reach high fitness peak, after which they are no longer advantageous, and lower mutation rates get fixed to localize the population in the fitness peak (Heo et al., 2009). However, many bacteria can increase their mutation rates 10–100-fold without noticeable loss of fitness (Matic et al., 1997; Sniegowski et al., 1997; Shaver et al., 2002). A quantitative analysis revealed that evolution rates increase linearly with mutation rates for slowly mutating viruses. This relationship plateaus for fast mutating viruses (Sanjuán, 2012). The evolution of bacterial populations may happen through alternating periods of high and low mutation rates (Giraud et al., 2001a; b). Highly illustrative is the behavior of bacteria kept under metabolic stress in stationary phase. To survive, cells must scavenge whatever becomes available through excretion or death in an otherwise starving population. After a period of mortality, mutants emerge that can survive and grow under such conditions (Zambrano et al., 1993; Finkel and Kolter, 1999; Zinser and Kolter, 2000). Null mutations in rpoS confer this advantage (Zambrano et al., 1993) (see chapter 12.1).

Binomial sampling error and background selection both generate random change in the frequency of a mutation across generations. The former is an inevitable consequence of finite population size and it causes the frequency of a mutation to change in a manner that is independent of direct selection on the mutation itself. Its magnitude is inversely proportional to population size, N. The latter causes additional noise due to a mutation’s chance association with genetic backgrounds of different selective value. Its magnitude is proportional to the genetic standard deviation in genome-wide fitness, σW(gen). The dilution of direct selection on a mutation due to the additional noise generated by background selection can be expressed by a lowered effective population size, Ne, in comparison to the census size N (Hill and Robertson, 1966; Felsenstein, 1974; Rice and Chippindale, 2001). The effective size of a population (Ne) is a fundamental determinant of nearly all aspects of evolution as it determines the probability of (and times to) fixation or removal of mutant alleles (Lynch, 2006; Charlesworth, 2009). Ne is generally much smaller than the actual size of a population, as a consequence of variation in family size, nonrandom mating, sex-ratio bias, and many other aspects of population structure (Crow and Morton, 1955; Frankham, 1995). Comprehensive estimates of the Ne/N ratio (that included the effects of fluctuation in population size, variance in family size and unequal sex-ratio) ranged from 0.11 (Frankham, 1995) to 0.25 and 0.75 (Nunney, 1993; 1996) to but can also be much smaller (Hedgecock, 1994; Waples, 1998; Hedrick, 2005). Response to selection is dependent on population size for both new mutations (Hill, 1982 a; b; Caballero et al., 1991; López and López-Fanjul, 1993; Mackay et al., 1994; Samani and Bell, 2010) and standing genetic variation (Jones et al., 1968; Weber, 1990; Weber and Diggins, 1990; Agashe et al., 2011; Lachapelle and Bell, 2012). Thus, the rate of adaptive evolution is positively correlated to Ne (Wright, 1931; Kimura, 1983; Keightley, 2004; Leimu et al., 2006; Willi et al., 2006). Small effective population sizes, on the other hand, are believed to have an increased risk of extinction (Lande, 1994; Kondrashov, 1995a; Lynch et al., 1995a; b; Willi et al., 2006; Bell and Gonzalez, 2009; Willi and Hoffmann, 2009). The ability of a population to incorporate beneficial mutations and to purge deleterious mutations should scale positively with population size, assuming that Ne scales positively with N (Lynch, 2006; 2007a). However, associations between population size and genetic variance may be complex (Barton and Turelli, 2004). Muller (1964) proposed that an asexual organism will inevitably accumulate deleterious mutations, resulting in an increase of the mutational load and an inexorable, ratchet-like, loss of the least mutated class. In large populations, Muller’s ratchet occurs more slowly, and even lower rates of recombination will effectively arrest mutation accumulation (Charlesworth et al., 1993). Large populations more rapidly produce variants carrying multiple mutations that can evade constraints such as fitness valleys (Weinreich and Chao, 2005). As the rate of long-term adaptive evolution appears to be limited by the supply of new mutations (Gossmann et al., 2012), species with larger Ne are expected to undergo more adaptive evolution than species with small Ne because a greater number of advantageous mutations appear in the population and a higher proportion of these mutations are effectively selected (Jensen and Bachtrog, 2011; Gossmann et al., 2012; Phifer-Rixey et al., 2012). Accordingly, mathematical models observe a low mean population size for low values of the maximum possible mutation size μ, a sharp rise in population size for intermediate values of μ, and a plateau in the population size for high values of μ (Dees and Bahar, 2010). For low μ, new organisms remain in tight clusters, unable to explore the fitness landscape far beyond the locations of their parents. This results in overcrowding, followed by a decrease of the population and eventual extinction. For high μ, the population cannot grow indefinitely, constrained by both the overpopulation limit and the finite fitness landscape size, which combine to play a role analogous to the carrying capacity in logistic population models (Dees and Bahar, 2010).

Levels of genetic diversity in plant and animal wildlife are related to population size (Frankham, 1996; 1997; Spielman et al. 2004; Leimu et al., 2006; Willi et al., 2007; Hoffmann and Willi, 2008; Charlesworth, 2009). Under assumptions of neutrality and additive gene effects, equilibrium additive genetic variance of a quantitative trait increases linearly with effective population size (Chakraborty and Nei, 1982; Lynch and Hill, 1986). On the other hand, the amount of genetic diversity a population contains has been shown to correlate with current fitness and, in the case of heritabilities, with evolutionary potential (Fisher, 1930; Falk and Holsinger, 1991; Frankham, 1995; Falconer and Mackay, 1996; Franklin and Frankham, 1998; Reed and Frankham, 2003; Johnson et al., 2006; Leimu et al., 2006). A certain amount of genetic load is considered the necessary cost of evolutionary responsiveness (Crow and Abrahamson, 1965). Literature suggests that allozyme heterozygosity is a good measure of population fitness and adaptive potential (Garten, 1976; Soulé and Wilcox, 1980; Beardmore, 1983; Allendorf and Leary, 1986; Houle, 1989; Merilä and Crnokrak, 2001) although the association between neutral diversity and population fitness is highly variable across studies (Arden and Lambert, 1997; Calero et al., 1999; Booy et al., 2000; Noel et al., 2007). Moreover, estimates of variation within natural populations often do not support the prediction that quantitative genetic variation declines at small Ne (Willi et al., 2006).

Bigger animals not only live longer but they also display a set of correlated life-history traits (Stearns, 1983; 1989; Promislow and Harvey, 1990; Promislow et al., 1992; Hendriks and Mulder, 2008; Jeschke and Kokko, 2009). Specifically, they mature more slowly and breed for the first time at a later age. Their reproductive rates are lower (longer intervals between reproductive events) and their productivity at each breeding event is also reduced (lower litter or clutch sizes) (Speakman and Król, 2010). Body mass can be taken as a proxy for Ne (Berlin et al., 2007; Popadin et al., 2007; Nabholz et al., 2008b; 2009), based on the nearly universal inverse relationship between body mass/size and population size and density (Damuth, 1981; 1987; Peters, 1983; Griffiths, 1998; Ackerman et al., 2004; Arim et al., 2011). Thus, increased multicellularity is generally associated with a reduction in Ne, and this in turn reduces the efficiency of natural selection (Lynch, 2010a). The nearly neutral theory predicts that the rate and pattern of molecular evolution will be influenced by Ne, because in small populations more slightly deleterious mutations are expected to drift to fixation (Ohta and Kimura, 1971; Ohta, 1992; 1993). As predicted by theory, island lineages have significantly higher ratios of non-synonymous to synonymous substitution rates than mainland lineages (Johnson and Seger, 2001; Woolfit and Bromham, 2005). The higher supply of beneficial mutations allows large microbial populations to adapt more rapidly to new environments compared to small populations (Gillespie, 1999; Orr, 2000a; Wilke, 2004, de Visser and Rozen, 2005). The adaptive walk of asexually reproducing populations can be described as travelling wave whose speed is determined by population size and mutation rate (Fisher, 2011; Hallatschek, 2011).

Similarity of mutation rates among lineages with vastly different generation lengths and physiological attributes points to a much greater contribution of replication-independent mutational processes to the overall mutation rate (Kumar and Subramanian, 2002). Species with long generation times, and small effective population size, appear to have higher rates of mutation per generation (Keightley and Eyre-Walker, 2000; Piganeau and Eyre-Walker, 2009). Direct estimates are limited, but suggest that the nuclear mutation rate per generation ranges over 100-fold, from 3.3 x 10-10 per site in Saccharomyces cerevisiae to 3.5 x 10-9 in Drosophila melanogaster, 2 x 10-8 in Homo sapiens, and 3.8 x 10-8 in Mus musculus (Keightley et al., 2009; Lynch, 2010a). Estimates of human germline base substitution rates range from 0.97 to 3.8 x 10-8 per base per generation (Haldane, 1935; Kondrashov and Crow, 1993; Crow, 1993; Nachman and Crowell, 2000; Kondrashov, 2003; Xue et al., 2009; Lynch, 2010b; Roach et al., 2010; The 1000 Genomes Project Consortium, 2010; Awadalla et al., 2011; Conrad et al., 2011; Kong et al., 2012; Sun et al., 2012). A human mutation rate of ~2.5 x 10-8 mutations per nucleotide site corresponds to 175 (range 91–238) mutations per diploid genome per generation (Nachman and Crowell, 2000). Assuming that (i) there is no selective bottleneck between gametogenesis and offspring (but see chapter 8) and (ii) 95 to 100% of all mutations have either neutral (~30% of amino-acid changing mutations in humans, ~16% in Drosophila [Eyre-Walker, 2002], and ~2.8% in enteric bacteria [Charlesworth and Eyre-Walker, 2006]) or deleterious effects on fitness (Eyre-Walker and Keightley, 2007) this ratio would mean that with each generation, fitness of human populations should decline. A recent study revealed a gradual decline in the proportion of nonsynonymous single nucleotide polymorphisms (SNPs) (considered deleterious) from tip to root of the human population tree. Up to 48% of nonsynonymous SNPs specific to a single genome were regarded deleterious in nature (Subramanian, 2012). It was estimated that at least 1.6 harmful new mutations per individual per generation have been arising in our lineage for the last several million years (Crow, 1999; Eyre-Walker and Keightley, 1999). Due to the effects of smaller Ne on the force of natural selection (Frankham, 1995; Lynch, 2010a), natural selection is deemed insufficient to deal with this mutational load (Eyre-Walker and Keightley, 1999; Eyre-Walker et al., 2002; Harris, 2010). It has been concluded that taking into account the human low reproductive rate it is difficult to envisage how such a high load can be tolerated by hominid populations (Kondrashov, 1995a; Crow, 1997b; Eyre-Walker and Keightley, 1999; Sunyaev et al., 2001; Eyre-Walker et al., 2002; Eöry et al., 2010). On the basis of genome-wide sequencing it has been estimated that, on average, each person carries approximately 250 to 300 loss-of-function variants in annotated genes and 50 to 100 variants previously implicated in inherited disorders (The 1000 Genomes Project Consortium, 2010).

Fisher (1930) concluded that sexual populations may have a more rapid rate of evolution than would an otherwise equivalent group of asexual organisms. Fisher's conclusion depends on the rate of mutation (Ridley, 2003):

• If favorable mutations are rare, each one will have been fixed in the population before the next one arises. New favorable mutations will always arise in individuals that already carry the previous favorable mutation. Sexual and asexual populations then evolve at the same rate.

• If favorable mutations arise more frequently, Fisher's argument works: the sexual population evolves faster. Each new favorable mutation will usually arise in an individual that does not already possess other favorable mutations; the greater speed with which the different favorable mutations combine together causes the sexual population to evolve faster. The higher the rate at which favorable mutations are arising, the greater the evolutionary rate of a sexual relative to an asexual population.

Given the deleterious effects of most mutations, it can be predicted that increasing the mutation rate should generate progressively larger reductions in fitness, assuming that the mutations accumulate in the genome and that the effects of the mutations are additive or act synergistically (Schultz and Lynch, 1997). This prediction has been confirmed by most studies that have examined the fitness effects of elevated mutation rates (Rosenbluth et al., 1983; Drake et al., 1998; Davies et al., 1999; Manoel et al., 2007; Morran et al., 2010). Ethyl methanesulfonate (EMS) mutagenesis experiments, in which controls are given identical treatment to mutagenized lines, other than a dose of mutagen, have also shown consistently strongly negative effects on fitness traits in Drosophila (Mukai, 1970; Ohnishi, 1977; Mitchell and Simmons, 1977; Temin, 1978; Keightley and Ohnishi, 1998; Yang et al., 2001) and C. elegans (Keightley et al., 2000). Similarly, transposable element insertional mutagenesis leads to reduced fitness in Drosophila (Mackay et al., 1992) and E. coli (Elena and Lenski, 1997). However, efficiency of artificial selection in agricultural plants can be enhanced following the increased genetic variation induced by mutagenesis with irradiation (Gregory, 1955; Brock, 1971; Chavan and Chopde, 1982; Micke et al., 1987; Ahloowalla et al., 2004; Patade and Suprasanna, 2008). Ionizing radiation was also used to increase genetic variation in a variety of insects. The response to selection was dependent on the genetic background, size of the irradiated population and dose of radiation. After an initial decline, population fitness and other responses to artificial selection increased in the irradiated populations compared with controls (Wallace, 1958; Mukai, 1961; Crenshaw, 1965; Ayala, 1966; 1967; 1969; van Delden and Beardmore, 1968; Blaylock and Shugart, 1972; Maruyama and Crow, 1975). Thus, it is essential for the evolutionary significance of mutations to take population size, mutation rate and the action of selection into account.

Cancer arises from a Darwinian process of mutation and selection among somatic cells (Frank and Nowak, 2004). Theoretical models and clinical observations confirm that mutator mechanisms are generally more efficient routes to tumorigenesis than non-mutator mechanisms. Mutations, generated via a mutator phenotype, accelerate carcinogenesis (Loeb, 1998; Strauss, 1998; Beckman and Loeb, 2006; Bielas et al., 2006; Loeb et al., 2008; Beckman, 2009). In carcinogenesis, the size of the stem pool affects the spread of deleterious and advantageous mutations (Frank, 2003a; Michor et al., 2003). In a large stem pool, oncogene and tumor suppressor mutations with increased rates of proliferation almost always succeed, whereas mutator mutations with decreased rates of proliferation rarely succeed. Put another way, natural selection among cell lineages deterministically takes its course in a large population. In small stem pools, chance events can influence which cell lineages succeed or fail. Small pools increase the probability that deleterious mutator mutations spread and decrease the probability that advantageous mutations spread (Frank and Nowak, 2004). Thus, in a variety of evolutionary events mutation rate and population size are the crucial determinants of adaptive evolution dynamics (Gerrish and Lenski, 1998; Hallatschek, 2011).

From these relationships a general framework for the fitness effects of  mutations in a variable environment can be sketched:

  • Due to the unfavorable ratio of deleterious to beneficial mutations, only a high amount of mutations gives the chance to generate some beneficial mutations.
  • There is both a lower and upper limit to mutagenesis intensity that may be beneficial for population fitness.
  • Due to this ratio, populations with a low effective population size should invest heavily into DNA repair. Otherwise they are doomed. However, their evolvability in variable environments is poor.
  • Populations with a high population size are able to engage in a mutagenic bet-hedging strategy.
  • Thus, a high effective size of the mutating population is prerequisite for what I would like to call the “selecting the pearl(s) among the pebbles” approach: An efficient selection regime, either due to natural selection or an experienced breeder (like in the case of irradiated crop plants) that is able to pick the “pearl(s)” of beneficial or at least neutral mutations out of the “pebbles” of deleterious ones may change the course of evolution

6. Developmental and reproductive biology hold the clue


… one finds two very different areas, which may be designated functional biology and evolutionary biology. To be sure, the two fields have many points of contact and overlap. Any biologist working in one of these fields must have a knowledge and appreciation of the other field if he wants to avoid the label of a narrow-minded specialist.

Ernst Mayr, 1961

Summary

Intriguingly, the major theories on the evolutionary rationale for the origin and maintenance of sexual reproduction treated the very processes involved in sexual reproduction either as a black box or reduced its evolutionary potential to the process of recombination. However, to gauge the full evolutionary importance of sexual reproduction, its varied ontogenetic and gametogenetic processes have to be taken into account. To this end, I embark on an inventory of the molecular toolboxes of developmental and reproduction biology from a phylogenetic point of view. The earliest both ontogenetic and reproductive events were responses to metabolic stress that occurred cyclically during the feast and famine life cycles of microorganisms. The molecular events created dormant seeds/propagules as a bet-hedging strategy with the ability to enter a reversible state of low metabolic activity when faced with unfavorable environmental conditions but capable of being resuscitated following environmental change. The morphogens that induced microbial dormancy such as cyclic AMP and differentiation-inducing factors (DIFs) elicit evolutionarily highly conserved ontogenetic events in metazoan cells. In this phylogenetic legacy, metazoan growth/differentiation factors induce cellular stress reactions. Cellular differentiation is associated with epigenetic reprogramming resulting, due to stochastic epigenetic instability, in significant epigenetic differences between differentiated cells. This variation impacts the cells’ competitive abilities for limiting amounts of trophic factors. The losers of these cell competitions are eliminated. Specification of primordial germ cells proceeds via different pathways, i.e. preformation or epigenesis. Spermatogenesis, the sequence of male germ cell development, is a long, orderly, and well-defined process in vertebrates occurring in seminiferous tubules of the testis whereby undifferentiated spermatogonial germ cells evolve into maturing spermatozoa. The temporal course of oogenesis is highly variable across taxa. In most species, oocytes are much larger than sperm, requiring a larger investment of resources. Importantly, during gametogenesis at various stages of the process a huge amount of germ cells are eliminated by apoptosis. Thus, depending on the species, only a limited number of gametes are allowed to pass their genetic information to the next generation.

In 1975 Feldman and Lewontin wrote: “There is a vast loss of information in going from a complex machine to a few descriptive parameters. Therefore, there is immense indeterminacy in trying to infer the structure of the machine from those few descriptive variables, themselves subject to error. It is rather like trying to infer the structure of a clock by listening to it tick and watching the hands”. The fundamental implications of this insight extend to virtual any complex phenomenon of life such as aging (Heininger, 2012) and sexual reproduction. Intriguingly, the major theories on the evolutionary rationale for the origin and maintenance of sexual reproduction treated the very processes involved in sexual reproduction as a black box or to stay in the picture of Feldman and Lewontin (1975) as a clock whose working they tried to infer from its tick and tack. It was one of the strange realizations gained from my literature work that both evolutionary biology, trying to decipher the evolutionary rationale of sexual reproduction, and reproduction biology, investigating the molecular processes of sexual reproduction, obviously lead completely independent lives with minimal, if any, information exchange (ignoring the a.m. wise advice of Ernst Mayr). Scientists’ specialization means that particular issues are looked at in isolation (sometimes called ‘silos’), rather than as an integrated whole (Joffe, 2010). Yet, since Aristotle's time, scientists have maintained that to understand a phenomenon in a scientific way we must know its ultimate cause(s) (Ruse, 2003).

In 1896, extended in 1904, A. Weismann devised the theory of Germinal Selection as an additional, hierarchic level to natural selection: competition and selection among the hereditary units within the germplasm (Weissman, 2011). Stephen Gould in his epic book “The Structure of Evolutionary Thought” (2002) extensively covered Weismann’s concept of germinal selection as an early hierarchical selection perspective. Selection means differential survival. What has often been overlooked when dealing with sexual reproduction is that it is invariably associated with death. I don’t mean the death of the reproductive organism that has repeatedly been addressed (Ruffié, 1986; Clark, 1996, Reznick and Ghalambor, 1999; Partridge et al., 2005). I mean the death of genetic units that during sexual reproduction are denied to pass their genetic information to the next generation: e.g. unicellular Dictyostelium discoideum that are engulfed by giant cells to form a multinucleated giant cell but disappear at the early stage of development (Okada et al., 1986) or the sometimes billionfold death of gametes in higher taxa.

6.1 A primer on the evolutionary roots of multicellular development

Knowledge is the object of our inquiry, and men do not think they know a thing till they have grasped the ‘why’ of it (which is to grasp its primary cause).

Aristotle

A variety of models for the evolution of multicellularity are based on the concept of division of labor (Willensdorfer, 2009; Gavrilets, 2010; Rossetti et al., 2010; Ispolatov et al., 2012). Such a division of labor does not necessarily need to occur in the form of soma and germ cells (Ispolatov et al., 2012). However, the basic physiological trade-off underlying the transition to differentiated multicellularity is between reproduction and viability, hence the germ–soma division of labor (Bonner, 2003; Kirk, 2003; Michod, 2006; 2007; Gavrilets, 2010). Studying the environmental conditions and processes underlying cellular differentiation at the unicellular-multicellular transition should provide valuable clues to understand the evolutionary roots and processes related to differentiation and reproduction (Buss, 1987; Heininger, 2001; Bonner, 2003). Studies of deep homology are showing that new structures need not arise from scratch, genetically speaking, but can evolve by deploying regulatory circuits that were first established in protists and early metazoans. The deep homology of shared genetic, biochemical, cellular and developmental mechanisms (Shubin et al., 1997; 2009; Gerhart, 2000; Gilbert and Bolker, 2001; Hall, 2003) is the “fossil record” (Runnegar, 1986; Buss, 1987) that each contemporary organism carries inside and that allows a glimpse into deep evolutionary time.

Cellular differentiation evolved as a stress response. Unicellular organisms have a “feast and famine” lifestyle, limiting amounts of nutrients being rather the rule than the exception and long periods of nutritional deprivation are punctuated by short periods that allow fast growth (Koch, 1971; Kolter et al., 1993, Morita, 1993; Msadek, 1999; Heininger, 2001). The evolutionary signature of persistent metabolic “feast and famine” cycling is still present in every eukaryotic cell (Johnston, 1999; Ravussin, 2002). The feast-to-famine transition is not merely a metabolic response to a drop in nutrient availability; this transition also involves cell-to-cell signaling pathways, the results of which range from sporulation to fruiting body and complex pattern formation (Shapiro and Dworkin, 1997; Shimkets, 1999). Arguably, the most ubiquitous and predictably recurrent environmental signal that triggers differentiation in unicellular organisms is nutrient shortage (Losick and Stragier, 1992; Hengge-Aronis, 1993; Yarmolinsky, 1995; Kaiser, 1996). At the crossroads of uni- and multicellularity, Dictyostelium discoideum, a non-metazoan social amoeba, is the model organism for the study of differentiation and development (Janssens and Van Haastert, 1987; Heininger, 2001; Williams, 2006; 2010; Bonner, 2009; Kawabe et al., 2009; Kessin, 2010), apoptosis/programmed cell death (Cornillon et al., 1994; Heininger, 2001; Bonner, 2009; Kessin, 2010), and both asexual and sexual reproduction (Godfrey and Sussman, 1982; Urushihara, 1992; Francis, 1998; Urushihara and Muramoto 2006; Bonner, 2009; Flowers et al., 2010; Amagai, 2011). In these life-cycle events, a variety of functional equivalences between proteins of Dictyostelium and other eukaryotic species were found (Annesley and Fisher, 2009). Intriguingly, metabolic stress is the trigger for Dictyostelium multicellular differentiation. Following nutrient deprivation, between 104 and 106 unicellular Dictyostelium discoideum aggregate into mobile slugs that are spatially heterogeneous with respect to cell fate. Upon extended starvation, slugs transform into stationary fruiting structures that consist of a round, spore-bearing sorus at the top of a long, thin stalk (Mohanty and Firtel, 1999; Williams, 2006). Differentiation to spores and apoptosis of the stalk cells are closely linked in a social stress response (Christensen et al., 1998). The stress response not only represents a primordial differentiation/apoptosis event but is also regarded as primordial reproduction event (Heininger, 2001; 2012; Foster et al., 2004; Jack et al., 2008). Cumulative evidence argues for the phylogenetic conservation of these processes. Dictyostelium-derived morphogens such as cyclic AMP and differentiation-inducing factors (DIFs) elicit dose-dependent and context-specific effects (Thomason et al., 1999; Williams, 2006; 2010) that are evolutionarily highly conserved. The conservation and ambiguity of the signals has been ascertained through their effects on various mammalian cell lines causing both growth arrest, differentiation and apoptosis (Asahi et al., 1995; Kubohara et al., 1995a; b; 1998; Kubohara, 1997; 1999; Miwa et al., 2000; Takahashi-Yanaga et al., 2003). Likewise, in the spheroidal green alga Volvox carteri, the most fundamental and primordial differentiation event creates germline and somatic cells (Kirk, 1998; Bonner, 2003; Heininger, 2012). The evolutionary conservation of these processes can be traced throughout phylogenesis following their continuity of genetic programming and the “fossil record” of the genome (Heininger, 2012).

Cellular stress is the overarching principle underlying cellular differentiation in metazoan organisms. Oxygen levels in the atmosphere only reached its present level approximately 350 million years ago (carboniferous period), clearly showing that cellular life on earth was well adapted to hypoxic conditions a long time before the present oxygen-dependent organisms appeared on earth (Wayne, 1985). Hypoxic microenvironments occur in both the developing embryo and adult, and often create specific ‘niches’ that regulate cellular differentiation (Maltepe and Simon, 1998; Simon MC et al., 2003). Hypoxia is a potent suppressor of mitochondrial oxidation and appears to promote “stemness” in adult and embryonic stem cells (Rehman, 2010). In adult stem cells, hypoxia prolongs the lifespan of the stem cells, increases their proliferative capacity, and reduces differentiation in culture (Fehrer et al., 2007; Jang and Sharkis, 2007). Embryonic stem cells are able to retain their pluripotency for a longer period of time when cultured in hypoxic conditions (Ezashi et al., 2005; Prasad et al., 2009). Components of the hypoxia-inducible factor-regulated system play essential roles in embryonic development and cellular differentiation (see chapter 7.2.4). Interestingly, Oct4, a transcription factor essential for maintaining stem cell pluripotency (see chapter 7.2.4), is a hypoxia- and hypoxia-inducible factor-regulated gene (Covello et al., 2006). On the other hand, hypoxia can also induce differentiation of cell lines (Short et al., 2004).

Mammal ontogeny occurs in a low oxygen environment (estimated to be roughly 1–6%; Guyton & Hall, 1966) during the initial stages of development similar to the environment in which early unicellular organisms lived (Jaffe, 1998). From fungi and plants to humans, oxidative stress is a trigger and marker of developmental events and cell differentiation (Sohal et al., 1986; Allen, 1991; 1998; Zs.-Nagy, 1992; Blackstone, 1999; 2009; Heininger, 2001; 2012; Orzechowski et al., 2002; Foyer and Noctor, 2005; de Magalhães and Church, 2006; Gapper and Dolan, 2006; Pitzschke et al., 2006; Hitchler and Domann, 2007; Covarrubias et al., 2008; Nasution et al., 2008; Scott and Eaton, 2008; Owusu-Ansah and Banerjee, 2009; Hernández-García et al., 2010; Vincent and Crozatier, 2010; Sardina et al., 2012). Generally, growth factors stimulate cellular growth and differentiation through the formation of reactive oxygen species (Burdon and Rice-Evans, 1989; Radeke et al., 1990; Thannickal et al., 1993; Lo and Cruz, 1995; Sundaresan et al., 1995; Bae et al., 1997; Zafari et al., 1998; Sattler et al., 1999; Sung et al., 2000; Suzukawa et al., 2000; Thannickal and Fanburg, 2000; Tolando et al., 2000; Sauer et al., 2001; Menon et al., 2003; Sturrock et al., 2006; Meng et al., 2008; Morgan and Liu, 2010). During cellular differentiation of two human embryonic stem cell lines, mitochondrial superoxide production and cellular levels of reactive oxygen species increased as result of increased mitochondrial biogenesis. The expression of major antioxidant genes was downregulated despite this increased oxidative stress and DNA damage levels increased during differentiation, whereas expression of genes involved in different types of DNA repair decreased (Saretzki et al., 2004; 2008).

The p53 protein family is activated by (Lu and Lane, 1993; Wang and Ohnishi, 1997; Vousden and Lu, 2002; Murray-Zmijewski et al., 2006; 2008; Berns, 2010; Hölzel et al., 2010; Marchenko et al., 2010; Lu et al., 2011) and integrates multiple cellular stress signals and assesses cellular damages to trigger different cellular outcomes — ranging from cell-cycle arrest (Aoubala et al., 2011; Blagosklonny, 2011; Smits, 2012), DNA repair (Götz and Montenarh, 1996; Albrechtsen et al., 1999; Seo et al., 2002; Sengupta and Harris, 2005), genome stability (Lane, 1992; Vousden and Lane, 2007; Riley et al., 2008; Dulic, 2011; Spinnler et al., 2011), tumor suppression (Barlev et al., 2010; Ho et al., 2010; Feldser et al., 2010; Junttila et al., 2010; Cheok et al., 2011; Madan et al., 2011), induction of autophagy (Crighton et al., 2006; Amaravadi et al., 2007), cell migration (Roger et al., 2006), reproduction (Hu et al., 2007; Hu, 2009; Amelio et al., 2012), regulation of metabolism (Jones RG et al., 2005; Bensaad et al., 2006; Green and Chipuk, 2006; Matoba et al., 2006; Bensaad and Vousden, 2007; Olovnikov et al., 2009; Vousden and Ryan, 2009; Feng and Levine, 2010; Hu W et al., 2010; Madan et al., 2011; Maddocks and Vousden, 2011; Liang et al., 2013a), angiogenesis (Teodoro JG et al., 2006), and senescence, differentiation (Liu Y et al., 2009; Darzynkiewicz, 2010; Mallette et al., 2010; Manning and Kumar, 2010; Morselli et al., 2010; Horikawa et al., 2011; Lee et al., 2011; Vilborg et al., 2011; McGee et al., 2012; Tucci, 2012), to apoptosis (Mihara et al., 2003; Karlberg et al., 2010; Meley et al., 2010; Trinh et al., 2010; Hill et al., 2011; Koster et al., 2011; Nardinocchi et al., 2011; Vaseva et al., 2011). p53 is a key regulator of mitochondrial function, including ROS production and associated repair of mtDNA oxidative damage, as well as mtDNA replication and mitochondrial biogenesis (Ralph et al., 2010). Thus, p53 couples energy metabolism and ROS formation by modulating the transcription of target genes that control the fluxes through mitochondrial respiration, glycolysis, or the pentose phosphate shunt (Liu B et al., 2008; Maddocks and Vousden, 2011). ROS are potent activators of p53 function and, indeed, the redox balance is believed to be a key factor in the modulation of p53 activity (Jayaraman et al., 1997; Martindale and Holbrook, 2002).

How can the fascinating variety of cellular functions in differentiated cell lineages be generated from a single genome? Emerging biochemical, genetic, and functional evidence suggests that both genetic and epigenetic reprogramming is crucial for diverse biological processes, including primordial germ cell reprogramming, pluripotent stem cell differentiation, hematopoiesis, and cancerogenesis (Ficz et al., 2011; Wu and Zhang, 2011; Tan and Shi, 2012). So far, developmentally programmed genome rearrangement that leads to the elimination of portions of chromosomes or the loss of entire chromosomes during embryonic development has been described in protists, crustaceans, insects, and vertebrates (fishes, birds, marsupials) (Beermann, 1959; Nakai et al., 1991; 1995; Prescott, 1994; Kubota et al., 1997; 2001; Goto et al., 1998; Kohno et al., 1998; Watson et al., 1998; Sanchez and Perondini, 1999; Müller and Tobler, 2000; Goday and Esteban, 2001; Kloc and Zagrodzinska, 2001; Degtyarev et al., 2002; Bachmann-Waldmann et al., 2004; Matzke and Birchler, 2005; Pigozzi and Solari, 2005; Yao and Chao, 2005; Zufall et al., 2005; Drouin, 2006; Duret et al., 2008; Itoh et al., 2009; Smith et al., 2009; 2012; Kojima et al., 2010; Nemetschke et al., 2010; Nowacki et al., 2011; Sémon et al., 2012; Wang et al., 2012; Bracht et al., 2013; Sabin et al., 2013). Intriguingly, the process of DNA elimination in Tetrahymena and Paramecium is orchestrated by small RNAs which are expressed during conjugation and direct chromatin modifications to mark internal eliminated segment sequences for elimination (Matzke and Birchler, 2005; Nowacki et al., 2011; Sabin et al., 2013). Epigenetic mechanisms, including DNA methylation, histone modifications, and non-coding RNA-mediated regulatory events, are phylogenetically conserved to effect the long-lasting changes in gene expression during development and tissue differentiation of a variety of taxa (Wu and Sun, 2006; Hitchler and Domann, 2007; Singh et al., 2009; Feng et al., 2010; Hu B et al., 2010; Iorio et al., 2010; Mahpatra et al., 2010; Ficz et al., 2011; Wu and Zhang, 2011; Briones and Muegge, 2012; Hu et al., 2012; Sabin et al., 2013). Genome-wide dynamic DNA methylation changes occur during cellular differentiation (Lister et al., 2009; Geiman and Muegge, 2010; Nagae et al., 2011; Su et al., 2012). It has been estimated that the mitotic transmission fidelity of DNA methylation patterns is about three orders of magnitude lower than that of DNA sequence (an error rate of 1 in 106 and 1 in 103 for DNA sequences and DNA modification, respectively) (Ushijima et al., 2003; Laird et al., 2004; Riggs and Xiong, 2004; Genereux et al., 2005; Fu et al., 2010; Petronis, 2010). Errors in methylation pattern were mainly due to de novo methylations in unmethylated regions (Ushijima et al., 2003). This stochastic epigenetic instability can result in significant epigenetic differences accumulating over time across cells, despite the DNA sequence identity of these cells (Wong et al., 2005; Petronis, 2010). Thus, epigenetic reprogramming of embryonic stem cells and differentiating cells has a significant stochastic element (Ushijima et al., 2003; Reiss and Mager, 2007; Raj and van Oudenaarden, 2008; Mohn and Schübeler, 2009) resulting in genetic and epigenetic mosaicism (Rakyan et al., 2002; Petronis et al., 2003; Wong et al., 2005; Feinberg and Irizarry, 2010; Weinhouse et al., 2011). It can be expected that such genetic and epigenetic variation impacts the competitive abilities of cells, rendering them the raw material for selective events.

No tissue differentiates without the loss of cells by apoptosis. In a variety of tissues, competition for limiting amounts of trophic factors, e.g. nerve growth factor, determines life or death of cells (Hamburger and Oppenheim, 1982; Raff, 1992; Deckwerth and Johnson, 1993; McLean et al., 1997; Barres and Raff, 1999; Casaccia-Bonnefil et al., 1999; Bonner-Weir, 2000; Meier et al., 2000; Lee FS et al., 2001; Li L et al., 2001; Frago et al., 2003; Imaizumi et al., 2004). The losers of this cell competition die (Adachi-Yamada and O’Connor, 2004). Although this cellular selection has a stochastic and spatiotemporal component (Milán et al., 2002), predominantly the competitively inferior cells are eliminated (Adachi-Yamada and O’Connor, 2004; Igaki, 2009).

6.2 A primer on reproductive biology

Surprisingly, almost none of the numerous publications attempting to elucidate the evolutionary origin and maintenance of sex made reference to the molecular processes of sexual reproduction. In fact, these molecular processes allow telling inferences on the evolutionary rationale of sexual reproduction. A corollary of the lack of information exchange between evolutionary biology and reproduction biology is the strange phenomenon that, to my knowledge, the millionfold “waste” production of gametes so far was not taken into account in any one of the many theories about the evolutionary rationale of sexual reproduction. And, strangely, that reproduction biology failed to interpret this “waste” gamete production in an ecological-evolutionary context.

Primordial germ cells (PGCs) are the precursor cells to the eggs and sperm that comprise the germline. In many animals, a population of PGCs is set aside early in embryogenesis that is dedicated for the germ cell lineage while other cells in the embryo differentiate into soma giving rise to the organ systems of the adult. PGCs migrate to the somatically derived gonads and proliferate into germline stem cells that can self renew and differentiate into gametes. At least three general mechanisms are used to specify PGCs within the animal kingdom (Extavour and Akam, 2003; Johnson AD et al., 2003; 2011; Seydoux and Braun, 2006; Extavour, 2007; Rosner et al., 2009; Gustafson and Wessel, 2010). The germline can form (i) early in embryogenesis from an inheritance of maternal factors (maternally derived, also referred to as preformation) as used in flies and nematodes, (ii) by cell-cell interactions early in embryogenesis (inductive, also referred to as epigenesis) as seen in mice, and (iii) any time in the animal’s life, even in adulthood, from a multipotent stem cell precursor (persistent multipotent cell-derived germ cells), such as in planaria and hydra (a mode of germ cell specification pathway that also occurs in plants) (Gustafson and Wessel, 2010). Despite these developmental differences, animals employ a group of conserved molecular determinants for PGC specification. The most common of these is the gene Vasa (Raz, 2000; Mochizuki et al., 2001; Noce et al., 2001; Gustafson and Wessel, 2010). Anisogamy or gamete dimorphism is the rule in multicellullar animals and plants (Bell, 1982; Parker, 1982) — egg-producing females make a larger contribution to the zygote compared with the small contribution made by the sperm of males, but both males and females contribute 50% of the genes.

6.2.1 Male gametogenesis

Sperm cells are, arguably, the most differentiated cells, and spermatogenesis may very well be one of the most complex biological processes undergone by any organism. Spermatogenesis, the sequence of male germ cell development, is a long, orderly, and well-defined process occurring in seminiferous tubules of the testis whereby undifferentiated spermatogonial germ cells evolve into maturing spermatozoa. The functions of all of the cell types in the testis, the Leydig cells, the Sertoli (and/or peritubular) cells, the germ cells, and the vasculature, are interwoven in a highly organized manner (Sharpe et al., 1990). Mammalian spermatogenesis can be divided into two phases. The first round of spermatogenesis, which starts after birth, is characterized by the sequential appearance within seminiferous tubules of cells corresponding to each stage of germinal cell. The second phase is that of mature spermatogenesis, made up of the subsequent rounds of spermatogenesis which then occurs permanently in adult animals. Seminiferous tubules are composed of somatic cells (myoid cells, Leydig cells and Sertoli cells), and germ cells (spermatogonia, spermatocytes, and spermatids). During spermatogenesis, three distinct stages can be differentiated: spermatocytogenesis by mitotic proliferation of spermatogonia, meiotic development of spermatocytes, and spermiogenesis, the postmeiotic development of spermatids and maturation of the spermatozoon (Johnson, 1995; Johnson et al., 2000). Spermatogenesis is dependent on a specific microenvironment (niche) contributed by Sertoli, myoid, and Leydig cells for proper development.

The Sertoli cells (the somatic cells of the seminiferous epithelium) play a key role in regulation of spermatogenesis and altering rates of spermatozoa produced (Johnson et al., 1984b; Orth et al., 1988; Sharpe, 1994; Rato et al., 2012). Damages to Sertoli cells lead to severe disruption of spermatogenesis and male infertility (Sridharan et al., 2007; Papaioannou et al., 2009). Sertoli cells are the target for both the pituitary follicle stimulating hormone and testosterone (almost exclusively produced by Leydig cells), illustrating their crucial role in supporting germ cell maturation (Sharpe, 1994; Walker and Cheng, 2005; Walker, 2009). Sertoli cells provide a specialized, protected environment within the seminiferous tubules of the testis for germ cell development. Adjacent Sertoli cells form tight junctions with each other such that nothing larger than 1000 daltons can pass from the outside to the inside of the tubule. This characteristic of Sertoli cells creates what is known as the blood-testis barrier. Serum macromolecules are effectively excluded from the adluminal section, which is an essential prerequisite for spermatogenesis, creating a microenvironment consisting exclusively of Sertoli cell secretions and germ cells (Dym and Fawcett, 1970; Russell and Peterson, 1985; Griswold, 1995). At the beginning of meiosis, germ cells located outside of the barrier pass through the tight junctions. Once beyond the blood-testis barrier, germ cells are dependent on Sertoli cells to supply nutrients and growth factors (Mruk and Cheng, 2004). Sertoli cells provide factors required to fuel germ cell metabolism (lactate, transferrin, androgen binding protein), growth regulatory factors (stem cell factor, transforming growth factors alpha and beta, insulin-like growth factor-I (IGF-I), fibroblast growth factor (FGF) and epidermal growth factor (EGF) and hormones that regulate the development of the male reproductive structures or feedback to regulate the hormonal signals affecting Sertoli cells (mullerian-inhibiting substance (MIS), and inhibin) (Jegou, 1992; Skinner, 2005). Sertoli cells are very active in elongation and desaturation of polyunsaturated fatty acids (PUFAs), while isolated germ cells are inefficient in synthesizing 22:5n-6 and 22:6n-3 (Retterstøl et al., 2001a; b). Thus, Sertoli cells appear to provide the bulk of PUFAs to pachytene spermatocytes and round spermatids (Retterstøl et al., 2001a; b). Glutathione levels (GSH) are a marker of oxidative stress susceptibility. GSH participates in diverse biochemical processes, including the maintenance of protein thiol groups, and protection of cells against peroxides, free radicals, and certain xenobiotics (Meister and Anderson, 1983). Sertoli cells are essential for the maintenance of spermatogenic cell GSH. The rate of GSH synthesis in isolated spermatogenic cells is insufficient to compensate for GSH turnover (Li et al., 1989). Sertoli cell functions include providing structural support and nutrition to developing germ cells, phagocytosis of degenerating germ cells and residual bodies, release of spermatids at spermiation and production of a host of proteins that regulate and/or respond to pituitary hormone release and that influence mitotic activity of spermatogonia (Amann, 1970; Dym and Raj, 1977; Feig et al., 1980; Jutte et al., 1982, 1983; Tres et al., 1986; Buch et al., 1988; Bellve and Zheng, 1989; Johnson, 1991; Russell and Griswold, 1993; Sharpe, 1994; Griswold, 1998; Walker and Cheng, 2005; Johnson et al., 2008).

In a germ cell’s path to make a spermatozoon from a spermatogonium, a spermatogonium divides by mitosis in the basal compartment of the seminiferous tubule, to produce either stem cells or committed spermatogonia that ultimately become primary spermatocytes. These cells pass through the blood–testis barrier of the Sertoli cell tight junctions as they move into the adluminal compartment. Germ cells continue their development in the immunologic-protected site of the adluminal compartment. In mammals, numerous (9 to 11) cell divisions of spermatogonia and 2 divisions of spermatocytes build the germ cell population of the testis. Spermatocytes undergo two meiotic divisions to form haploid spermatids. During meiosis, spermatocytes undergo chromosomal pairing, synapsis, and genetic exchange as well as transforming into haploid cells. Each cell progresses through a series of cytologically identifiable stages known as leptotene (chromosomes begin to condense), zygotene (pairing of homologous chromosomes is initiated), pachytene (complete synapsis of homologous chromosomes which undergo recombination), diplotene (homologous chromosomes begin to separate), diakinesis (chromosomes condense), metaphase I (separation of homologous chromosomes) and metaphase II (separation of sister chromatids). The germ cells at the end of meiosis produce round spermatids (de Kretser and Kerr, 1988; 1994; Holstein et al., 2003). During spermiogenesis, the nucleus of round haploid spermatids undergoes major chromatin restructuring (Meistrich, 1989; Wouters-Tyrou et al., 1998). The somatic and meiotic histones that remain in spermatids are replaced ~14 days before ejaculation in the mouse (~21 days in human) with basic transition proteins and then with protamines (Meistrich, 1989; Churikov et al., 2004; Kimmins and Sassone-Corsi, 2005), which are arginine-rich proteins that condense the chromatin and cause DNA to become transcriptionally inactive (Kierszenbaum and Tres, 1978). Sperm are then released into the seminiferous lumen and undergo a final process of maturation within the epididymis where they acquire motility and the ability to fertilize the egg (Khole, 2003; Gatti et al., 2004). The result of these elaborate processes is transcriptionally inert mature sperm DNA, the most highly condensed DNA structure known. During spermatogenesis, loss of a certain percentage of germ cells via apoptosis is normal in all species investigated, and it plays a critical role in determining spermatogenic efficiency (Roosen-Runge, 1973; Sharpe, 1994). Germ cell death occurs exclusively or preferentially in certain developmental stages, also varying in a species-specific manner in quality and quantity (Roosen-Runge, 1977). It takes approximately 35 days in mice (~64 days in humans) for germ cells to develop from spermatogonia to spermatozoa. The last round of DNA synthesis occurs in preleptotene spermatocytes. Meiotic prophase lasts about two weeks in the mouse (less than 4 weeks in man) and is followed by the first and second meiotic divisions, which occur within 24 hr of each other, while spermiogenesis takes about three weeks (over 5 weeks in man). It has been calculated that, assuming that a man produces 100 million (108) mature sperm per day, during an average reproductive life of sixty years he would produce well over two trillion (2 x 1012) mature sperm in his lifetime (Martin, 1991) (and a multiple of this figure of immature germ cells).

6.2.2 Female gametogenesis

The mature oocyte is one of the largest cell types in most eukaryotic species. While the nucleus of an oocyte contains the same amount of DNA as any other post-replication diploid cell in an organism, the cytoplasm is larger in volume and often contains much more RNA and protein than is found in most somatic cells. The mature oocyte is prepared to sustain the early development of the embryo immediately following fertilization. In some species the embryo develops externally, e.g. in frogs or sea urchins. Thus the oocyte must contain sufficient yolk reserves to sustain the embryo until it begins feeding. In the mouse, the embryo develops internally, is nourished by the mother and little or no yolk is stored in the oocyte (Bakken and McClanahan, 1978). In most teleosts, all amphibians, most reptiles and relatively few mammals, oogenesis appears to continue either uninterruptedly or cyclically throughout reproductive life. The other variant is that the oogenesis occurs only in fetal gonads, and oogonia neither persist nor divide mitotically during sexual maturity - e.g. cyclostomes, elasmobranchs, a few teleosts, perhaps some reptiles, all birds, mono-tremes, and with a few possible exceptions, all eutherian mammals (Zuckerman, 1951; 1971; Franchi et al., 1962). Others maintain that neo-oogenesis and follicular renewal during the prime reproductive period exists throughout the animal kingdom, including higher vertebrates (Bukovsky et al., 2005). In the developing mouse fetus, primordial germ cells (PGCs) that have colonized the developing ovaries (the germ cells are now referred to as oogonia) continue to proliferate and thus expand the female germline. Peak numbers of germ cells are reached by around e13.5, at which time many of the oogonia have left the mitotic cell cycle and initiated the first steps of meiosis (Byskov, 1986). After passing through the leptotene, zygotene, pachytene, and diplotene stages of meiosis I, the oocytes become arrested in diakinesis coincident with their enclosure by a specialized population of somatic (pregranulosa) cells to form nongrowing primordial follicles (Byskov, 1986). An identical situation is believed to proceed in human fetal ovaries (Briggs et al., 1999), with peak germ cell numbers of around 7 x 106 reached by approximately the 20th week of gestation (Block, 1952; Baker, 1963; Tilly and Ratts, 1996). In humans, it has been estimated that the ~7 million germ cells in the fetal ovaries at around week 20 of gestation are decimated to 1–2 million viable oocytes in early neonatal life (Baker, 1963). A proportion of oocytes and follicles either do not become, or do not stay, arrested (probably by escaping local inhibitors, Wandji et al., 1997; Yang and Fortune, 2008), or they initiate growth during prenatal or prepubertal life. These growing follicles are a consistent feature of ovaries during childhood (Peters et al., 1976). They always become atretic due to the prevailing immaturity of the hypothalamo-pituitary axis (Djahanbakhch et al., 2007). By menarche, the time of in-principle commencement of reproductive competence, the number of primordial follicles is down to about 300,000 (Block, 1952; Faddy et al., 1999; Jansen and Burton, 2004). Comparable proportions of germ cells are lost during fetal ovarian development in other vertebrate species (Beaumont and Mandl, 1961). In all vertebrate species examined to date, females are born with far fewer oocytes than the peak numbers present during early oogenesis. In fact, it has been estimated that over two-thirds of the potential female germ cell pool is lost by the time of birth in rats (Beaumont and Mandl, 1961), mice (Borum, 1961; Bakken and McClanahan, 1978), and humans (Pinkerton et al., 1961; Baker, 1963; Forabosco et al., 1991), resulting from attrition of both oogonia and oocytes (Borum, 1961; Bakken and McClanahan, 1978). In the follicular phase of the reproductive cycle, a cohort of primordial follicles are recruited for further development and maturation, with the end result being ovulation of one or more dominant mature follicles. Folliculogenesis is a process starting from primordial follicles, through primary, secondary, and antral stages to the largest Graafian or preovulatory follicles. Less than 0.1% of follicles present at the beginning of puberty develop into follicles capable of being ovulated (Tilly et al., 1997; Marti et al., 1999). It has been estimated that the duration of time required for the growth of a human follicle from the primordial stage to the large preantral stage takes in excess of 150 days (Gougeon, 2004). Thus, a follicle which ovulates in any given menstrual cycle will actually have begun to grow at least five menstrual cycles earlier. At ovulation, the human egg has entered its second meiotic division. The secondary oocyte just before fertilization lies within a comparatively massive, cystic structure, approximately 2 cm in diameter, the preovulatory Graafian follicle. This follicle is composed of up to about one million follicular (granulosa) cells, a fluid-filled antrum that accounts for most of the follicle’s volume, and an eccentrically placed egg that is surrounded by a specialized group of closely applied granulosa cells, the cumulus oophorus. In response to the luteinizing hormone surge, the preovulatory follicles ovulate and release oocytes. After ovulation, the granulosa cells and theca cells remaining within the follicles differentiate into a transient endocrine gland, the corpus luteum (Murphy, 2000). The corpus luteum is critical for successful maintenance of pregnancy in mammals because it is the primary source of the progestational hormone progesterone. Granulosa cells play a nurturing role in supporting oocyte development and follicle maturation by providing essential nutrients and estrogen (Buccione et al., 1990; Binelli and Murphy, 2010). However, during follicular growth and development, more than 99% of follicles become atretic and regress, primarily due to apoptosis of granulosa cells (Baker, 1963; Hughes and Gorospe, 1991; Tilly et al., 1991; Faddy et al., 1992; Kaipia and Hsueh, 1997; Jiang JY et al., 2003). Thus, in each cycle of folliculogenesis most primordial follicles undergo a process of regression known as atresia and only one or a few (depending on the species) numbers of them are destined to become preovulatory follicles (McGee and Hsueh, 2000). Considering that a woman will generally ovulate a single egg during each monthly cycle, between 400 and 500 oocytes advance to the point of ovulation in adult life. In summary, of the 7 million or so oocytes formed in human fetal ovaries, fewer than 500 will ever ovulate. A quick calculation thus reveals that greater than 99.99% of fetal germ cells and 99.9% of the oocytes endowed in the human ovaries at birth face death as their ultimate fate, a process which, when completed, leads to exhaustion of the follicle reserve and to menopause (Morita and Tilly, 1999; Tilly, 2001). Follicular atresia is also a common feature of nonmammalian vertebrate ovaries although the mode of atresia may differ in different species (Saidapur, 1978).

If there is any meaning to evolutionary theory and to MacArthur and Wilson’s notion (1967, p. 149): “Evolution … favours efficiency of conversion of food into offspring”. this “waste” female and male gamete overproduction must bear an evolutionary meaning, since otherwise these organisms should have been replaced by organisms that deal with their resources more economically. In a world of limited resources (Heininger, 2012), “The best competitor in an environment will be that population that is able to persist using the lowest level of resource supply” (Sears et al., 2004).

7. Sex: Proactive mutagenesis


It would be very strange indeed to believe that everything in the living world is the product of evolution except one thing  the process of generating new variation!

Jablonka and Lamb, 2005 p.101

Summary

The environmental signals that resulted in the reactive bacterial transformation were both accommodated and assimilated in eukaryotes as proactive bet-hedging tools. The oxidative stress was “domesticated” by higher taxa as “fire in the hearth” to institutionalize the generation of genetic variation. To this end both male and female gametogenesis proceed in a stressful milieu. The poor vascularization of the testes and the maturing follicle together with the high metabolic investment particularly for spermatogenesis result in functional hypoxia and metabolic and oxidative stress. Hypoxia-inducible factors are constitutionally expressed in male and female germ cells. Various molecular and structural features such as cytokines and nuclear factor-kappaB, heat shock proteins, germ granules, mitochondrial uncoupling and aquaporin-8, and upregulated, but poorly coordinated DNA repair are footprints of the (epi-)mutagenic, DNA instability-fostering milieu. Importantly, germ granules are the germline-specific version of stress granules, the stress-dependent assembly of ribonucleoproteins. In essence, sexual reproduction evolved a gender-dimorphic “division of labor”. The smaller male gametes are exposed to high oxidative stress resulting in male-biased (epi-)mutagenesis and innovation of genetic information. The larger female gametes, affording a higher investment of resources, rather are the elements of information conservation. Germline cell differentiation is controlled by a specific set of genes whose expression is tightly locked into the repressed state in somatic cells. Large-scale epigenome alterations, now evidenced in nearly all cancers, lead to aberrant activation of these normally silenced genes, known as cancer/testis genes. The observation of shared characteristics between germline cells and tumor cells has led to the concept that recapitulation of portions of the germline gene-expression programme might contribute characteristic features to the neoplastic phenotype, including immortality and DNA instability.

Evolution is opportunistic and favors the prepared (Caporale, 1999). Due to its blindness and purposelessness, evolution does not "know" in advance which evolutionary path will lead to the increase of fitness (or at least avoid the loss of fitness) in fluctuating often unpredictable environments. Therefore, retrospectively, the best "strategy" to increase fitness is to take every possible path at every next step, sampling the ‘sequence space’ with as many lottery tickets as is reasonable in a bet-hedging strategy. As a result, no fit configurations will be missed. In contrast, if the evolutionary entity only takes a part of paths at the current location, only configurations downstream of these paths can be reached: all other configurations will be missed. From the angle of blind evolution the greater the incompleteness in configuration sampling, the smaller is the probability for blind evolution to increase fitness (Fu, 2007). Evidently, bacteria respond to stress resorting to a bet-hedging strategy, seeking new genetic material that might help them to survive (Johnsborg and Håvarstein, 2009). Three main bet-hedging strategies have been described: conservative bet-hedging (play it safe), diversified bet-hedging (don't put all eggs in one basket) and adaptive coin flipping (choose a strategy at random from a fixed distribution) (Cohen, 1966; Cooper and Kaplan, 1982, Seger and Brockmann, 1987; Olofsson et al., 2009). Bet-hedging is found in organisms ranging from bacteria to humans (Cohen, 1966; Gillespie, 1973; 1974a; Slatkin, 1974; Tonegawa, 1983; Hairston and Munns, 1984; Seger and Brockmann, 1987; Moxon et al., 1994; Danforth, 1999; Meyers and Bull, 2002; Friedenberg, 2003; Balaban et al., 2004; Kussell and Leibler, 2005; Wolf et al., 2005; Venable, 2007; Acar et al., 2008; Ackermann et al., 2008; Beaumont et al., 2009; Olofsson et al., 2009; Childs et al., 2010; Gremer et al., 2012; Morrongiello et al., 2012; Starrfelt and Kokko, 2012).

7.1 Sexual reproduction: domesticating the fire

It has been argued that exchange of genetic material can only speed up evolution if donors and recipients use the same system to encode, store and process genetic information. Consequently, prokaryotic ‘‘sex’’ must have played a significant role in preserving the near universality of the genetic code (Johnsborg et al., 2007). Exchange of genetic materials by two individual members of the same species is considered to be the origin of primitive sex. Genetic assimilation occurs when an acquired trait loses its dependency on environmental triggers and becomes an inherited trait (Masel, 2004). Already at the level of bacteria, genetic assimilation of competence and transformation is observed (Majewski, 2001; Birdsell and Wills, 2003). Neisseria gonorrhoeae, Helicobacter pylori, and Thermus thermophilus are some of at least 44 species of bacteria that are naturally competent for genetic transformation and are able to take up DNA independently of their growth phase (Sparling, 1966; Hidaka et al., 1994; Lorenz and Wackernagel, 1994; Israel et al., 2000; César et al., 2011). Another Gram-negative bacterium, Haemophilus influenzae becomes competent under defined physiological conditions, but is not sensitive to external signaling (Lorenz and Wackernagel, 1994). Transformation in the pathogenic Neisseria has fuelled high rates of recombination (Smith et al., 1993) that has been estimated to change an allele of the Neisseria meningitides genome ten times more likely than point mutation (Feil et al., 1999; Jolley et al., 2005). In gonococci, transformation has been harnessed as a powerful mechanism for generating genetic diversity, spreading advantageous alleles and mediating some forms of antigenic variation (Hobbs et al., 1994; Fudyk et al., 1999; Snyder et al., 2004). Analyses of the four published neisserial genomes revealed high densities of repeated elements (Parkhill et al., 2000; Tettelin et al., 2000; Achaz et al., 2002; Bentley et al., 2007). Intrachromosomal recombination between these repeats is a major source of variability in Neisseria, resulting in frequent adaptive changes in gene expression profiles (Moxon et al., 1994; Saunders et al., 2000) and even reoccurring states of hypermutability (Richardson et al., 2002; Davidsen et al., 2007). Transformation in Neisseria spp. and H. influenza requires the presence of a specific DNA uptake sequence (Scocca et al., 1974; Danner et al., 1980; 1982; Goodman and Scocca, 1988; Elkins et al., 1991) in the incoming DNA. These signals allow discrimination between DNA from closely related strains or species and foreign/unrelated DNA. When exposed to a mixture of homologous and foreign DNAs, these human pathogens show preferential uptake of DNA uptake sequence-containing DNA (Scocca et al., 1974; Elkins et al., 1991; Duffin and Seifert, 2010). In Neisseria meningitidis, Neisseria gonorrhoeae and Haemophilus influenza by far the most frequent 9– or 10mer repeat genomic sequences residing within coding regions are the DNA uptake sequences required for natural genetic transformation. A significantly higher density of DNA uptake sequences was found within genes involved in DNA repair, recombination, restriction modification and replication than in any other annotated gene group in these organisms (Davidsen et al., 2004). Increased DNA uptake sequences density is expected to enhance DNA uptake and the over-representation of DNA uptake sequences in genome maintenance genes might reflect facilitated recovery of genome preserving functions. For example, transient and beneficial increase in genome instability can be allowed during pathogenesis simply through loss of antimutator genes, since these DNA uptake sequences-containing sequences will be preferentially recovered. Furthermore, uptake of such genes could provide a mechanism for facilitated recovery from DNA damage after genotoxic stress (Davidsen et al., 2004). Evidence was also presented that DNA uptake sequences are implicated in genome stability rather than in generating adaptive variation (Treangen et al., 2008). This dual evidence for a role of DNA uptake sequences in DNA repair and diversity may reflect the ambiguity of sexual reproduction in generating both genetic conservation and innovation depending on the evolutionary demands of an organism (Bedau and Packard, 2003; Buchanan et al., 2004; Clune et al., 2008; Dees and Bahar, 2010). During evolution, this primitive form of molecular sex has been transformed into a complex biological function involving specialized sexual structures and multiple hormonal interactions (Roy et al., 1996). While genetic recombination is a key feature of meiosis, it is not unique to this process. Recombinational capacity is found throughout the prokaryotes and therefore must considerably predate eukaryotes and meiosis (Levin, 1988; Cavalier-Smith, 2002; Marcon and Moens, 2005; Wilkins and Holliday, 2009). A crucial set of molecules for genetic recombination, the recA family of proteins, is utilized for recombination in both prokaryotes and eukaryotes (Shinohara et al., 1992; Wilkins and Holliday, 2009).

Genetic assimilation is the evolutionary process by which a phenotype that is produced specifically in response to some environmental stimulus, such as a stressor, becomes stably expressed independently of the evoking environmental effect (Waddington, 1942; 1953a; 1956; 1957; Scharloo, 1991; Masel, 2004; Braendle and Flatt, 2006; Pigliucci et al., 2006). Genetic assimilation is a special case of a more general phenomenon, called genetic accommodation (West-Eberhard, 2003; Braendle and Flatt, 2006). This scenario of phenotypic evolution posits that (1) a mutation or environmental change triggers the expression of a novel, heritable phenotypic variant, (2) the initially rare variant phenotype starts to spread (in the case of an environmentally induced change, due to the consistent recurrence of the environmental factor), creating a subpopulation expressing the novel trait, and (3) selection on existing genetic variation for the regulation or form of the trait causes it to become (a) genetically fixed or to remain (b) phenotypically plastic (West-Eberhard, 2003). According to Braendle and Flatt (2006), only process (3) represents genetic accommodation in the strict sense as it was defined by West-Eberhard (2003) but, for the sake of conceptual simplicity, they refer to genetic accommodation as the entire sequence of steps (1) to (3). Genetic assimilation describes only scenario (3a), i.e. the fixation of the response leading to environmental insensitivity, also called ‘‘environmental canalization’’ (West-Eberhard, 2003), whereas genetic accommodation can describe both the evolution of environmentally insensitive (3a) and sensitive (3b) trait expression (Braendle and Flatt, 2006). Another difference between the two concepts is that the model of genetic accommodation assumes that the trigger uncovering previously cryptic or novel phenotypes is either genetic or environmental, whereas the concept of genetic assimilation typically assumes only an environmental trigger. Thus, genetic accommodation is a generalization of genetic assimilation (Braendle and Flatt, 2006). Sexual trait expression can be both environmentally insensitive, i.e. genetically assimilated, as in birds and mammals, and environmentally sensitive, i.e. genetically accommodated, as in cyclically parthenogenetic taxa like daphnia and aphids and geographically parthenogenetic taxa.

Co-option, exaptation, and preadaptation are related terms referring to shifts in the function of a trait during evolution. For example, a trait can evolve because it served one particular function, but subsequently it may come to serve another. Time and again evolution had to master an arduous task: what had evolved as a facultative response of unicellular organisms to environmental cues had to be fixed after the multicellular transition in an obligate developmental context (Heininger, 2001; 2002; 2012). What had been an environmental challenge evolved to be relayed by signaling molecules as a cellular message in the internal milieu of a multicellular organism.

In the same way fire is dangerous and humans learned how to control and use it, cells control and use ROS (de Magalhães and Church, 2006). Sexual reproduction evolved as stress response and the “wildfire” that had served to generate genetic diversity was “tamed” as “domestic fire in the hearth” to constitutively generate genetic variation. What evolved as reactive process became a proactive evolutionary motor. Importantly, by no means this implies that evolution may be foresighted.

7.2 The oxidative stress of gametogenesis

That oxidative stress represents an essential feature of gametogenesis is widely acknowledged (Riley and Behrman, 1991; Aitken, 1995; Tilly and Tilly, 1995; Chainy et al., 1997; Fisher and Aitken, 1997; Kugu et al., 1998; Knapen et al., 1999; Behrman et al., 2001; Gil-Guzman et al., 2001; Orozco et al., 2003; Ford, 2004; Juan et al., 2005; Aitken and Roman, 2008; Gupta et al., 2008). However, so far the focus of scientific interest was rather on the detrimental, pathophysiological, aspects of oxidative stress, particularly in mammalian gametogenesis, largely ignoring its physiological implications. Oxidative stress is an inherent feature of gametogenesis in all taxa. Importantly, from lower to higher taxa there is a substantially incremental use of this general principle, culminating in human male gametogenesis that balances at the verge of mutational error catastrophe (see chapter 14.1). Scientific focus has been on mammalian gametogenesis. Thus, evidence for the relevance of oxidative stress in gametogenesis is most compelling in mammals while it is often rather circumstantial in other taxa. One of the main difficulties relates to the absence of generally agreed methodologies to measure ROS, especially in vivo (Taylor and Moncada, 2010). But the fire develops enough smoke to be detected indirectly. For instance, DNA and RNA oxidation as indicators of cellular damage from free radicals (Halliwell and Gutteridge, 1999; Joyner-Matos et al., 2007), and up-regulated expression of specific stress proteins, such as heat shock proteins, are biomarkers of a homeostatic cellular response to a stressor (e.g. Hofmann and Somero, 1995; Downs et al., 2001; Abele and Puntarulo, 2004; Joyner-Matos et al., 2007).

A multitude of cellular stress-related features witnesses the pronounced developmental stress during gametogenesis:

  • metabolic and replicative stress
  • functional hypoxia with constitutive expression of hypoxia-inducible factors
  • oxidative stress markers
  • expression of heat shock factors and heat shock proteins
  • RNA granules
  • epigenetic reprogramming
  • recombination requiring double-strand breaks
  • membrane lipid unsaturation
  • mitochondrial aquaporin-8
  • DNA repair
  • autophagy and apoptosis

During mammalian gametogenesis, at least 3 major periods with increased oxidative stress can be delimited:

       • period of epigenetic reprogramming in PGCs and ensuing quality selection
       • recombination during meiosis requiring double strand breaks
       • sperm maturation or follicle selection

7.2.1 Functional hypoxia

7.2.1.1 Testis

Whereas ambient air contains 21% O2, most tissues maintain O2 tensions between 2% and 9%. The resistance in the unusually long and narrow testicular artery is high, leaving pressure in the unfenestrated testicular capillaries lower than in all other organs, only marginally higher than venous pressure (Sweeney et al., 1991). Since blood vessels are located exclusively between the tubuli, oxygen reaches the lumen of the tubuli seminiferi only by diffusion. The poor vascularization of the testes means that oxygen tensions in this tissue are low (Cross and Silver, 1962; Setchell and Waites, 1964; Free et al., 1976; Setchell, 1978, p. 300; Setchell et al., 1994; Klotz et al., 1996; Zheng and Olive, 1997; Lysiak et al., 2000a; b; Giaccia et al., 2004; Wenger and Katschinski, 2005; Reyes et al., 2012). The safety margin for disturbances in the testicular blood flow is particularly narrow (Damber and Bergh, 1992; Bergh et al., 2001; Lissbrant et al., 2006). The capacity of the testis to autoregulate its blood flow is limited (Lissbrant et al., 2006), and reductions in systemic blood pressure are, therefore, accompanied by reductions in testicular blood flow (Free, 1977; Lissbrant et al., 1997a). In line with this, circulatory shock is followed by damage to the spermatogenetic epithelium (Pfitzer et al., 1982). Reported mean interstitial oxygen tensions in mammalian testes ranged from 10.6 to 15.2 mmHg (Cross and Silver, 1962; Massie et al., 1969; Free et al., 1976; Lysiak et al., 2000a), close to the brink of hypoxia (Setchell, 1978; Collin et al., 2000). Even pO2 values as low as 2 mmHg have been reported, which are among the lowest values found in the body and otherwise occur only in the vicinity of oxygen consuming mitochondria (Max, 1992). Even if other groups reported higher testicular pO2 values, the pO2 values within the tubuli seminiferi are clearly below the pO2 values outside of the tubuli (Wenger and Katschinski, 2005). Moreover, testicular microvascular pO2 (which represents the driving force for transcapillary O2 flux; Behnke et al., 2001) was found reduced by ~50% with old age (Dominguez et al., 2011). As an adjustment of mammalian testes to low O2 pressure in the seminiferous tubules, testicular mitochondria seem to consume less oxygen to generate the same electrical membrane potential when compared to mitochondria from other tissues (Moreira et al., 2005; 2006; Teodoro J et al., 2006). Cytochrome c oxidase is the terminal complex of the mitochondrial respiratory chain responsible for about 90% of oxygen consumption in mammals (Babcock and Wikström, 1992). Inhibition by NO of O2 binding to cytochrome c oxidase (Brown and Cooper, 1994; Cleeter et al., 1994; Schweizer and Richter, 1994; Brown, 2001; Cooper et al., 2008; Taylor and Moncada, 2010) might be responsible for the inability of mitochondria to consume O2 readily at low O2 concentrations (Clementi et al., 1999). Inhibition of cytochrome c oxidase has been linked to increased ROS release (Kowaltowski and Vercesi, 1999; Echtay et al., 2002; Inoue et al., 2003; Turrens, 2003; Facundo et al., 2006; Kowaltowski et al., 2009). Moreover, endogenous NO mediates ROS production at low oxygen concentrations by modifying the redox state of cytochrome c oxidase (Moncada and Erusalimsky, 2002; Palacios-Callender et al., 2004; Mason et al., 2006). NO production in rat liver and heart mitochondria increases under hypoxic conditions (Schild et al., 2003; Valdez et al., 2004). Mitochondria from a variety of eukaryotes are capable of reducing NO2 to NO when incubated at low oxygen concentrations (Kozlov et al., 1999; Nohl et al., 2000; Tiravanti et al., 2004; Tischner et al., 2004; Planchet et al., 2005; Castello et al., 2006) and this reaction has been shown to be catalyzed by cytochrome c oxidase in a pH-dependent fashion (Castello et al., 2006; 2008; Poyton et al., 2009a). In the testis, NO regulates various functions, including Leydig cell steroidogenesis (Del Punta et al., 1996), spermatogenesis and germ cell apoptosis (Zini et al., 1996; El-Gohary et al., 1999; Lue et al., 2003), Sertoli cell tight-junction dynamics (Lee and Cheng, 2004) and regulation of testicular blood flow (Lissbrant et al., 1997b). In addition to its mitochondrial generation, NO is synthesized from L-arginine by three different nitric oxide synthases: inducible NOS (iNOS), constitutive neuronal nitric oxide synthase (nNOS) and constitutive endothelial nitric oxide synthase (eNOS). Endothelial and neuronal NOS are Ca2+-dependent and often produce NO at the nanomolar level every few hours. Inducible NOS produces more NO (micromolar levels) and its productive activity can last for days. In the mammalian testis, all three isoforms of NOS are constitutively expressed and they play important roles in the biology of Sertoli and Leydig cells as well as in spermatogenesis and germ cell apoptosis (Zini et al., 1996; O’Bryan et al., 2000; Wang Y et al., 2002; Ha et al., 2004; Kim HC et al., 2007; Costur et al., 2012; Doshi et al., 2012). It has been shown that iNOS can be induced by factors released from round spermatids, implicating a regulatory role of germ cells on Sertoli and Leydig cell NOS function (Lee and Cheng, 2004; Türker et al., 2004).

Testicular blood flow exhibits vasomotion, the rhythmic dilation and constriction of precapillary sphincters, which in turn results in cyclical variations in blood flow through capillaries (Damber et al., 1982; 1983; Collin et al., 1993; 2000; Turner et al., 1996). Mean rat testicular interstitial pO2 was 12.5 ± 2.6 mm Hg, which displayed a cyclical variation of 11.9 ± 0.4 cycles per minute with a mean amplitude of 2.8 ± 0.8 mm Hg (Lysiak et al., 2000a). Vasomotion is testosterone-dependent (Damber et al., 1987; 1992; Collin et al., 1993), is induced by local factors (Bergh et al., 1999), is not seen prior to puberty (Damber et al, 1990) and is inhibited by stress, hyperthermia, hypoxia, cryptorchidism, and varicocele (Setchell et al., 1995; Collin et al., 1996; 2000; Collin and Bergh, 1996b). There is evidence for an increase in oxidative stress as a consequence of cycling hypoxia in tumors (Kalliomäki et al., 2008).

Oxygen consumption in the testis is high because of the energetic demands of spermatogenesis (Setchell and Waites, 1964; Wenger and Katschinski, 2005). Hypoxia occurs when the metabolic demand for oxygen exceeds the supply and both low oxygen tension and high oxygen consumption together result in functional hypoxia (Abele et al., 2007). Cells detect decreases in oxygen concentrations to activate a variety of responses that help cells to adapt to low oxygen levels. Oxidative stress may be induced not only by a rise but also by a fall in oxygen tension. Paradoxically, hypoxic tissues generate a high amount of oxidative stress (Chandel et al., 1998; 2000; Waypa and Schumacker, 2002; Schumacker, 2002; Turrens, 2003; Guzy and Schumacker, 2006; Cash et al., 2007; Bell et al., 2007; Clanton, 2007; Kulkarni et al., 2007), at least in part mediated by NO (Palacios-Callender et al., 2004; Mason et al., 2006). Mitochondria have been implicated as potential oxygen sensors by increasing the generation of ROS, which regulate a variety of hypoxic responses (Chandel et al., 1998; 2000; Agani et al., 2000). Both in Saccharomyces cerevisiae (Dagsgaard et al., 2001) and mammals (Brunelle et al., 2005; Guzy et al., 2005; Klimova and Chandel, 2008) the mitochondrial respiratory chain is required for hypoxic gene expression independent of oxidative phosphorylation. The transcriptional response to hypoxia activates a microtubule-dependent and dynein motor-driven mechanism that redistributes mitochondria from the central cytoplasm to the perinuclear region (Gutsaeva et al., 2008; Liu and Hajnóczky, 2011; Al-Mehdi et al., 2012; Murphy, 2012). Perinuclear clustering of mitochondria is also a reponse of cells to oxidative stress (Hallmann et al., 2004). This perinuclear mitochondrial clustering is associated with accumulation of nuclear ROS that is required for hypoxia-induced transcription (Al-Mehdi et al., 2012; Murphy, 2012; Sena and Chandel, 2012). In a multitude of taxa throughout phylogenesis, clusters of perinuclear mitochondria have been identified during gametogenesis (Fukuda et al., 1975; Guraya, 1979; D'Herde et al., 1995; Eckelbarger et al., 1998; Dabiké and Preller, 1999; de Smedt et al., 2000; Pepling and Spradling, 2001; Wilding et al., 2001; Eckelbarger and Young, 2002; Kloc et al., 2004a; Pepling et al., 2007; Zelazowska et al., 2007; Taguchi et al., 2012) mimicking the perinuclear distribution of mitochondria of hypoxic tissues. Intriguingly, perinuclear redistribution of mitochondria is an early event of cell death pathways in many cell types (De Vos et al., 1998; Desagher and Martinou, 2000; Wakabayashi and Spodnik, 2000; Li J et al., 2004; Golstein and Kroemer, 2007), facilitating the transfer of apoptosis initiating factor (AIF), which activates caspases in the nucleus, Endonuclease G, an apoptotic DNase that degrades nuclear DNA, and ROS directly from the mitochondria to the nucleus to promote genomic destruction (Ferri and Kroemer, 2001; Li LY et al., 2001; Li J et al., 2004; Aslan and Thomas, 2009). Disrupted expression of beta-actin reduced TNF-induced mitochondria clustering, ROS production and apoptosis dramatically (Li J et al., 2004).

7.2.1.2 Ovary

The ovary is one of the best vascularized organs of the body (Banwell, 2009). However, the large pre-ovulatory follicle, where the oocyte undergoes maturation, remains avascular with all capillaries being excluded from the basal lamina and the oocyte is removed from any direct oxygen supply (Hazzard and Stouffer, 2000; Plendl, 2000; Tamanini and De Ambrogi, 2004; Banwell, 2009). Hypoxia of the granulosa cells is a normal event during the growth of ovarian follicles (Tropea et al., 2006). Follicular fluid oxygen partial pressure is reported to decrease with increasing follicular size in mammals (Fischer et al., 1992; Basini et al., 2004a). A positive correlation between glucose utilization and lactate production exists, and it is postulated that, as the follicle grows, energy requirements increase with decreasing O2 availability (due to thickening of the avascular epithelium), leading to an increase in glycolysis and increased lactate production (Boland et al., 1993; Gull et al., 1999). This is accompanied by a 2-fold decrease in O2 tension (59.8 mmHg in follicular fluid versus 102 mmHg in maternal blood) and higher CO2 tension (46.9 mmHg in follicular fluid versus 38.3 mmHg in blood), resulting in a lower pH of follicular fluid compared with blood (7.33 and 7.41, respectively) (Fischer et al., 1992; Sutton et al., 2003). Theoretical modeling also suggested that the follicle becomes increasingly hypoxic (Gosden and Byatt-Smith, 1986; Redding et al., 2007; 2008). Oxygen limitation is known to stimulate follicular angiogenesis, which is important for follicular growth and development. Impairment of angiogenesis within ovarian follicles contributes to follicular atresia (Greenwald and Terranova, 1988). ROS may act as signal transducers (Schroedl et al., 2002) or intracellular messengers (Pearlstein et al., 2002) of the angiogenic response.

7.2.2 Metabolic and replicative stress

Metabolic activity is a primary source of free radicals, which are unavoidable by-products of ATP synthesis. Commoner et al. (1954) published the first direct evidence that free radicals were produced in living cells and found that free radical levels were higher in tissues that were more metabolically active (Swartz, 1998; Gomez-Cabrera et al., 2009). Endotherms exhibit generally a higher metabolic activity (Else et al., 2004) and 10 to 100 fold higher ROS production in different tissues compared to invertebrates and fishes (Wilhelm Filho et al., 2000; 2007; Abele and Puntarulo, 2004), and this difference corresponds roughly to differences in specific metabolic rates (Schmidt-Nielsen, 1977). Typically, a cell becomes mitochondria-rich when it is subjected to high levels of metabolic demand (Lane, 2002). The ‘germinal cytoplasm’ of a variety of organisms is crowded with mitochondria (Czolowska, 1969; Fukuda et al., 1975; Guraya, 1979; D'Herde et al., 1995; Eckelbarger et al., 1998; Dabiké and Preller, 1999; de Smedt et al., 2000; Pepling and Spradling, 2001; Wilding et al., 2001; Eckelbarger and Young, 2002; Kloc et al., 2004a; Pepling et al., 2007; Zelazowska et al., 2007; Taguchi et al., 2012). An increased mass-specific metabolic rate elevates the generation of reactive oxygen and nitrogen species (RONS) both at the cellular (Sohal et al., 1990; Mohanty et al., 2000) and organismal level (Adelman et al., 1988). Mitochondrial ROS generation correlates well with metabolic rate (Sohal and Allen, 1995; Perez-Campo et al., 1998), suggesting that a faster metabolism simply results in more respiratory chain electron leakage. Conversely, dietary restriction results in a decrease in mitochondrial substrate oxidation activity, a concomitant decrease in the production rate of ROS and oxidative DNA damage (Simic and Bergtold, 1991; Chung et al., 1992; Sohal and Weindruch, 1996; Yu, 1996; Dandona et al., 2001; Heilbronn and Ravussin, 2003; Bevilacqua et al., 2004; Lambert and Merry, 2004; Masoro, 2005). Likewise, a dietary restriction mimetic, the glucose antimetabolite 2-deoxy-D-glucose, that competitively inhibits the uptake and utilization of glucose thereby suppressing glycolysis and reducing energy production in mitochondria, leads to a decrease in production of ROS (Halicka et al., 1995; Lee J et al., 1999a; Roth et al., 2005) and oxidative DNA damage (Tanaka et al., 2006a). Various intraindividual, intraspecies and interspecies comparative data revealed that the rate of ROS generation, oxidative DNA damage and DNA evolution is directly related to the mass-specific metabolic rate (Adelman et al., 1988; Shigenaga et al., 1989; Cutler, 1991; Simic and Bergtold, 1991; Avise et al., 1992a; Cortopassi et al., 1992; Adachi et al., 1993; Martin and Palumbi, 1993; Loft et al., 1994; Rand, 1994; Martin, 1999; Gillooly et al., 2001; 2005; 2007; Cooke et al., 2003; Foksinski et al., 2004; Olinski et al., 2006; Rosa et al., 2008).

Quite obviously, for sexually reproducing taxa reproduction affords a high investment for gamete production in terms of resource and energy utilization (Williams, 1966b; Calow, 1979; Bell, 1980; Reznick, 1985; Bell and Koufopanou, 1986; Berglund and Rosenqvist, 1986; Loudon and Racey, 1987; Gittleman and Thompson, 1988; Reiss, 1989; Geber et al., 1999; Rocheleau and Houle, 2001; Barnes and Partridge, 2003; Lester et al., 2004; Wheelwright and Logan, 2004; Roff et al., 2006; Speakman, 2008; Szymanski et al., 2009; Bergeron et al., 2011). In the oviparous lizard, Sceloporus undulatus, for instance, the metabolic rate of females when gravid was elevated by 122% compared with that when non-gravid (Angilletta and Sears, 1999). In birds, reproduction is associated with significant metabolic costs. Because most avian species maintain atrophied reproductive organs when not active, reproduction requires major tissue remodeling in preparation for breeding. Females undergo rapid (days) recrudescence and regression of their reproductive organs at each breeding attempt, while males grow their organs ahead of time at a much slower rate (weeks) and may maintain them at maximal size throughout the breeding season. Egg production leads to a 22%–27% increase in resting metabolic rate over non-reproductive values. In male birds, gonadal recrudescence may lead to a 30% increase in resting metabolic rate (Vézina and Salvante, 2010). In free-ranging reproductive male North American red squirrels (Tamiasciurus hudsonicus) mean energy expenditure of males approximately doubled during the breeding season (from 290 ± 7 to 579 ± 73 kJ/day) (Lane et al., 2010).

Testis-specific morphogenetic events suggest that male gonads have a higher energy requirement than ovaries (Matoba et al., 2008). Approximately 4×106 sperm cells are produced every day by the male mouse (Thayer et al., 2001). Between age 20 and 50 years, the human testes produce approx. 2,000 spermatozoa per second each day with wide ranges (Amann and Howards, 1980; Amann, 2010). For 50% of healthy men, 21–50 years old, their daily output of mature sperm is between 68 and 250 x 106 (i.e. 2.8 to 10.4 x 106 sperm/h), but for 25% of men daily sperm production is <68 x 106 and for another 25% of men is between 251 and >600 x 106 (Johnson et al., 1984a). In addition, almost half of the potential sperm production in men is lost by apoptosis during postprophase of meiosis (Johnson et al., 1983a) and around two-thirds to three-quarters of spermatogonia are eliminated during mitotic proliferation (Oakberg, 1956; Clermont, 1962; Huckins and Oakberg, 1978; Allan et al., 1992; Dym, 1994; Blanco-Rodríguez et al., 2003). Daily sperm production per gram of testicular parenchyma varies between species but appears to be generally high, and as far as comparisons can be carried out (Clermont, 1972; Berndtson, 1977; 2011), is significantly higher in other mammalian species than in humans (Amann et al., 1976; Berndtson, 1977; Johnson et al., 1980; 1983b; 1992; 2000; Johnson, 1986; Gopalkrishnan et al., 1987). Thus, given that daily output of mature sperm is only the peak of the iceberg of daily spermatogenetic activity, there is a huge energetic investment into mammalian male gametogenesis.It has been calculated that, assuming that a man produces 100 million (108) mature sperm per day, during an average reproductive life of sixty years he would produce well over two trillion (2 x 1012) mature sperm in his lifetime (Martin, 1991) (and a multiple of this figure of immature germ cells). In mice it has been estimated that if all the cells produced as spermatogonia were to become sperm, there would be a two- to fivefold increase in sperm production (de Rooij and Lok, 1987). Mammalian male germ cells produce ATP in the mitochondria at a close to maximal rate (Grootegoed et al., 1984). Spermatogenesis in vivo is impaired by lowered germ cell ATP levels after inhibition of the citric acid cycle and uncoupling of oxidative phosphorylation. Fluoroacetate is converted to fluorocitrate and then inhibits the enzyme aconitase. Gossypol acts as an uncoupling agent on oxidative phosphorylation in different cell types. The more or less specific effects of fluoroacetate and gossypol on spermatogenesis in vivo (Sullivan et al., 1979; Qian and Wang, 1984) may be related to a high sensitivity of spermatogenic cell types, as compared to other cell types, to compounds which interfere with mitochondrial energy metabolism and respiratory control (Grootegoed et al., 1984).

Oscillations in oxygen consumption, energy metabolism, and redox state are intimately integrated with cell cycle progression, establishing the redox control of the cell cycle (Menon and Goswami, 2007; Burhans and Heintz, 2009; Sarsour et al., 2009). Because signaling pathways play specific roles in different phases of the cell cycle and the hierarchy of redox-dependent regulatory checkpoints changes during cell cycle progression, the effects of ROS on cell fate vary during the cell cycle. ROS navigate cells between Scylla and Charybdis. A role for oxidative stress has been demonstrated in the stimulation of cell proliferation (Burdon, 1995; Irani et al., 1997; Cotgreave and Gerdes, 1998; Shackelford et al., 2000; Thannickal et al., 2000; Sauer et al., 2001; Ushio-Fukai et al., 2002; Kreuzer et al., 2003; Immenschuh and Baumgart-Vogt, 2005; Rhee, 2006; Buggisch et al., 2007; Matés et al., 2008). Regarding cellular proliferation, oxidative stress affects several biochemical pathways (from epidermal growth factor receptor to mTOR) that involve key signaling proteins, such as nuclear factor erythroid 2-related factor 2 (Nrf2), kelch-like protein 19 (Keap1), Ras, Raf, mitogen activated protein kinases (MAPK) such as ERK1/2, MEK, p38alpha, c-Jun N-terminal kinase (JNK), c-myc, p53 and PKC (Matsuzawa and Ichijo, 2008; Nguyen et al., 2009; Wiemer, 2011; Sosa et al., 2013). And oxidative stress is the key orchestrator of cellular deletion by apoptosis (Buttke and Sandstrom, 1994; Jacobson, 1996; Tan et al., 1998; Jabs, 1999; Kannan and Jain, 2000; Simon et al., 2000; Jones, 2001; Ueda et al., 2002; Kern and Kehrer, 2005; Le Bras et al., 2005; Orrenius, 2007; Matés et al., 2008; Trachootham et al., 2008; Circu and Aw, 2010; Doyle et al., 2010). Intriguingly, thyroid hormones that increase the metabolic rate, calorigenesis, and exacerbate oxidative stress due to the acceleration of aerobic metabolism, trigger reproductive activity in response to environmental cues as phylogenetically highly conserved signals (see chapter 14.2.2).

Mammalian spermatogenesis balances at a very narrow edge in relation to cell proliferation on one hand and lipid peroxidation, DNA damage and cell death on the other. Progression to a more prooxidant state whilst initially leading to enhanced proliferative responses results subsequently in increased cell death (Burdon, 1995). Increased mitochondrial metabolic activity has been shown to enhance oxidative stress (Commoner et al., 1954; Hagen et al., 1998; Swartz, 1998; Gomez-Cabrera et al., 2009). Accordingly, the high metabolic effort during the reproductive period is reflected by increased systemic prooxidant state, susceptibility to stress and oxidative stress (Boonstra et al., 2001; Salmon et al., 2001, Wang Y et al., 2001; Wingfield and Sapolsky, 2003; Alonso-Alvarez et al., 2004a; 2006; Koochmeshgi et al., 2004; Wiersma et al., 2004; Bertrand et al., 2006; Klose et al., 2006; Delaporte et al., 2007; Harshman and Zera, 2007; Rush et al., 2007; Samain et al., 2007; Bizé et al., 2008; Costantini, 2008; Dowling and Simmons, 2009; Monaghan et al., 2009; Bergeron et al., 2011; Isaksson et al., 2011; Heiss and Schoech, 2012).

Many observations have demonstrated the greater vulnerability of single-stranded DNA as occurs during DNA replication and transcription (Cohen et al., 1991). For example, in vitro single-stranded DNA exhibits a >100-fold greater rate of depurination (Lindahl and Neiberg, 1972) or of deamination (Lindahl and Neiberg, 1974) than double-stranded DNA. Similarly, a review of chemical mutagenesis that is often associated with oxidative stress stated that "Most, if not all, mutagens are much more reactive in single-stranded nucleic acids, so that these regions are probably preferentially modified in replicating DNA" (Singer and Kusmierek, 1982). The occurrence of nonrandom distribution/clustering of oxidative DNA damage sites in the genome of highly proliferative cells indicates that there may be regions in the genome with an increased vulnerability to ROS damage (Chastain et al., 2006; Redon et al., 2010). These vulnerable regions are in areas undergoing replication, transcription, and/or between nucleosomes (Hanawalt et al., 1979; Friedberg et al., 1995; MacLeod, 1995; Chastain et al., 2006). DNA synthesis is a remarkably vulnerable phase in the cell cycle. In addition to introduction of errors during semi-conservative replication, the inherently labile structure of the replication fork, as well as numerous pitfalls encountered in the course of fork progression, make the normally stable double stranded molecule susceptible to collapse and recombination. Estimates of the extent of endogenous DNA damage due to oxidants produced during metabolic activity vary widely (Nohl, 1994; Beckman and Ames, 1997; Vilenchik and Knudson, 2003; Barzilai and Yamamoto, 2004; Møller and Loft, 2004; Tanaka et al., 2006b). According to one of the relatively low estimates, during a single cell cycle approximately 5,000 DNA single-strand lesions are generated per nucleus by endogenous ROS (Vilenchik and Knudson, 2003; Tanaka et al., 2006b). Approximately 1% of these single-strand lesions become converted to DSBs, predominantly at the time of DNA replication, while the remaining 99% are repaired by essentially error-free mechanisms. Thus, on average, about 50 DSBs per nucleus (~0.8 DSBs per 108 bp) are generated during a single cell cycle in human cells (Vilenchik and Knudson, 2003). Cellular demand for DNA repair correlates with the cell’s potential to replicate. gamma-H2AX-enriched regions (that mark DSBs, see chapter 7.2.8) of endogenous origin in replicating cells include sub-telomeres and active transcription start sites, apparently reflecting replication- and transcription-mediated stress during rapid cell division (Seo et al., 2012). Activated oncogenes including ras, myc, cyclin E, mos, cdc25A, and E2F1, result in the continuous formation of mitochondrial ROS and endogenous DSBs (Denko et al., 1994; Karlsson et al., 2003; Bartkova et al., 2005; 2006; Di Micco et al., 2006; Halazonetis et al., 2008; Ralph et al., 2010; González et al., 2013) due to increased replication stress (Bartkova et al., 2006; Di Micco et al., 2006). Male spermatogonia and spermatocytes have an extreme replication activity compared with any other cell type (humans produce approximately 50–200 million spermatozoa per individual per day). This huge energetic investment, coupled with testicular high oxygen consumption and low oxygen tension generates substantial oxidative stress (Agarwal et al., 2003; Fujii et al., 2003) and associated DNA damage (Tanaka et al., 2006a; b).

7.2.3 Oxidative stress

Oxidative stress is a constitutive feature of male (Riley and Behrman, 1991; Aitken, 1994; 1995; Chainy et al., 1997; Fisher and Aitken, 1997; Gil-Guzman et al., 2001; Aitken et al., 2003; Orozco et al., 2003; Ford, 2004; Juan et al., 2005; Aitken and Roman, 2008) and female gametogenesis (Riley and Behrman, 1991; Tilly and Tilly, 1995, Kugu et al., 1998, Knapen et al., 1999; Behrman et al., 2001; Agarwal et al., 2005; Gupta et al., 2008; Ruder et al., 2008; Lázár, 2012). A thorough examination of the available literature suggests the general tendency that in female gametogenesis the physiological role of ROS is appreciated while in male gametogenesis rather the pathophysiological role of ROS is emphasized. Particularly, the oxidative stress within the seminiferous tubules is often treated as black box, in part because the massive ROS-dependent carnage of germ cells is highly cryptic due to the fast phagocytosis of cell debris by Sertoli cells (Maeda et al., 2002) and in part because oxidative stress during the various later stages of sperm maturation is highly visible and may have a detrimental impact for male fertility. On the other hand the high sensitivity of spermatogenetic cells to oxidative stress may explain their extremely limited viability in cell culture (Chapin and Phelps, 1990). Testosterone has been shown to have pro-oxidant properties in a variety of animal tissues (e.g. Chainy et al., 1997; Royle et al., 2001; Pansarasa et al., 2002; Aydilek et al., 2004; Gil et al., 2004; Rutkowska et al., 2005; Prasad et al., 2006; 2008; Alonso-Alvarez et al., 2007; Iliescu et al., 2007; Chignalia et al., 2012) while in other tissues, particularly in the brain, anti-oxidant actions have been described (Ahlbom et al., 1999; 2001; Calderón Guzmán et al., 2005; Chisu et al., 2006). Testosterone also induces testicular oxidative and nitrosative stress (Peltola et al., 1996; Chainy et al., 1997; Aydilek et al., 2004; Guo et al., 2009) that, on the other hand, is indispensable for spermatogenesis (Chainy et al., 1997; Aydilek et al., 2004). Androgens may cause oxidative stress via cytochrome P450 4A and NADPH oxidase-dependent mechanisms (Iliescu et al., 2007; Singh et al., 2007; Ojeda et al., 2012) that was considered causal to the sexual dimorphism of blood pressure dysregulation (Chen and Meng, 1991; Bowles, 2004; Dantas et al., 2004; Iliescu et al., 2007; Miller et al., 2007; Singh et al., 2007; Ojeda et al., 2012). Testosterone also stimulated NADPH-oxidase activity in a variety of tissues (Brown et al., 1976; Lu JP et al., 2010; Chignalia et al., 2012).

Oxidative and nitrosative stress, at least in part mediated by gonadal hormones (Chainy et al., 1997; Felty et al., 2005; Guo et al., 2009), is not only a hallmark of gametogenesis but of other sexual reproduction-related events as well (Riley and Behrman, 1991; Heininger, 2001; Agarwal et al., 2003; 2005; Lue et al., 2003; Nedelcu, 2005; Aitken and Roman, 2008; Metcalfe and Alonso-Alvarez, 2010; Lázár, 2012). In vitro, male germ cells at various stages of differentiation were found to spontaneously generate hydrogen peroxide as they progressed through the epididymis, maximal activity being observed on the release of mature cells from the caudal region into a modified Krebs-Ringer's solution. Superoxide production could be dramatically enhanced by the addition of exogenous NADPH, in a manner that was closely correlated with the stage of epididymal development being maximal for immature cells recovered from the caput epididymis in all species. Precursor germ cells (pachytene spermatocytes, round and elongate spermatids) similarly generated chemiluminescent signals compatible with the low level generation of ROS. Superoxide generation in these cells could again be stimulated by NADPH, via mechanisms that were inversely related to the stage of germ cell differentiation, the greatest activity being observed in pachytene spermatocytes. These results demonstrate that differentiating male germ cells have the potential to generate ROS, and have implications for the redox regulation of gonadal function and the development of reproductive pathologies involving oxidative stress (Fisher and Aitken, 1997).

As will be discussed in more detail in chapter 10, a multitude of processes related to gametogenesis require oxidative stress: mutagenesis, epimutagenesis, reprogramming, recombination, and apoptosis. Testicular germ cells are intimately associated with the free radical generating phagocytic Sertoli cells (Bauché et al., 1994). Compared to Sertoli cells, germline cells are much more susceptible to oxidative stress resulting in germ cell-specific apoptosis following oxidative stressor challenge (Turner et al., 1997; Lysiak et al., 2000a; Lue et al., 2002; Payabvash et al., 2008). A strong mutational bias from G/C to A/T nucleotides implicates oxidative DNA damage as a major endogenous mutagenic force in a variety of eukaryotes (Petrov and Hartl, 1999; Haddrill and Charlesworth, 2008; Lynch et al., 2008; Denver et al., 2009; Keightley et al., 2009; Lynch, 2010b; Ossowski et al., 2010). 5-Hydroxyuracil (resulting from the oxidative deamination of cytosine) and 8-oxoguanine (8-oxoG) are the two most common types of oxidative DNA damage in animal genomes and they cause G:CA:T transitions and G:CT:A transversions, respectively. 8-oxoG is unevenly distributed in the normal human genome and the distribution pattern is conserved among different individuals. Regions with a high frequency of recombination and SNPs are preferentially located within chromosomal regions with a high density of 8-oxoG (Ohno et al., 2006). This evidence suggests that 8-oxoG is one of the main causes of frequent recombinations and SNPs in the human genome, which largely contribute to the genomic diversity in human beings (Ohno et al., 2006). Management of ROS to prevent excessive damage, yet enabling its signaling function is achieved through numerous enzyme systems e.g. peroxidases, superoxide dismutases etc., and small molecules e.g. glutathione that collectively form the cellular anti-oxidant system (Hoogeboom and Burgering, 2009). The key role of glutathione in protecting various cells against free radical injury or chemically induced damage is now well established (Meister and Anderson, 1983; Vina, 1990; Sies, 1999).

The nuclear factor erythroid 2-related factor 2 (Nrf2), a Cap'n'Collar basic leucine zipper transcription factor, activates expression of cytoprotective genes that protect against oxidative stress (Banning et al., 2005; Kohle and Bock, 2007; Sporn and Liby, 2012; Cui et al., 2013). Nrf2 modulates the expression of hundreds of genes, including not only the familiar antioxidant enzymes but also a large number of genes that control several processes, including immune and inflammatory responses, tissue remodeling and fibrosis, carcinogenesis, and metastasis (Hybertson et al., 2011; Sosa et al., 2013). ROS levels are tightly controlled and predominantly regulated by Nrf2 and its repressor protein Keap1. When treated with oxidation-inducing drugs, mice that lack Nrf2 develop more severe intestinal inflammation than controls, with increased aberrant crypts, suggesting a function for Nrf2 in the prevention of inflammation and carcinogenesis (Khor et al., 2006). Similarly, Nrf2-deficient mice display an increased susceptibility to oxidative stress with a disrupted spermatogenesis in an age-dependent manner (Nakamura et al., 2010). A recent report described a strong association between functional polymorphisms in Nrf2 promoters with defective spermatogenesis in humans and lower Nrf2 mRNA expression and decreased levels of antioxidant gene glutathione S-transferaseM1 and SOD2 mRNA in spermatozoa (Yu et al., 2012).

In mammals, antioxidant enzymes such as Sod1, Sod2, and Gpx are present during different stages of oogenesis (El Mouatassim et al., 1999). All known peroxiredoxins are expressed in the oocytes; particularly, Prx6 is upregulated during in vitro maturation (Leyens et al., 2004a; b). Interestingly, Sod1-deficient female mice have drastically compromised fertility, with oogenesis halted at the middle of follicle development (Matzuk et al., 1998). The lack of Sod1 in Drosophila also causes reduction in female fertility (Phillips et al., 1989). Infertility is often the only abnormality in unstressed animals with knock-out of genes for antioxidant enzymes (Ho et al., 1998; Matzuk et al., 1998). On the other hand, transgenic male mice that express higher levels of mitochondrial MnSOD exhibit a decreased fertility (Raineri et al., 2001). At puberty, a number of primary oocytes begin to grow each month. One primary oocyte outgrows the others and resumes meiosis I. Interestingly, resumption of meiosis I is induced by an increase in ROS and inhibited by antioxidants (Takami et al., 1999, 2000; Kodaman and Behrman, 2001). Attendant to the increase in steroid hormone production of developing follicles is an increase in the activity of cytochrome P450, which in turn generates ROS such as H2O2 (Ortega-Camarillo et al., 1999). On the other hand, large goat ovarian follicles (>6 mm) exhibited greater catalase activity than granulosa cells from small (<3 mm) or medium (3–6 mm) sized follicles. After a uniform dose of FSH (200 ng/ml), both catalase activity and estradiol release were greater in large follicles than in medium or small follicles (Behl and Pandey, 2002). Since the dominant follicle will be the follicle with the highest estrogen concentration, the concomitant increases in catalase that is able to metabolize H2O2, and estradiol in response to FSH suggest a role for catalase in follicular selection and prevention of apoptosis (Behl and Pandey, 2002). Moreover, antioxidants are beneficial for meiosis II (Yoshida et al., 1993; Eppig, 1996; Behrman et al., 2001), which suggests a complex role for ROS indicating that regulated generation of ROS by the pre-ovulatory follicle is an important promoter of the ovulatory sequence (Ruder et al., 2008). An investigation of ROS regulation by the preovulatory follicle in response to LH indicated that a gonadotrophin-stimulated, protein kinase C-activated, NADPH/NADH oxidase-type superoxide generator in the preovulatory follicle exists and may be a regulating factor in ROS production during ovulation (Kodaman and Behrman, 2001). In females, oxidative stress is not limited to gametogenesis but is involved in virtually every aspect of female reproductive activity from oogenesis to parturition (Heininger, 2001; Rizzo et al., 2012). Oxidative stress is involved in granulosa cell estrogen production (LaPolt and Hong, 1995) and estrogen-mediated oocyte maturation (Tarin et al., 1998; Behrman et al., 2001) but, on the other hand, may contribute to ovarian senescence (Tarin et al., 1998; Behrman et al., 2001). Oxidative stress-responsive NF-kappaB modulates the expression of an androgen receptor (Roy et al., 1996) which further stresses the links between sex and oxidative stress. On the other hand, mild oxidative stress as trigger of reparative events may be the mechanism (similar to tolerance induction) which ensures the DNA repair (Schwartz et al., 1988; Dutta et al., 1996; Zhang et al., 1996; Fabisiewicz and Janion, 1998; Korzets et al., 1999) and thus rejuvenation and essential immortality of germ cell lines. Control of intracellular redox balance has emerged as a primary function of the p53 network (Trinei et al., 2002; Holley et al., 2010; Pani and Galeotti, 2011). Strong immunoreactivity for p53, the guardian of the genome, was observed in the nuclei of a number of spermatogonia, of some premeiotic spermatocytes and probably in all spermatids (Sjöblom and Lähdetie, 1996; Stephan et al., 1996). p53 physically interacts with key factors of homologous recombination: the human RAD51 protein and its prokaryotic homologue RecA (Stürzbecher et al., 1996). p53 is the link both between oxidative stress-generated DNA repair, recombination (Sjöblom and Lähdetie 1996, Schwartz et al., 1999), and quality control (Yin et al., 1998; 2002) during spermatogenesis.

7.2.4 Hypoxia-inducible factors

The increase in mitochondrial ROS formation during hypoxia modulates the activation of one or more hypoxia-inducible factors (Agani et al., 2000; Chandel et al., 2000; Schroedl et al., 2002; Brune and Zhou, 2003; Zhou et al, 2003; Guzy et al., 2005; Kietzmann and Görlach, 2005; Mansfield et al., 2005; Poyton et al., 2009a; b; Kuphal et al., 2010; Majmundar et al., 2010). The key transcription factor, hypoxia-inducible factor (HIF), originally identified in erythropoietin producing hepatoma cells (Semenza and Wang, 1992), is expressed very widely (probably universally) in mammalian cells (Wang and Semenza, 1993; Firth et al., 1994). The HIF system is phylogenetically highly conserved, even in primitive animal species that lack erythropoietin, red blood cells or even any specialized oxygen-delivery apparatus (Nagao et al., 1996; Huang and Bunn, 2003; Chuang et al., 2011; Loenarz et al., 2011; Ratcliffe, 2013). Under normoxic conditions, HIF-1alpha is rapidly degraded which is mediated by the ubiquitin–proteasome pathway and the von-Hippel Lindau protein (Salceda and Caro, 1997; Huang et al., 1998; Semenza, 1999; Tanimoto et al., 2000). Acute cellular hypoxia stabilizes and activates both the ubiquitous HIF-1alpha and the more tissue-specific HIF-2alpha (also known as EPAS1) transcription factors. Once stabilized by hypoxia, HIF-1alpha forms a heterodimer with HIF-1beta (also known as ARNT) on hypoxic responsive elements (HREs) within target gene promoters to drive the expression of HIF-1alpha targets (Williams and Benjamin, 2000; Powell et al., 2002). Inhibition of mitochondrial electron transport in hypoxia leads to increased production of ROS and this, rather than lack of molecular oxygen itself, is responsible for the reduction in HIF hydroxylase activity that signals hypoxia (Chandel et al. 1998; Bell et al., 2007; Ratcliffe, 2013). ROS produced by the redistributed mitochondria cause oxidative modification of the promoter regions of HIF-1 target genes. The introduction of oxidative modifications in these promoters enhanced HIF-1alpha association and gene expression (Al-Mehdi et al., 2012; Murphy, 2012). The expression of glutathione peroxidase or catalase (that convert hydrogen peroxide into water), but not superoxide dismutases 1 or 2 (that convert superoxide into hydrogen peroxide), prevented the hypoxic stabilization of HIF-1alpha in mammalian cells (Brunelle et al., 2005). Intriguingly, a pattern of antioxidant enzymes (normal SOD levels but low GPx and catalase levels) that enhances the stabilization of HIF-1-alpha by H2O2 (Brunelle et al., 2005) is exactly what Bauché et al. (1994) and others (see chapter 7.3) found in male germ cells. HIFs regulate the expression of genes involved in the adaptation to hypoxic conditions (Chandel et al., 1998; Semenza, 1999; 2009; Kaelin and Ratcliffe, 2008; Kaluz et al., 2008; Ratcliffe, 2013). HIF-1 is considered a key regulator of the ROS-mediated response to metabolic, hypoxic, or inflammatory stress (Seagroves et al., 2001; Zagórska and Dulak, 2004; Kietzmann and Görlach, 2005; Lum et al., 2007; Aragonés et al., 2009; Chen D et al., 2009; Görlach, 2009; Semenza, 2009; Kuphal et al., 2010; Shay and Simon, 2012). In mammalian cells, HIF-1alpha mRNAs are induced by the ubiquitous redox sensitive transcription factor nuclear factor-kappaB (NF-kappaB), the cellular ‘sensor’ for oxidative stress (Li and Karin, 1999; Mercurio and Manning, 1999; Storz and Toker, 2003; Gloire et al., 2006), so that NF-kappaB regulation plays an important role in the response to hypoxia (Jung et al., 2003; BelAiba et al., 2007; Bonello et al., 2007; Görlach and Bonello, 2008; van Uden et al., 2008; Görlach, 2009). HIF targets include members of stress-response gene families that mediate acute and chronic hypoxic adaptations, such as glucose transporters (Bashan et al., 1992; Ebert et al., 1995), glycolytic enzymes (Firth et al., 1994; 1995; Semenza et al., 1994), angiogenic factors such as vascular endothelial growth factor (VEGF) (Shweiki et al., 1992; Levy et al., 1995; 1996; Liu et al., 1995; Shima et al., 1995) and platelet-derived endothelial cell growth factor (Griffiths et al., 1997), haematopoietic growth factors such as erythropoietin (Semenza and Wang, 1992; Bunn and Poyton, 1996), and molecules that affect cell growth, survival, and motility. ROS are dose- and time-dependent inducers of VEGF gene and protein expression in vascular smooth muscle cells, human retinal pigment epithelial cells, human melanoma cells and glioblastoma cells (Bassus et al., 2001; Kuroki et al., 1996) while the antioxidant N-acetylcholine can suppress VEGF induction (Chua et al., 1998; Redondo et al., 2000). VEGF stimulates the expression of a cluster of nuclear-encoded mitochondrial genes, suggesting a role for VEGF in the regulation of mitochondrial biogenesis (Wright et al., 2008). On the other hand, HIF-1 actively represses mitochondrial function and O2 consumption by inducing pyruvate dehydrogenase kinase 1, which inactivates the tricarboxylic acid cycle (TCA) for cell survival (Kim et al., 2006; Papandreou et al., 2006). Therefore, energy conservation, in addition to energy generation, is an integral part of the hypoxic response. Under conditions of limited oxygen supply, anaerobic glycolysis becomes the predominant form of cellular ATP generation (Pasteur effect). HIF-1 is a necessary mediator of the Pasteur effect in mammalian cells (Seagroves et al., 2001). Many genes involved in glucose uptake and glycolysis were identified as HIF-1 target genes (Wenger, 2000; Zhong et al., 2010). Importantly, the Pasteur effect was observed in the adult rat testes whereas the Crabtree effect (the occurrence of anaerobic glycolysis despite aerobic conditions [Crabtree, 1929]) appeared only in immature testes (Leiderman and Mancini, 1968). HIF is also important for the establishment of the Warburg effect (Warburg et al., 1924; Sosa et al., 2013). In this metabolic shift from aerobic respiration, the machinery of glycolysis is upregulated in cancer cells and other rapidly proliferating cells (Wang et al., 1976; McKeehan, 1982), giving them a metabolic advantage for surviving and thriving in the oxygen poor microenvironment (Gladden, 2004; Sola-Penna, 2008). Aerobic glycolysis is an inefficient way to generate ATP. Under standard conditions, in which oxygen (O2) is abundant, and for long-term maintenance, healthy cells choose aerobic respiration. In aerobic respiration, which takes place in the cytoplasm and mitochondria, oxygen and glucose are used to generate 38 ATP molecules and carbon dioxide (CO2) is released and O2 is absorbed. In anaerobic respiration, which occurs only in the cytoplasm, 2 molecules of ATP are synthesized in the electron transport chain, using inorganic molecules other than O2; CO2 is released, but O2 is not required. Glycolytic enzymes are evolutionarily older and have reached catalytic perfection, thus suggesting that glycolysis may be the pathway of choice if glycolytic substrates are plentiful and oxygen is low. Anaerobic glycolysis generates lactate, and due to the low levels of energy produced only low levels of ROS are produced (Sosa et al., 2013). The major role of glycolysis in highly proliferating cells is to provide substrates to the pentose phosphate pathway for nucleotide synthesis, a great priority considering their high division rate (Kondoh et al., 2007; Weinberg and Chandel, 2009). Hence, the metabolism of proliferating cells is adapted to facilitate the uptake and incorporation of nutrients into the biomass (e.g., nucleotides, amino acids, and lipids) needed to produce a new cell (Vander Heiden et al., 2009; Shlomi et al., 2011; Krisher and Prather, 2012). Stem cells are more glycolytic than primary cells and possess reduced mitochondrial oxygen consumption, which helps these cells to control oxidative stress (Iyer et al., 1998; Kondoh et al., 2005; Funes et al., 2007; Diehn et al., 2009). Proliferating spermatogonia exhibit high glycolytic activity and fit into this metabolic pattern (Warburg et al., 1924; Roosen-Runge, 1953; Sosa et al., 2013). During spermatogenesis, the dependence of germ cells on lactate/pyruvate and glucose for energy metabolism keeps changing. Sertoli cells metabolize various substrates, preferentially glucose, the majority of which is converted to lactate and not oxidized via the TCA cycle (Rato et al., 2012). A gradient in mitochondrial activity has been described in testicular germ cells with stem cell spermatogonia presenting the least active mitochondria (Meinhardt et al., 1999). In Xenopus, mitochondria in the germ plasm have lower respiratory activity than mitochondria destined for the soma (Kogo et al., 2011). After passage to the luminal compartment germ cells rely on the breakdown of lactate and pyruvate provided by Sertoli cells (Boussouar and Benhamed, 2004; Rato et al., 2012). A high concentration of exogenous lactate apparently is required for spermatocytes and spermatids to use endogenous pyruvate as the predominant energy-yielding substrate via the TCA cycle (Grootegoed et al., 1984). The use of lactate as their primary substrate for producing ATP (Nakamura et al., 1982; 1984a) protects spermatocytes and spermatids against apoptotic cell death (Bustamante-Marín et al., 2012). Spermatogonia, mature spermatozoa and the somatic Sertoli cells exhibit high glycolytic activity, whereas spermatocytes and spermatids produce ATP mainly by oxidative phosphorylation (Mann and Lutwak-Mann, 1981; Robinson and Fritz, 1981; Grootegoed et al., 1984; Nakamura et al., 1984b; Trejo et al., 1995; Bajpai et al., 1998; Meinhardt et al., 1999; Harris, 2002; Marin et al., 2003; Ramalho-Santos et al., 2009). Interestingly, inhibition of ATP synthase decreased ATP levels and suppressed cell death, an effect not seen with inhibition of glycolysis, indicating that mitochondrial ATP production plays a role in regulating male germ cell apoptosis (Erkkila et al., 2006; Ramalho-Santos et al., 2009).

HIF-1 is known to have bimodal effects on cell physiology because it can activate either cell survival or cell death genes depending on the extent and duration of oxygen debt (Piret et al., 2002; Wiesener and Maxwell, 2003; Wang Y et al., 2004; Kilic et al., 2007). p53 has been implicated in the expression and degradation of HIF-1 (Blagosklonny et al., 1998; Ravi et al., 2000; Schmid et al., 2004) and, on the other hand, is HIF-dependent in hypoxic cells (An et al., 1998; Carmeliet et al., 1998; Chen et al., 2003b).

Hypoxia can drive genomic instability and alter DNA damage repair pathways. Intriguingly, hypoxic cells can acquire a mutator phenotype that consists of decreased DNA repair, an increased mutation rate and increased chromosomal instability (see chapter 10.1), a phenomenon which has been well established in carcinogenesis (Ralph et al., 2010) and tumor progression (Huang LE et al., 2007; Bristow and Hill, 2008). This might be particularly true in proliferating cells that have adapted to low O2 levels and continue to proliferate in the context of compromised DNA repair (Blais et al., 2006). Components of the HIF-1 system play essential roles in embryonic development and cellular differentiation (Ke and Costa, 2006; Cicchillitti et al., 2012). A clear link has been demonstrated between hypoxia, HIFs and molecules that are crucial for the regulation of the differentiation of stem and/or progenitor cells, including Notch, beta-catenin, Sox2, Oct4 and c-MYC (Simon and Keith, 2008; McCord et al., 2009; Bussolati et al., 2012). HIF-1alpha expression increased between embryonic days 8.5 and 9.5 in normal mouse embryos (Iyer et al., 1998). Knockout of either HIF-1alpha (Iyer et al., 1998; Ryan et al., 1998; Kotch et al., 1999), HIF-2alpha (Tian et al., 1998; Peng et al., 2000), or HIF-1beta (Maltepe et al., 1997) resulted in abnormal vascular development and lethality in mice. Embryos deficient in HIF-1alpha died by embryonic day 11 as a consequence of lack of blood vessel formation, defective formation of the neural fold, and cardiovascular malformation (Iyer et al., 1998; Ryan et al., 1998). OCT4 can be directly activated by HIF-1alpha and HIF-2alpha (Covello et al., 2006; Bussolati et al., 2012). OCT4 is essential for maintaining the undifferentiated state of embryonic stem cells, inner cell mass cells, the embryonic epiblast and PGCs (Scholer et al., 1990; Nichols et al., 1998). OCT4 expression is tightly controlled during embryogenesis and adult life, and its downregulation is required for differentiation of the trophectoderm lineage and subsequent gastrulation by the epiblast; however, OCT4 expression is maintained in PGCs. In the adult, outside the stem-cell populations OCT4 is exclusively expressed in germ cells (Simon and Keith, 2008). OCT4 is essential for germ-cell maintenance, but dispensable for somatic stem-cell self-renewal (Kehler et al., 2004; Lengner et al., 2007). HIF-1alpha directly inhibits the transcription factor c-Myc, causing de-repression of its targets p21 and p27 (Koshiji et al., 2004; Dang et al., 2008). c-Myc targets involved in mismatch repair are also modulated by HIF-1alpha, suggesting a role for HIF in hypoxia-induced genetic instability (Koshiji et al., 2005; To et al., 2006; Huang LE, 2008).

Hypoxia induces specific microRNAs collectively referred to as hypoxamirs (Chan and Loscalzo, 2010). HIF-1 expression is also controlled by specific microRNAs and, in turn, controls the expression of other microRNAs, which fine-tune adaptation to low oxygen tension (Loscalzo, 2010). miR-210 is the master hypoxamir (Chan et al., 2012). miR-210 is consistently upregulated in both normal and transformed hypoxic cells (Kulshreshtha et al., 2007; Camps et al., 2008; Corn, 2008; Giannakakis et al., 2008; Ivan et al., 2008; Pulkkinen et al., 2008, Crosby et al., 2009; Huang et al., 2010; Devlin et al., 2011; Gorospe et al., 2011; Mutharasan et al., 2011; Pocock, 2011; Voellenkle et al., 2012). miR-210 is involved in repressing mitochondrial respiration (Chan et al., 2009; Chen et al., 2010; Favaro et al., 2010; Puisségur et al., 2011), exaggerates production of mitochondrial ROS (Favaro et al., 2010) and impairs DNA repair (Crosby et al., 2009; Fasanaro et al., 2009; Devlin et al., 2011). miR-210 appears to be able to bypass hypoxia-induced cell cycle arrest and partially reverse the hypoxic gene expression signature (Zhang et al., 2009). In these particular cells, miR-210 activates the proto-oncogene myc pathway, via downregulation of the c-Myc antagonist MNT, and loss of MYC abolished miR-210-mediated override of hypoxia-induced cell cycle arrest. Importantly, miR-210 expression has been identified in gametogenesis of Drosophila, Xenopus, fishes and mammals (Grün et al., 2005; Madison-Villar and Michalak, 2011; Torley et al., 2011; Ambady et al., 2012; Bizuayehu et al., 2012; El Naby, 2012; Miles et al., 2012).

The hypoxic response is characterized by a rapid inhibition of protein synthesis which occurs through the repression of the initiation step of mRNA translation (Wouters et al., 2005; Buchan and Parker, 2009; Spriggs et al., 2010). However, regulation of translation also results in a specific increase in the synthesis of a subset of hypoxia-induced proteins (e.g. heat shock proteins; Borger and Essig, 1998; Yueh and Schneider, 2000; Koritzinsky and Wouters, 2007). These regulatory pathways link hypoxia with the stress granules (see chapter 7.2.7), the site of RNA triage (Anderson and Kedersha, 2002b; 2006; 2008; Kedersha and Anderson, 2002).

7.2.4.1 Male gametogenesis

Decreased O2 levels and HIF-1α activation are critical components of spermatogenesis (Gruber et al., 2010). Various HIF-1 isoforms are expressed constitutively by spermatocytes, spermatids, spermatozoa and Leydig cells (Marti et al., 2002; Powell et al., 2002; Depping et al., 2004; Wenger and Katschinski, 2005; Turner and Lysiak, 2008; Lysiak et al., 2009; Palladino et al., 2011), reflecting the stressful conditions under which spermatogenesis takes place. In cancer cells, constitutive HIF-1α protein expression and activity is mediated by ROS and the NFkappaB pathway (Kuphal et al., 2010). Under normoxic conditions, the von Hippel–Lindau tumor suppressor protein (VHL) targets the HIF-1α subunit for rapid ubiquitination and proteasomal degradation (Maxwell et al., 1999). The significance of stringent regulation of HIF-1α in the testis is supported by findings in male mice with conditional inactivation of the VHL gene (Wenger and Katschinski, 2005). These animals show oligospermia, reduction in testicular weight and infertility, suggesting that impaired regulation of HIF-1alpha results in defects in spermatogenesis (Ma et al., 2003). Because HIF-1α knock-out mice die during embryogenesis (Iyer et al., 1998), the role of HIF-1alpha in testis cannot be investigated in this animal model. However, HIF-2alpha knock-out mice suffer among others from azoospermia (Scortegagna et al., 2003). Postnatal HIF-1alpha ablation leads to male infertility, with reduced testis size and weight. While immature spermatogonia and spermatocytes are present in HIF-1α-/- testes, spermatid and spermatozoan numbers are dramatically reduced. This is not due to germ cell-intrinsic defects. Rather, HIF-1α-/- Sertoli cells exhibit decreased ability to form tight junctions, thereby disrupting the blood-testis barrier necessary for proper spermatogenesis (Gruber et al., 2010). Hypoxia may also be a physiological signal for testosterone production by Leydig cells (Gonzales et al., 2009; 2011). The promoter of the mouse 3beta-hydroxysteroid dehydrogenase type 1 gene, which encodes a key enzyme in testosterone production, is a potential target of HIF-1 (Lysiak et al., 2009). Intermittent hypoxia increased testosterone secretion from Leydig cells in vivo and in vitro through the activation of adenylate cyclase, enhanced activities of P450scc, 3beta-hydroxysteroid dehydrogenase, and 17beta-hydroxysteroid dehydrogenase, increased levels of calcium ion, and influx into Leydig cells (Hwang et al., 2007; 2009). Many HIF-1 target genes, especially the glycolytic enzymes, are also expressed in the testis (Gold et al., 1983; McCarrey et al., 1996; Li et al., 1998; Welch et al., 2000). Various testis-specific isoforms of glycolytic enzymes are expressed in the haploid stages of spermatogenesis and are still active in mature spermatozoa, including phosphoglycerate kinase 2 (Gold et al., 1983; McCarrey et al., 1996), glyceraldehyde 3-phosphate dehydrogenase-2 (Welch et al., 2000) and lactate dehydrogenase C (Li et al., 1998). Interestingly, all four genes encoding the bifunctional enzymes 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase (PFKFB-1 to -4) are highly expressed in the testis in a hypoxia-inducible manner (Minchenko et al., 2003; 2004). PFKFB-1 to -4 regulate the levels of fructose-2,6-bisphosphate, a potent allosteric regulator of 6-phosphofructo-1-kinase and hence a key regulator of the glycolytic flux. Indeed, sperm capacitation, motility changes, acrosome reaction and fertilization are exclusively dependent on anaerobic glycolysis and can occur under strictly anaerobic conditions (Fraser and Quinn, 1981).

While somatic Sertoli and Leydig cells produce VEGF-A, a VEGF isoform (Ergün et al., 1997; Liu and Yang, 2004; Reddy et al., 2012), VEGF receptors (VEGF-R1 and VEGF-R2) are expressed on testicular blood vessels (Shweiki et al., 1993; Collin and Bergh, 1996a; Ergün et al., 1997; Korpelainen et al., 1998; Marti and Risau, 1998) and testicular germ cells in humans (Ergün et al., 1997), rats (Rudolfsson et al., 2004), mice (Nalbandian et al., 2003), and cattle (Caires et al., 2009). The presence of VEGF receptors on Sertoli and Leydig cells suggests an autocrine regulatory effect of VEGF on the activity of both cell types (Ergün et al., 1997). In this context it is of interest that intermittent hypoxia (that could result from the autocrine negative feedback of VEGF) increased Leydig cell testosterone secretion in vivo and in vitro (Hwang et al., 2007; 2009) while chronic hypoxia leads to decreased testosterone and reproductive dysfunction (Fahim et al., 1980; Semple et al., 1980; Aasebo et al., 1993; Soukhova-O’Hare et al., 2008; Liao et al., 2010) possibly in conjunction with testicular vascularization and hyperthermia (Farías et al., 2005; Madrid et al., 2012). Intriguingly, an intermittent hypoxic training regimen consisting of 14 consecutive days with four repetitions of 5–7 min each of induced hypoxia interspersed with 5 min periods of ambient inspiration produced a stimulatory effect on male reproductive function and remedied male subfertility (Swanson and Serebrovska, 2012). VEGF, however, had no effect on testicular vasomotion (see chapter 7.2.1) (Rudolfsson et al., 2004). VEGF expression is conserved in vertebrate testes of fish, reptiles and mammals (Reddy et al. 2012). VEGF increases the number of proliferating testicular endothelial cells (Rudolfsson et al., 2004). Seasonal regression and regrowth of the white-footed mouse testes positively correlated with VEGF expression (Young and Nelson, 2000). Adult roe deer males show VEGF mRNA expression dependent on season, reaching its highest level at the peak of spermatogenesis during the pre-rutting period and had its nadir at the end of the rut when involution already began (Wagener et al., 2010; Schön et al., 2010). These and other findings (Dolci et al., 2001; Guo R et al., 2004; Schmidt et al., 2006; 2007) suggested that VEGF may directly affect the proliferation and differentiation of germ cells, an effect that may, at least in part, be independent from the formation of testicular microvasculature (Korpelainen et al., 1998; Wagener et al., 2010). Ischemia-reperfusion injury of rat testis significantly increased VEGF protein and mRNA in a time-dependent manner in testicular vascular endothelial cells and germ cells (Hashimoto et al., 2009). Overexpression of VEGF has harmful effects on spermatogenesis in transgenic mice (Korpelainen et al., 1998; Huminiecki et al., 2001) and in human testes with varicocele (Shiraishi and Naito, 2008) implicating VEGF in male fertility (Korpelainen et al., 1998; Huminiecki et al., 2001; Bott et al., 2006; Shiraishi and Naito, 2008).

7.2.4.2 Female gametogenesis

Low oxygen tension during meiotic arrest improves the developmental competence of mammalian oocytes (Eppig and Wigglesworth, 1995; Hashimoto et al., 2000; 2002). The functional hypoxia and the growing-follicle HIF signaling is involved in follicular differentiation and luteinization (Tam et al., 2010). Within non-atretic follicles, HIF-1alpha mRNA was highly expressed in the granulosa cell layer, while weaker labeling was evident in the theca interna. These results suggest that HIF-1alpha may play a role in the regulation of cellular metabolism during follicular growth (Boonyaprakob et al., 2005). Significant differences in dissolved oxygen content occur in follicular fluids aspirated from follicles of equivalent size and ultrasonographic appearance. Oocytes from severely hypoxic follicles were associated with high frequencies of DNA damage. Oocytes with cytoplasmic and chromosomal disorders and embryos with multinucleated blastomeres and limited developmental ability were derived predominantly from underoxygenated follicles (Van Blerkom, 1998). HIF-1alpha expression appears to be the signal that provides trophic support and competitive advantage to growing follicles: human oocytes collected from follicles that contain a higher concentration of VEGF have a higher viability (Van Blerkom et al., 1997; Van Blerkom, 2000; Monteleone et al., 2008).

Initiation and maintenance of follicular growth depends on development of the follicular microvasculature (Clark, 1900; Andersen, 1926). The density of capillaries surrounding the dominant, maturing follicle is greater than that of other smaller follicles. Preovulatory follicles of monkeys were found to have similar concentrations of gonadotropin-binding sites; however, only the follicle that was destined to ovulate became heavily labelled after intravenous injection of labelled gonadotropin consistent with an increased vascularity of the dominant follicle (Zeleznik et al., 1981). Follicles undergoing atresia have decreased vascularity (Zeleznik et al., 1981; Moor and Seamark, 1986) and reduced DNA synthesis of follicular endothelial cells, in association with reduced follicular vascularity, is one of the earliest signs of atresia (Greenwald, 1989). Decreased vascularity may limit access of atretic follicles to nutrients, substrates, and tropic hormones, thereby maintaining these follicles in an atretic state (Moor and Seamark, 1986). VEGF measurements of follicular fluid indicated a potentially important role for this factor both in perifollicular angiogenesis and in the regulation of intrafollicular oxygen levels (Van Blerkom et al., 1997). Metabolic/hypoxic stress that involves ROS in hypoxic signaling plays a pivotal role in the follicular angiogenic process and stimulates VEGF synthesis by granulosa cells (Neeman et al., 1997; Tempel-Brami and Neeman, 2002; Basini et al., 2004a; b; 2007). However, in contrast with solid tumors of similar size, the spatial and temporal discrepancy between VEGF expression and angiogenesis suggests that cumulus cells secrete to the follicular fluid, in addition to VEGF, material with antiangiogenic activity. Hyaluronan as a high molecular weight suppressor of angiogenesis, produced by the cumulus cells, can account for this antiangiogenic activity maintaining an avascular follicular antrum (Tempel et al., 2000; Tempel-Brami and Neeman, 2002). In addition to hypoxia, FSH and LH stimulate VEGF production by granulosa cells (Christenson and Stouffer, 1997). VEGF mRNA is expressed in large preantral follicles (Ravindranath et al., 1992). Direct injection of VEGF into the mouse ovary resulted in the development of an enhanced vascular network promoting follicular development and diminishing apoptosis (Quintana et al., 2004), acting as a survival factor for granulosa cells (Greenaway et al., 2004). Formation of antrum coincides with continued follicular angiogenesis resulting in the development of an intricate vascular mesh that secures an increasing supply of gonadotropins, growth factors, oxygen, steroid precursors, as well as other substances to the growing follicle (Geva and Jaffe, 2000a; b; Wulff et al., 2001). VEGF is essential for angiogenesis and the generation of healthy ovulatory follicles and corpora lutea. Even in the presence of gonadotropins, administration of substances that inactivate VEGF and its signaling, block the development and function of preovulatory follicles caused by arrests to both angiogenesis and antrum formation (Wulff et al., 2001; 2002; Zimmermann et al., 2001; 2002; 2003; Fraser et al., 2005; Taylor et al., 2007).

Elevated VEGF concentration in follicular fluid is a marker of follicular hypoxia, reduced perifollicular blood flow, and ovarian aging (Friedman et al., 1998; Battaglia et al., 2000; Ocal et al., 2004). Women of advanced reproductive age undergoing follicular aspiration after superovulation showed increased follicular fluid VEGF concentrations compared with younger women consistent with a hypoxic environment within follicles of older women (Friedman et al., 1997). Taking into account that VEGF expression is associated with oxidative stress (Neeman et al., 1997; Basini et al., 2004a; b; 2007) these markers may also be related to oocyte DNA damage (Van Blerkom, 1998). Thus the complex causal relationship between follicular hypoxia and its timing make VEGF both a marker of oocyte high quality (Van Blerkom et al., 1997; Van Blerkom, 2000; Monteleone et al., 2008) and damage (Van Blerkom, 1998). Therefore, follicular blood flow is a better predictor for the outcome of in vitro fertilization than follicular fluid VEGF (Kim KH et al., 2004).

7.2.5 Cytokines and nuclear factor-kappaB

The cytokines interleukin-1 (IL-1) and tumour necrosis factor-alpha (TNF) exhibit a variety of stimulatory activities on maturation, differentiation and growth of many cell types involved in development and inflammation, such as fibroblasts, synovial cells, endothelial and epithelial cells, bone marrow cells, and T- and B-lymphocytes (Dinarello, 1985; Krakauer, 1986; Akahoshi et al., 1988; Chouaib et al., 1988; Bonavida and Granger, 1990; Le Hir et al., 1996; Muñoz-Fernández and Fresno, 1998; Azuma et al., 2000). Both cytokines have profound effects on inflammatory reactions (Beutler and Cerami, 1986; 1987; Billingham, 1987; Eastgate et al., 1988; Marx, 1988; Tracey et al., 1986; 1987; Fiers, 1991; Beutler, 1995). There seems to be an analogy between hematopoietic and spermatogenic systems (Huleihel and Lunenfeld, 2004). In the testes, cytokines are involved in the Sertoli-germ cell interaction, Leydig cells steroidogenesis, blood flow, and vascular permeability as well as in local control of immune cells (Spangelo et al., 1995). Macrophages are closely associated with Leydig cells in the testicular interstitium and comprise about 25% of the interstitial cell population in the testes of mammals including man, monkeys, rat and boar (Fawcett et al., 1973; Niemi et al., 1986). IL-1 and TNF are released by testicular macrophages, Sertoli and Leydig cells (Khan et al., 1987; Gérard et al., 1991; Söder et al., 1991; Wang et al., 1991; Hales et al., 1992; Hutson, 1992; Lin et al., 1993; Haugen et al., 1994; Kern et al., 1995; Cudicini et al., 1997). IL-1 receptors have been reported on Sertoli cells, Leydig cells, testicular macrophages, and germ cells suggesting both autocrine and paracrine functions. IL-1 was shown to promote proliferation and differentiation of spermatogonia and preleptotene spermatocytes (Pollanen et al., 1989; Parvinen et al., 1991) and is correlated with spermatogonial DNA synthesis in the rat seminiferous epithelium (Söder et al., 1991). Interleukin-1 and TNFalpha are also potent growth factors for immature rat Sertoli and Leydig cells (Khan et al., 1992; Petersen et al., 2002; 2004) and stimulate testosterone biosynthesis and secretion in adult Leydig cells (Verhoeven et al., 1988; Warren et al., 1990; Svechnikov et al., 2001). IL-1 and TNF both induce oxygen radical formation in a variety of cell types (Klempner et al., 1979; Tsujimoto et al., 1986; Ferrante et al., 1988; Meier et al., 1989; Radeke et al., 1990; Feng et al., 1995; Lo et al., 1998; Lee FY et al., 1999; Bonizzi et al., 2000; Corda et al., 2001; Gloire et al., 2006; Ishikawa and Morris, 2006; Davies et al., 2008; Morgan and Liu, 2010), e.g. by the activation of NAD(P)H oxidases (De Keulenauer et al., 1998; Gauss et al., 2005; Ammons et al., 2007; Yang D et al., 2007). A key cytokine signaling pathway is through NF-kappaB (Schreck et al., 1991; Bonizzi et al., 2000; Baud and Karin, 2001; Janssens et al., 2002; Senftleben and Karin, 2002; Tian and Brasier, 2003; Gloire et al., 2006).

The metazoan transcription factor NF-kappaB signaling pathway shows some homology to the yeast mitochondrial retrograde response (Srinivasan et al., 2010). NF-kappaB activation is part of the cellular stress response to a variety of factors including cytokine stimulation, irradiation, and ischemia-reperfusion. NF-kappaB is a master regulator for inflammatory responses, mediating cellular defense against infectious agents and environmental and cellular stress (Mercurio and Manning, 1999; Wang et al., 2002; Piva et al., 2006), and is highly conserved in innate immunity (Silverman and Maniatis, 2001). Numerous studies have demonstrated that in cells subjected to oxidative stress there is a potent NF-kappaB response. As such, NF-kappaB is often referred to as the cellular ‘sensor’ for oxidative stress (Li and Karin, 1999; Mercurio and Manning, 1999; Storz and Toker, 2003; Oliveira-Marques et al., 2009; Morgan and Liu, 2011; Siomek, 2012) that is linked to ROS by reciprocal, both negative and positive controls (Bubici et al., 2006; Basak and Hoffmann, 2008; Morgan and Liu, 2011). Oxidative stress-sensitive NF-kappaB is upregulated during gamete maturation (Delfino and Walker, 1998; 1999a; b; Lilienbaum et al., 2000; Shalini and Bansal, 2007; Paciolla et al., 2011). Nuclear NF-kappaB is present in Sertoli cells and in the late meiotic and postmeiotic germ cells of the rodent testis (Delfino and Walker, 1998; Lilienbaum et al., 2000) at higher levels than in any other tissue assayed (Budde et al., 2002). Intense cytoplasmic staining for NF-kappaB was found in the spermatogonia and early meiotic germ cells of the human testis. In light of the importance of the stable pool of immature germ cells for continuous spermatogenesis, the extensive deposition of NF-kappaB in the cytoplasm of spermatogonia and immature spermatocytes may be used for rapid nuclear translocation and transcriptional activation of protective genes on certain stimuli (Pentikäinen et al., 2002). In Sertoli cells, NF-kappaB elements in the androgen receptor promoter have been identified as being responsible for increased androgen receptor expression, representing an important (cell type-specific) regulatory mechanism required to maintain efficient spermatogenesis (Delfino et al., 2003; Zhang L et al., 2004).

7.2.6 Heat shock response

All organisms share a common molecular stress response that includes a dramatic change in the pattern of gene expression and the elevated synthesis of a family of stress-induced proteins called heat shock proteins (Craig, 1985; Lindquist and Craig, 1988; Kalmar and Greensmith, 2009). Genes encoding a variety of molecular chaperones, and proteins that catalyse ROS and disulfide bond metabolism are induced in response to oxidative stress (Barford, 2004). Cells react to stressful conditions by activation of heat-shock factors (HSFs), of which there are three mammalian (HSF1, HSF2, and HSF4) and one avian (HSF3) isoforms (Pirkkala et al., 2001; Kalmar and Greensmith, 2009). Functional conservation of HSFs among eukaryotes has been revealed by the finding that HSFs from various organisms, including insects, mammals and plants, can substitute for yeast HSF in Saccharomyces (Pirkkala et al., 2001; Abane and Mezger, 2010; Björk and Sistonen, 2010; Fujimoto and Nakai, 2010). Activated HSFs bind to heat-shock elements (HSEs) within the promoters of their target genes and induce synthesis of protective molecular chaperones called heat shock proteins (Hsps). Hsps are highly conserved proteins present in organisms ranging from bacteria to man. Hsps prevent protein misfolding and are required for stress resistance and healthy cell growth, development, and aging (Morimoto, 1993; 2008; Bukau et al., 2006; Prahlad and Morimoto, 2009). In addition to extracellular stimuli, several ‘nonstressful’ conditions induce Hsps during normal cellular growth and development. The eukaryotic transcription factor that appears to be most directly redox regulated is HSF1, a highly conserved transcription factor which mediates the transcription of a complex of genes in response to heat, oxidative stress and a variety of other stressors (Fedoroff, 2006). Using HSF1-null mice, Izu et al. (2004) showed that apoptosis of pachytene spermatocytes was markedly inhibited in testes with a single exposure to heat and in cryptorchid testes, indicating that HSF1 promotes apoptotic cell death of pachytene spermatocytes exposed to thermal stress. In marked contrast, HSF1 acts as a cell-survival factor of more immature germ cells, probably including spermatogonia, in testes exposed to high temperatures. These results demonstrate that HSF1 has two opposite roles in male germ cells independent of the activation of heat shock genes (Izu et al., 2004).

HSF2 regulates the activity of heat shock genes under perceived non-stressful conditions, such as differentiation and development (Sistonen et al., 1994). Moreover, HSF2 modulates HSF1-mediated regulation of a variety of other hsp genes, clearly demonstrating a functional role for HSF2 in stress responses (Sistonen et al., 1994; Mathew et al., 2001; He et al., 2003; Åkerfelt et al., 2007; 2010; Östling et al., 2007; Sandqvist et al., 2009). The first indication of a role for HSFs in oogenesis was suggested by studies in Drosophila, which demonstrated that the unique Drosophila HSF is essential for oogenesis and implied that its role in oogenesis is mediated not only by the regulation of Hsp genes (Jedlicka et al., 1997; Abane and Mezger, 2010). HSF1 is highly expressed in mouse nonfertilized ovulated oocytes arrested at metaphase II and in pre-implantation embryos (Mezger et al., 1994; Christians et al., 1997). The deficiency in HSF1 provokes an oxidative stress to which oocytes are particularly sensitive (Liu L et al., 2000; Dumollard et al., 2007a). In gonads, HSF2 is abundantly expressed and plays a role in spermatogenesis (Sarge et al., 1994; Fiorenza et al., 1995; Kallio et al., 2002; Wang G et al., 2003; Åkerfelt et al., 2008) and oogenesis (Kallio et al., 2002; Wang G et al., 2003; Abane and Mezger, 2010). During mouse postnatal testis development, the HSF2-beta mRNA isoform (that is primarily expressed in the heart and brain) is switched to the HSF2-alpha isoform. HSF2-alpha protein, the predominant isoform expressed in testis cells, is a more potent transcriptional activator than the HSF2-beta isoform (Goodson et al., 1995). In contrast to hsf1-/- mice, which exhibit normal spermatogenesis, targeted disruption of hsf2 results in reduced testicular size but only a small impairment in male fertility. Hsf2-/- mice displayed increased apoptosis at pachytene, meiotic M and type A spermatogonia stage (Kallio et al., 2002: Wang G et al., 2004). Meiosis is affected by HSF2 deficiency, in both males and females. Disruption of both hsf1 and hsf2 results in a more severe phenotype associated with male sterility due to severe defects in spermatogenesis. Earliest defects observed are the reduced number of germ cells in juvenile mice, and germ cells that enter the meiotic prophase fail to progress beyond the pachytene stage. This is associated with a reduction or absence of transcription of genes critically involved in spermatogenesis. The findings suggest that additive or synergistic transcriptional activity of both hsf1 and hsf2 is required for normal mammalian spermatogenesis and male fertility (Wang G et al., 2004).

The Hsps are a group of highly conserved proteins that are induced in both prokaryotes and eukaryotes by elevated temperatures or a variety of cellular stressors (Samali and Orrenius, 1998; Feder and Hofmann, 1999; Kalmar and Greensmith, 2009). When cells are exposed to elevated temperatures the heat shock protein HSP70 is the most prominently expressed. Variation in Hsp70 gene expression and polymorphisms has been positively correlated with variation in thermotolerance in Drosophila melanogaster, in Caenorhabditis elegans, and in mammals (Hashmi et al., 1997; Maloyan et al., 1999; Sonna et al., 2002; Gong and Golic, 2004; Singh et al., 2006; Bernabucci et al., 2010). Hsps are constitutively expressed during both female and male germ cell maturation in invertebrates and vertebrates (Ambrosio and Schedl, 1984; Glaser and Lis, 1990; Ohsako et al., 1995, Paranko et al., 1996, Dix, 1997, Michaud et al., 1997; Sarge and Cullen, 1997; Joanisse et al., 1998; Son et al., 1999; Neuer et al., 2000; Aguilar-Mahecha et al., 2001; Nonoguchi et al., 2001; Christians et al., 2003; Kamaruddin et al., 2004; Ma et al., 2007; Huo et al., 2008; Meistertzheim et al., 2009; Abane and Mezger, 2010; Åkerfelt et al., 2010; Lachance et al., 2010) and in plant male gametogenesis (Schöffl et al., 1998). Several HSPs have been found to be constitutively expressed in germ cells at specific stages of development. Two HSP70-related genes, Hsp70-2 and Hsp70t, are expressed at high levels in spermatocytes and spermatids, respectively, and at considerably lower or nondetectable levels in other tissues (Krawczyk et al., 1988; Zakeri et al., 1988; Matsumoto and Fujimoto, 1990; Rosario et al., 1992; Son et al., 1999; Fourie et al., 2001). Male mice deficient for Hsp70-2 are sterile due to a failure of spermatocytes to proceed through the first meiotic division and increased spermatocyte apoptosis (Dix et al., 1996), and a decreased level of Hsp70-2 mRNA in human testis has been associated with abnormal spermatogenesis and sterility (Son et al., 2000). Hsp70-2 is also expressed at moderate levels in a variety of tumor cells and controls tumor cell proliferation (Rohde et al., 2005). The temperature threshold for induction of HSP72 encoded by the hsp70 gene is lower in male germ cells than in somatic cells (Sarge et al., 1995; Sarge, 1995). Similarly, Hspa4l, a member of the heat shock protein 110 family that responds to a lower temperature heat shock rather than the traditional elevated temperatures is abundantly expressed from late pachytene spermatocytes to postmeiotic spermatids (Kaneko et al., 1997a; b; Held et al., 2006; 2011). HSP90 and HSP60 are expressed in spermatogonia and spermatocytes (Gruppi et al., 1991; Meinhardt et al., 1995; Ohsako et al., 1995; Sarge and Cullen, 1997). Hsp90 is one of the most highly expressed cytosolic molecular chaperones, comprising 1% of the total cellular protein even in non-stressed conditions. It interacts with several hundred client proteins, simultaneously recruiting its own battery of co-chaperones, thus maintaining normal cellular functions in an ATP-dependent manner (Taipale et al., 2010; Echeverria and Picard, 2010). HSP90 is highly expressed in PGCs and continues to be expressed in both male and female pre-meiotic germ cells (Ohsako et al., 1995). Hsp90alpha is the major Hsp expressed by fully grown oocytes and markedly down-regulated by the absence of HSF1 (Metchat et al., 2009). Interestingly, Hsp90 is constitutively expressed at 2–10-fold higher levels in tumors compared to normal non-gonadal tissues (Ferrarini et al., 1992). Spermatogenesis is the developmental process most sensitive to the loss of Hsp90 function (Yue et al., 1999). Male KO mice without the Hsp90alpha isoform are sterile because spermatogenesis arrests specifically at the pachytene stage of meiosis I (Grad et al., 2010; Kajiwara et al., 2012). Birds are unique among homeothermic animals in developing spermatogenesis at the elevated avian internal body temperature of 40-41° C. While the expression of Hsp70 and ubiquitin did not change upon heat shock in mouse testicular cells, both the amount and polyadenylation of Hsp70 and ubiquitin transcripts increased when male germ cells from adult chicken testis were exposed to elevated temperatures (Mezquita et al., 1998).

Heme oxygenase (HO) is a heat shock protein and its induction occurs together with the induction of other HSPs during various physiopathological conditions (Shibahara et al., 1987; Keyse and Tyrrell, 1989; Mitani et al., 1990; Sharp et al., 1999). HO oxidatively cleaves heme (Fe-protoporphyrin IX) to produce CO, biliverdin, and free iron (Maines, 1997). Endogenously produced CO has been shown to possess intriguing signaling properties affecting numerous critical cellular functions including but not limited to inflammation, cellular proliferation, and apoptotic cell death (Ryter et al., 2006). Depending on the cellular milieu, HO activity can be considered as a proponent of oxidative stress with both pro-oxidant and anti-oxidant activities: (i) by liberating chelated iron from the heme molecule creating catalytically active free iron causing formation of ROS by Haber–Weiss chemistry (Mello-Filho and Meneghini, 1984; Maines and Gibbs, 2005) with devastating effects in a cell or, (ii) as a means to generate the antioxidant bilirubin (Stocker et al., 1987; Maines, 1997; Baranano et al., 2002). Accordingly, numerous studies proposed that the activity of the HO system might provide cellular protection against oxidative stress (Stocker, 1990; Maines, 1997; Niess et al., 1999; Pomeraniec et al., 2004; Shiraishi and Naito, 2005) and, on the other hand, may augment apoptosis (Liu XM et al., 2000; Ozawa et al., 2002; Harada et al., 2004a). HO-2 is the constitutive form of the stress inducible HO-1 gene and is abundantly expressed throughout spermatogenesis (Trakshel et al., 1986; Kurata et al., 1993; Ewing and Maines, 1995; McCoubrey et al., 1995; Aguilar-Mahecha et al., 2001) and in ovarial theca, granulosa cells, and corpora lutea as well as the ovarian stroma (Alexandreanu and Lawson, 2003; Harada et al., 2004a; Malone and Michalak, 2008) suggesting that HO-2 plays an important role in germ cell development. Moreover, HO-1 expression in the ovary plays a pivotal role in the process of oocyte ovulation, fertilization, and corpora lutea maintenance (Zenclussen et al., 2012). Intriguingly, HO-2 levels in the testis are controlled by glucocorticoids and developmental and tissue-specific factor(s) determine generation of transcripts unique to the organ (Liu N et al., 2000). The glucocorticoid element is the only demonstrated functional response element in the promoter sequence of HO-2 (McCoubrey and Maines, 1994; Weber et al., 1994; Raju et al., 1997). Leydig cells play a key role in oxidative stress-induced downregulation of spermatogenesis (Ozawa et al., 2002). Under stress conditions, HO-1 is upregulated predominantly in Leydig cells coinciding with CO generation, and microsomal cytochromes P450 which are required for steroidogenesis are suppressed. Under these circumstances, diploid and tetraploid germ cells in peripheral regions of seminiferous tubules, suggesting involvement of spermatogonia and primary spermatocytes, exhibited apoptosis (Ozawa et al., 2002).

7.2.7 Stress and germ granules

In response to environmental stress (e.g. heat, hypoxia, hyperosmolarity and oxidative conditions), eukaryotic cells shut down protein synthesis in a stereotypic response that conserves anabolic energy for the repair of stress-induced damage, enhancing their ability to withstand stress (Holcik and Sonenberg, 2005; Lavut and Raveh, 2012). This results in a notable saving of cellular energy, which is mainly consumed in the process of translation (estimated as up to 50% of the cellular energy, depending on the organism) (Warner, 1999; Mathews et al., 2000; Rudra and Warner, 2004). Cytoplasmic RNA granules in germ cells and somatic cells have emerged as important players in the posttranscriptional regulation of gene expression. RNA granules contain various ribosomal subunits, translation factors, decay enzymes, helicases, scaffold proteins, and RNA-binding proteins, and they control the localization, stability, and translation of their RNA cargo (Anderson and Kedersha, 2006). RNA-binding proteins and miRNAs can block translation by mobilizing mRNAs to subcytoplasmic domains where translation is inhibited; such sites include processing bodies (P-bodies), neuronal RNA granules, and stress granules (SGs). These are cytoplasmic foci that form transiently in response to cell stress and damage and harbor mRNAs that are typically stable and not translated (Gottschald et al., 2010; Gorospe et al., 2011). SGs have been observed in yeast (such as Saccharomyces pombe), protozoa (Trypanosoma brucei) and metazoa (such as Homo sapiens and Caenorhabditis elegans). They have also been observed in plants and in chloroplasts, suggesting that they may be assembled in prokaryotes as well (Anderson and Kedersha, 2009). In fact, stress induces the assembly of RNA granules in an organelle with bacterial ancestry, the chloroplast of Chlamydomonas reinhardtii (Uniacke and Zerges, 2008). SGs are phase-dense particles that appear in the cytoplasm of both plant and animal cells subjected to a wide variety of stresses (eg. heat, UV irradiation, oxidative conditions, hyperosmolarity) (Anderson and Kedersha, 2002a; b; Moeller et al., 2004; Kedersha and Anderson, 2007; Arimoto et al., 2008). SGs form during hypoxia both in vitro and in vivo and reoxygenation leads to disaggregation of these granules and restoration of protein synthesis (Moeller et al., 2004). SGs are hallmarks of stalled translational initiation (Brengues et al., 2005; Kedersha et al., 2005; Yamasaki and Anderson, 2008). SGs are a major adaptive defense mechanism that negatively regulates the stress-activated p38 and JNK MAPK apoptotic response (Arimoto et al., 2008; Tsai and Wei, 2010). The RNA silencing machinery redistributes into SGs in cells that go through mitosis after UV irradiation (Pothof et al., 2009). The involvement of oxidative stress in the causation of translational arrest and SG assembly has been shown in yeast, plants and animals (Anderson and Kedersha, 2002b; Kedersha and Anderson, 2002; 2007; Gilks et al. 2004; Cohen et al., 2005; McEwen et al., 2005; Shenton et al., 2006; Kedersha and Anderson, 2007; Kedersha et al., 2008, Uniacke and Zerges, 2008; Pothof et al., 2009; Wolf et al., 2010; Brown et al., 2011; Emara et al., 2012). The stress-induced phosphorylation of the eukaryotic translation initiation factor 2 (eIF2) induces SG assembly by preventing or delaying translational initiation (Anderson and Kedersha, 2002a; 2009). SGs form when translation is initiated under conditions in which the concentration of the active eIF2–guanosine triphosphate (GTP)–transfer RNA for methionine (tRNAMet) ternary complex is reduced. The assembly of translationally inactive initiation complexes lacking eIF2 allows the RNA-binding proteins TIA-1 or TIAR (or both) to redirect untranslated mRNAs from polyribosomes to SGs. By regulating the equilibrium between polysomes and SGs, TIA-1 and TIAR may influence the frequency with which individual transcripts are sorted for translation or triage in both stressed and unstressed cells (Anderson and Kedersha, 2002a; 2009). Sorting of mRNAs for future translation or decay by individual cells could generate potentially different phenotypes in a genetically identical population (Lavut and Raveh, 2012). The translational arrest that accompanies environmental stress is selective: whereas translation of constitutively expressed “housekeeping” transcripts is turned off, translation of stress-induced transcripts encoding heat shock proteins such as HSP70 and some transcription factors is maintained or enhanced (Anderson and Kedersha, 2002b; 2006; Kedersha and Anderson, 2002; Moore, 2005). mRNAs encoding constitutively expressed “housekeeping” proteins are redirected from polysomes to these discrete cytoplasmic foci, a process that is synchronous with stress-induced translational arrest (Anderson and Kedersha, 2002b; 2008; Kimball et al., 2003; Kedersha et al., 2005). SGs have been shown to contain the Argonaute proteins, microRNAs, a number of mRNA-editing enzymes, and proteins required for transposon activity (Anderson and Kedersha, 2009). “Germ granules” are cytoplasmic, nonmembrane-bound organelles unique to the germline. The term ‘‘germ (or germinal) granule’’ encompasses what are known as P granules in Caenorhabditis elegans and zebrafish, polar granules in Drosophila melanogaster, germinal granules in Xenopus laevis, and the perinuclear nuage in mammalian germ cells. Germ granules are thought to be a signature feature of germ cells in animals (Mahowald, 1968; al-Mukhtar and Webb, 1971; Eddy, 1974; 1975; Wilsch-Brauninger et al., 1997; Snee and Macdonald, 2004; Updike and Strome, 2010; Gao and Arkov, 2012). In some organisms (i.e., Drosophila, Xenopus, Caenorhabditis elegans, and zebrafish), germ granules are present continuously throughout development, with the exception of mature sperm. Germ granules share components with the P-bodies and SGs of somatic cells and often appear as compositional hybrids of SGs and P-bodies (Kotaja et al., 2006; Gallo et al., 2008; Buchan and Parker, 2009; Lim et al., 2009; Voronina et al., 2011; Olszewska et al., 2012; Schisa, 2012). For instance, various isoforms of eukaryotic translation initiation factors are components of SGs and germ granules and are required for gametogenesis in C. elegans, Drosophila, mammals and plants (Amiri et al., 2001; Schisa, 2012; Sengupta and Boag, 2012). The relationship of stress and germ granules is reflected not only by their shared components but also by mechanisms of their formation. Like SGs (see above) germ granules require TIAR for their assembly (Beck et al., 1998). TIAR nullizygotes that survive to birth are sterile because of defective germ cell maturation (Beck et al., 1998). Germinal granules of germline cells in various animals, from sponges to vertebrates, were found to contain a protein product (RNA-helicase) of the vasa gene or related genes, a key determinant and a universal marker of germline cells in metazoans, which is necessary for the formation and maintenance of the structural organization of germ granules. In animals that reproduce only sexually, the expression of vasa-related genes is always exclusively confined to the germ cell line during the entire course of development, from early embryo up to gametogenesis (Ding and Lipshitz, 1993; Ikenishi, 1998; Castrillon et al., 2000; Raz, 2000; Matova and Cooley, 2001; Mochizuki et al., 2001; Extavour and Akam, 2003; Xu et al., 2005: Linder, 2006; Linder and Lasko, 2006; Seydoux and Braun, 2006; Sunanaga et al., 2006; Strome and Lehman, 2007; Ewen- Campen et al., 2010). Intriguingly, Vasa has a role in somatic stress responses of plants (Vashisht and Tuteja, 2006) and Botryllus schlosseri, a colonial urochordate (Rosner et al., 2009), organisms that form germline cells from a multipotent stem cell precursor. Nanos is an evolutionarily conserved RNA-binding protein essential for germ cell development and is a component of germ plasm in organisms, in which germ cells are specified by germ plasm-based ‘preformation’ (Wang and Lehmann, 1991; Subramaniam and Seydoux, 1999; Seydoux and Braun, 2006). Many proteins containing Tudor domains have been identified in stress and germ granules from different model organisms (Gao and Arkov, 2012). The Tudor domain is a small, 50–55 amino acid beta-barrel module that forms a pocket lined with aromatic amino acids (forming an aromatic cage) (Chen C et al., 2011; Pek et al., 2012a). The aromatic cage interacts with methylated amino acids, for example, lysines or arginines, of target proteins. In various model organisms, Tudor domain proteins play crucial roles in the assembly of stress granules (De Leeuw et al., 2007; Goulet, et al., 2008; Linder et al., 2008; Gao et al., 2010; Weissbach and Scadden, 2012) and germ granules (Boswell and Mahowald, 1985; Chuma et al., 2006; Strasser et al., 2008; Vasileva et al., 2009; Siomi et al., 2010; Pillai and Chuma, 2012). During most of germline development, germ granules appear as rounded fibrillar aggregates that cluster around nuclei (Voronina et al., 2011). This arrangement has been documented to persist from the stage of primordial germ cells (PGCs) of Drosophila (Mahowald, 1968; 1971), C. elegans (Strome and Wood 1982, 1983), Xenopus (Ikenishi et al., 1996), and zebrafish (Knaut et al., 2000). In mice, small granules surrounded by fibrillar matrix and mitochondria become apparent around the nuclei of PGCs two days after their formation at 9–9.5 days of gestation (Spiegelman and Bennett, 1973; Clark and Eddy; 1975). The germ granules are in intimate association with mitochondria (and are often described as “intermitochondrial cement or material/bar/cloud”) and, taking into account the role of oxidative stress in SG assembly (Anderson and Kedersha, 2002b; Kedersha and Anderson, 2002; 2007; Gilks et al. 2004; Cohen et al., 2005; McEwen et al., 2005; Shenton et al., 2006; Kedersha and Anderson, 2007; Kedersha et al., 2008, Uniacke and Zerges, 2008), can be expected to be causally related to ROS, particularly H2O2 generation.

The putative function of germ granules in transgenerational epigenetic information transfer will be discussed in chapter 16.

7.2.8 DNA damage and repair

In contrast to other biomolecules DNA cannot be replaced, only repaired. Cells have acquired a variety of DNA repair mechanisms with broad, often overlapping substrate specificities to counteract the harmful effects of DNA injuries (Hoeijmakers, 2001; Friedberg et al., 2006). Increased formation of DSBs brings about a high demand for DNA damage response. Repair of DNA plays a large role in regulating mutant frequency (Klungland et al., 1999; Wang et al., 2006; Ikehata et al., 2007). A vast number of genes encoding proteins that take part in different DNA repair mechanisms show enhanced or specialized expression during spermatogenesis (Hirose et al., 1989; Menegazzi et al., 1991; Weeda et al., 1991; Alcivar et al., 1992; Engelward et al., 1993; Walter et al., 1994; 1996; Chen et al., 1995; Zhou and Walter, 1995; Li WH et al., 1996; van der Spek et al., 1996; Wilson et al., 1996; Mackey et al., 1997; Shannon et al., 1999; Kim et al., 2000; Baarends et al., 2001; Intano et al., 2001; Olsen et al., 2001; 2005; Hsia et al., 2003). For instance, nuclear extracts prepared from mixed germ cells exhibited a substantially higher uracil-DNA glycosylase-initiated BER activity than other mitotically active cells/tissues, namely 5-fold higher than thymocytes and 6-fold higher than small intestine, a finding that cannot be explained simply by germ cell division rates (Intano et al., 2001). The removal of oxidized bases is performed predominantly by the BER pathway (Dianov et al., 2001; Mitra et al., 2001; Bohr et al., 2002) and it has been shown that induction of DNA repair enzymes occurs in response to oxidative stress (Samson and Cairns, 1977; Rusyn et al., 2000; Schärer and Jiricny, 2001; Wood et al., 2001; Barzilai and Yamamoto, 2004; Powell et al., 2005). The link is so close that changes in expression of BER genes have been proposed as sensitive in vivo biomarkers for oxidative DNA damage (Rusyn et al., 2004; Powell et al., 2005). There are a number of consequences of induction/deficiency in DNA repair that are important for the process of mutagenesis. Although induction would lead to enhanced repair, it has been suggested that this can be deleterious and promote mutagenesis (Cairns, 2000). If enzymes that act consecutively on different steps of repair are up-regulated unevenly, a state of imbalanced DNA repair might occur and lead to accumulation of both mutagenic and clastogenic lesions (Glassner et al., 1998; Posnick and Samson, 1999). Thus, at face value, the high levels of DNA repair proteins do an excellent job: spontaneous mutation frequencies are significantly lower for mixed germ cells compared to somatic tissues (Kohler et al., 1991; Dycaico et al., 1994; Walter et al., 1998; Murphey et al., 2013). (But importantly, what was measured in these studies was the small population of high-quality survivors of the germ cell quality-control carnage [e.g. Walter et al., 1998; see chapter 8]).

Given the key role of DNA repair for the faithful transmission of genetic information, DNA repair during mammal gametogenesis, particularly spermatogenesis, is far from optimal. In particular, the BER pathway plays a major role in regulating mutant frequency in the rodent male germline (Huamani et al., 2004; Allen et al., 2008). Importantly, there is no safety margin for BER in the germline: haploinsufficiency of DNA polymerase-beta that is required for the short-patch BER pathway results in reduced BER activity and elevated spontaneous mutagenesis in the mouse male germline but not in somatic tissues (Allen et al., 2008). BER by DNA glycosylases is the main pathway for repair of oxidative base lesions in DNA. In human, rat and mouse nuclear and mitochondrial extracts, the highest DNA glycosylase activities were in testis (Olsen et al., 2001; Karahalil et al., 2002), arguing for the highest oxidative stress in testes (regrettably ovaries were not included in the comparator tissues). The poor removal of oxidized purines (mainly repaired by BER) and bulky DNA adducts (mainly repaired by NER) in human testicular cells are reflected in an accumulation of DNA damage such as 8-oxoG and benzo(a)pyrene adducts in human sperm (Sun et al., 1997; Zenzes et al., 1999; Irvine et al., 2000; Evenson et al., 2002; Olsen et al., 2005). DNA base lesions were preferentially removed from transcriptionally active genes compared to inactive regions (Bohr et al., 1985; Tu et al., 1996).

In vitro and in vivo, ROS exposure results in high frequencies of DNA single-strand breaks (SSBs) and double-strand breaks (DSBs) (Aitken et al., 1998a; Lopes et al., 1998; Kemal Duru et al., 2000; Karanjawala et al., 2002; Mills et al., 2003; Sawyer DE et al., 2003). DNA SSBs are frequent endogenous DNA lesions in human cells (Lindahl, 1993). In normal human cells, it is estimated that approximately 1% of DNA single-strand lesions are converted to approximately 50 endogenous DNA DSBs per cell per cell cycle (Vilenchik and Knudson, 2003). This number is similar to the estimate of the number of exogenous DSBs produced by 1.5–2.0 Gy of ionizing radiation (Vilenchik and Knudson, 2003). In various E. coli and yeast systems, repair of endogenous DSBs and break-induced replication were shown to be highly mutagenic with rates of mutation close to 10-5 per nt (Strathern et al., 1995; Rattray et al., 2002; Rattray and Strathern, 2003; Ponder et al., 2005; Galhardo et al., 2007; Yang et al., 2008; Hicks et al., 2010; Burch et al., 2011; Deem et al., 2011). DSBs are repaired in mammalian cells by two major pathways, namely non-homologous end joining (NHEJ) and homologous recombination (HR) using the sister chromatid as template (Critchlow and Jackson, 1998; Khanna and Jackson, 2001; Pastwa and Blasiak, 2003; Jeggo and Lobrich, 2005; Mao et al., 2008; Moynahan and Jasin, 2010; Brandsma and Gent, 2012). NHEJ and HR pathways are often described as “error-prone” and “error-free”, respectively, but this is an oversimplification (Shrivastav et al., 2008). While “error-prone” NHEJ functions equally well throughout the cell cycle, the “error-free” HR operates in S and G2 phases when sister chromatids become available (Mao et al., 2008). There are two different NHEJ pathways known to date that nevertheless display a similar potential to induce mutations. The canonical pathway, known as DNA-PKcs-dependent NHEJ uses DNA ligase IV, KU70, KU80 and XRCC4 to complete the DNA repair. NHEJ may proceed without some of the canonical factors using PARP1, DNA ligase III and XRCC1 as the alternative ‘back-up’ mechanism known as B-NHEJ (Iliakis, 2009).

Accumulating evidence indicates that DSBs trigger a number of histone modifications around the DSB sites and these modifications may facilitate DSB repair (Lukas et al., 2011; Polo and Jackson, 2011). Oxidative stress induces H2AX phosphorylation in human spermatozoa (Li Z et al., 2006). Phosphorylation of the histone variant H2AX forms gamma-H2AX that marks DNA damage, particularly if the damage involves formation of DSBs (Rogakou et al., 1998; Sedelnikova et al., 2002). Yet, the deposition of H2AX itself may be a better indicator of endogenous DSB hotspots than H2AX phosphorylation (Seo et al., 2012). Within minutes of the induction of DNA DSBs in somatic cells, histone H2AX becomes phosphorylated at serine 139 and forms gamma-H2AX foci at the sites of damage (Rogakou et al., 1998; Cook et al., 2009; Xiao et al., 2009). These foci then play a role in recruiting DNA repair and damage-response factors and changing chromatin structure to accurately repair the damaged DNA (Rogakou et al., 1998; Paull et al., 2000; Hamer et al., 2003; Fernandez-Capetillo et al., 2004) but are also involved in genomic instability (Takahashi and Ohnishi, 2005). Studies of NHEJ efficiency in the nuclear extracts of primary cells from several myeloid malignancies that are associated with increased ROS production, demonstrate a significant increase in repair efficiency, compared to normal hematopoietic cells. However, this increased NHEJ activity was accompanied by a significant increase in the frequency of errors (misrepair) (Gaymes et al., 2002; Brady et al., 2003; Sallmyr et al., 2008). Moreover, abnormally stimulated HR may be leading to mutations (Nowicki et al., 2004).

HR repair activity is present at all stages of mouse male germ cell development, spermatocyte being the most proficient stage (Srivastava and Raman, 2007). In mouse testicular germ cell extracts, NHEJ activity was demonstrated (Sathees and Raman, 1999; Raghavan and Raman, 2004), and NHEJ is active in late spermatocytes (late pachytenes and early diplotenes). Probably there is an interplay between various DSB repair pathways (Goedecke et al., 1999; Lankenau, 2007; Ahmed et al., 2010a). In fact, in Drosophila each stage of germ cell differentiation might exhibit its own characteristic usage of different DNA repair pathways (Preston et al., 2006a; Lankenau, 2007). Moreover, this differential use appears to be modulated by age: HR increased linearly from less than 14% in young individuals to more than 60% in old ones (Preston et al., 2006b; Engels et al., 2007).

Undifferentiated spermatogonia are in the G0/G1 phase of the cell cycle, hence DSBs are expected to be processed predominantly by NHEJ (Rübe et al., 2011). In spermatogonial stem cells lacking compact heterochromatin, histone-associated signaling components of the DNA repair machinery are completely absent and DSBs are rejoined predominantly by DNA-PK-independent pathways, suggesting the existence of alternative repair mechanisms. As a complimentary mechanism, the differentiating progeny, but not the spermatogonial stem cells themselves, promote apoptosis in response to low levels of DNA damage (Rübe et al., 2011). Phosphorylated histone H2AX was detected in a higher proportion of normal fetal gonocytes, and a wider range of adult spermatocyte differentiation stages (Bartkova et al., 2005b). gamma-H2AX occurs in all intermediate and B spermatogonia and in preleptotene to zygotene spermatocytes. Type A spermatogonia and round spermatids do not exhibit gamma-H2AX foci but show homogeneous nuclear gamma-H2AX staining, whereas in pachytene spermatocytes gamma-H2AX is only present in the sex vesicle (Hamer et al., 2003; Forand et al., 2004). DNA DSBs repair proteins differ for the various types of spermatogenic cells, no germ cell type possessing the complete set (Ahmed et al., 2007). The authors concluded that this likely leads to a compromised efficiency of DSB repair relative to somatic cell lines. In addition, the DNA damage response and efficiency of HR-based DNA repair can be expected to be downregulated (Song L et al., 2011) by the abundant expression of miR-18 in testicular germ cells (Björk et al., 2010) (see chapter 10.6.1). Oxidative stress can induce H2AX phosphorylation in human spermatozoa following DSB induction. However, the surveillance system involving gamma-H2AX, Rad50, and 53BP1 in human spermatozoa does not function effectively in DNA repair (Li Z et al., 2006). A significant number of DNA lesions may escape repair. During DNA replication these lesions may cause arrest of the replication fork, and/or the formation of replication gaps, which must be processed to complete replication and enable cell division (Livneh et al., 1993; Kreuzer, 2005; Lopes et al., 2006). One of the central mechanisms to overcome such lesions is translesion DNA synthesis (TLS), also termed translesion replication (TLR), or error-prone repair. The key components in this process are specialized DNA polymerases, characterized by a low fidelity and the ability to replicate across DNA lesions, even bulky ones. Because coding information of the modified bases is usually distorted, TLS is inherently an error-prone process, hence the term error-prone repair: the gap is repaired, but a mutation is produced (Livneh, 2001; 2006; Friedberg, 2005; Prakash et al., 2005; van der Laan et al., 2005; Waters et al., 2009). TLS polymerases are highly expressed in testicular germ cells (McDonald et al., 1999; Yamada et al., 2000; Velasco-Miguel et al., 2003; van der Laan et al., 2005; Sun J et al., 2009a) and are intrinsically required for the long-term maintenance of spermatogenesis (Sun J et al., 2009a).

7.2.9 Mitochondrial ROS, uncoupling and aquaporin-8

Oxidative stress is a hallmark of gametogenesis and other sexual reproduction-related events (Riley and Behrman, 1991; Heininger, 2001; Agarwal et al., 2005; Nedelcu, 2005; Metcalfe and Alonso-Alvarez, 2010; Shkolnik et al., 2011). Oxidative stress regulates mitochondrial respiration (Richter, 1997) and mitochondrial respiration is required for gametogenesis (Treinin and Simchen, 1993; Jambhekar and Amon, 2008). The generation of RONS by sperm and oocyte mitochondria (Riley and Behrmann, 1991; Liu L et al., 2000a; Vera et al., 2004; Liu Z et al., 2006; Koppers et al., 2008; Nabholz et al., 2008a; De Iuliis et al., 2009; Ramalho-Santos et al., 2009) and by non-mitochondrial mechanisms (De Iuliis et al., 2006; Sabeur and Ball, 2007) drive the process. In Drosophila, PINK1/Parkin that regulate mitochondrial function and oxidative stress and have a role in mitochondrial quality control, are required for male and female fertility and in particular for proper sperm maturation (Greene et al., 2003; Pesah et al., 2004; Clark et al., 2006).

Increased mtDNA rearrangements and deletions in human gametes (Brenner et al., 1998; Reynier et al., 1998; Barritt et al., 1999) witness this oxidative stress exposure. The mitochondrial mutation rate is orders of magnitude higher than the nuclear one (Brown et al., 1979; 1982; Ballard and Whitlock, 2004; Lynch et al., 2006). This hypermutability largely explains the high level of within-species mtDNA heterozygosity (Bazin et al. 2006) and the high amount of homoplasmy detected in animal mtDNA phylogenetic analyses (Springer et al., 2001; Delsuc et al., 2003; Galewski et al., 2006). The mtDNA neutral substitution rate varies by 2 orders of magnitude across mammalian species (Nabholz et al., 2008a). Mammal and bird species maximal longevity correlate with mitochondrial mutagenesis and evolutionary patterns of mitochondrial DNA diversity (Nabholz et al., 2008a; 2009), arguing for the fundamental role of a mitochondrial molecular clock in both the germline and soma and its role for phylogenetic and organismal dynamics and life history trajectories (Heininger, 2001; 2012). Mitochondria play a key role in steroidogenesis (Yacobi et al., 2007). The first and rate-limiting step in the biosynthesis of steroid hormones in the adrenals and gonads is the transfer of cholesterol across the inner mitochondrial space from the outer mitochondrial membrane to the inner mitochondrial membrane, a process known to depend on the action of the steroidogenic acute regulatory protein (StAR) (Christenson et al., 2001; Stocco, 2001; Diemer et al., 2003a; Miller WL, 2007). Mitochondrial membrane potential, mitochondrial ATP synthesis, and mitochondrial pH are all required for acute steroid biosynthesis, suggesting that mitochondria must be energized, polarized, and actively respiring to support Leydig cell steroidogenesis (Allen JA et al., 2006). Steroidogenesis in Leydig cells produces ROS, largely from mitochondrial respiration and the catalytic reactions of the steroidogenic cytochrome P450 enzymes (Quinn and Payne, 1984; 1985; Peltola et al., 1996; Hales, 2002; Hanukoglu, 2006). Besides functioning as key enzymes in steroidogenesis, mitochondrial P450-type enzymes can function as a futile NADPH oxidase and leak electrons, thus producing superoxide and ROS (Hanukoglu, 2006). The primary source of ROS in sperm is the respiratory chain of the mitochondrion (Turrens et al., 1985; Gavella and Lipovac, 1992; Cadenas and Davies, 2000; Baker et al., 2005). The main ROS form in male germ cells is hydrogen peroxide (H2O2), the concentration of which in sperm has not been measured (Jones et al., 1979; Sharma and Agarwal, 1996; Aitken et al., 1998a). Both the perinuclear clustering of mitochondria in germline cells and long-range action of H2O2 are compatible with a potential role of H2O2 in the nuclear compartment during gametogenesis.

Mitochondrial ROS activate mitochondrial uncoupling proteins (Echtay et al., 2002a; b; Brand et al., 2004). Uncoupling proteins (UCPs) are phylogenetically conserved (Stuart et al., 1999; Vercesi, 2001; Czarna and Jarmuszkiewicz, 2005) mitochondrial inner membrane proteins that are important in regulating cellular fuel metabolism and as attenuators of ROS production through strong or mild uncoupling (Boss et al., 2000; Sluse et al., 2006; Azzu et al., 2010; Mailloux et al., 2011; Mailloux and Harper, 2011). UCPs uncouple proton flux through the inner mitochondrial membranes and ATP synthesis, providing a route for proton re-entry (Mattiasson and Sullivan, 2006), lowering the proton-motive force (Brand and Esteves, 2005). UCP2 and UCP3 diminish oxidative stress by lowering the mitochondrial membrane potential (Nègre-Salvayre et al., 1997; Gustafsson et al., 2004; Brand and Esteves, 2005). Acute increases in ROS production increase proton conductance through both UCP2 and UCP3, providing a negative feedback loop to limit further mitochondrial ROS formation (Echtay et al., 2002a; b). UCP2 was detected in all germ cells under normal conditions, predominantly in elongate spermatids (Ricquier and Bouillaud, 2000; Zhang et al., 2007), possibly expressed in response to prooxidant TNFα (Lee FY et al.,1999). Germ cell UCP2 is upregulated by aging (Amaral et al., 2008), and hyperthermia (Zhang et al., 2007) and protected the cells from ROS-induced apoptosis (Zhang et al., 2007). Although the physiological role of UCP2/3 in a variety of tissues is still controversial (Pecqueur et al., 2001; Brand et al., 2004; Cannon et al., 2006; Echtay, 2007; Azzu et al., 2010; Mailloux and Harper, 2011; 2012; Shabalina and Nedergaard, 2011; Sluse, 2012), testicular UCP2 expression appears to be a marker of oxidative stress (Pecqueur et al., 2001; Brand et al., 2004; Echtay, 2007; Zhang et al., 2007; Amaral et al., 2008; Mailloux and Harper, 2011; 2012).

Mitochondrial aquaporins may facilitate the diffusion of H2O2 across biological membranes, modulating the outcome of cellular stress. Aquaporins (AQPs), are a family of small integral membrane proteins that primarily transport water across the plasma membrane (Agre and Kozono, 2003). There are 13 members (AQP0–12) in humans. They are divided into three subgroups based on the primary sequences: water selective AQPs (AQP0, 1, 2, 4, 5, 6, 8), aquaglyceroporins (AQP3, 7, 9, 10), and superaquaporins (AQP11, 12) (Ishibashi et al., 2009). The striking physicochemical similarity between water and H2O2 points to AQPs as likely candidates for H2O2 permeation in plants and animals (Henzler and Steudle, 2000; Bienert et al., 2007; Hooijmaijers et al., 2012). Experimental evidence suggests that certain AQPs act as peroxoporins and thus facilitate the diffusion of H2O2 across biological membranes (Henzler and Steudle, 2000; Bienert et al., 2007; Dynowski et al., 2008). AQPs are also involved in swelling of tissues under stress (Verkman, 2005; 2008). The AQP8 paralogue is characterized by an unusual primary structure, a location of the exon–intron boundaries in the gene different from those of other AQPs, suggesting that AQP8 has a separate phylogenetic origin from other AQPs (Zardoya and Villalba, 2001). AQP8 shares the highest sequence homology (38–40%) with the plant water channel AQP-TIP (Ishibashi et al., 1997; Ma et al., 1997; Koyama et al., 1998). AQP8 seems to be present in most metazoans (Zardoya and Villalba, 2001; Campbell et al., 2008; Tomkowiak and Pienkowska, 2010; Ishibashi et al., 2011). Membrane diffusion of H2O2 is facilitated by aquaporins AQP3 and AQP8 in mammals (Bienert et al., 2007; Miller et al., 2010) and TIP1;1 and PIP2;1 in plants (Bienert et al., 2007; Dynowski et al., 2008). Certain isoforms of Arabidopsis thaliana AtPIPs, including AtPIP2;2, AtPIP2;4, AtPIP2;5, and AtPIP2;7, are also permeable for H2O2 (Hooijmaijers et al., 2012). AQP8 expression appears to be modulated by conditions involving oxidative stress (Yamamoto et al., 2001; Hardin et al., 2004; Yamamoto et al., 2007; te Velde et al., 2008; Yang M et al., 2009; Kortner et al., 2012). AQP8 is intracellular (Ishibashi, 2006), particularly localized at the inner mitochondrial membrane (Ferri et al., 2003; Calamita et al., 2005; 2007; Lee et al., 2005; La Porta et al., 2006; Lee and Thévenod, 2006; Gena et al., 2009) and endoplasmatic reticulum (Ferri et al., 2003). Although AQP8 has been reported to transport water (Liu K et al., 2006), overall water or glycerol permeabilities of mitochondria isolated from AQP8 knockout were reported not to differ from those of mitochondria from wildtype mice (Yang et al., 2006). Moreover, the overall water permeability of brain mitochondria that lack AQP8 corresponds to that of liver and testis mitochondria with abundant AQP8 (Calamita et al., 2006) suggesting that water transport may not be the primary function of AQP8. Mitochondrial AQP8 may have a role in mitochondrial H2O2 release, mitochondrial permeability transition and cell death (Lee and Thévenod, 2006; Younts and Hughes, 2009; Marchissio et al., 2012). In fact, AQP8 upregulation may sensitize hepatocellular carcinoma cells to apoptotic insults (Jablonski et al., 2007; Younts and Hughes, 2009). Intriguingly, while gastrointestinal inflammation was found to downregulate AQP8 expression (Hardin et al., 2004; Yamamoto et al., 2007; te Velde et al., 2008; Kortner et al., 2012), neuronal ischemia and lactacystin (a proteasome inhibitor)-induced neuronal toxicity and apoptosis upregulated AQP8 expression (Chen et al., 2008; Yang M et al., 2009). Intriguingly, AQP8 is abundantly expressed by male and female gonadal tissues and germline cells (Ishibashi et al., 1997; Calamita et al., 2001a; b; Elkjær et al., 2001; Kageyama et al., 2001; McConnell et al., 2002; West-Farrell et al., 2009; Yeung, 2010). In the developing rat testis, AQP8 mRNA was first detected between 13 and 16 days post-partum, consistent with the reported first appearance of spermatocytes (13–14 days) and AQP8 protein was observed in adult rat spermatocytes (Kageyama et al., 2001). In the developing rat testis, transcripts of AQP7 became detectable between 23 and 25 days post-partum, when round spermatids have been reported to appear (Calamita et al., 2001b; Kageyama et al., 2001). However, genetic deletion of AQP7 and 8 in mice did not result in obvious abnormalities in sperm morphology and function (Yang B et al., 2005; Sohara et al., 2007; Yeung et al., 2009). However, AQP8-null mice had increased testicular weight/size and a greater ratio of spermatogenic cells to Sertoli cells in seminiferous tubules but water content measured by wet-to-dry weight ratios showed no significant difference between AQP8-null and wild-type mice (Yang B et al., 2005). The latter findings may indicate a role of AQP8 in germline cell apoptosis. AQP8 is expressed in granulosa cells and upregulated by follicle stimulation hormone (McConnell et al., 2002; Mock et al., 2004; Su W et al., 2010) but is not expressed in oocytes (Edashige et al., 2000). AQP8 deficiency significantly lowered apoptosis rate in AQP8-/- granulosa cells and increased corpus luteums in mature AQP8-/- ovaries suggesting increased maturation and ovulation of follicles (Su W et al., 2010), thus increasing the fertility of female mice (Su W et al., 2010; Sha et al., 2011). During an 8 days’ period of follicle growth in vitro, antrum formation and steroid production were monitored, and mRNA was isolated. Follicles that developed an antrum and had a steroidogenic profile similar to that of follicles in vivo had unchanging expression of AQP8 and HIF-1α. Follicles that did not form an antrum or produce appropriate levels of estrogen and progesterone, however, demonstrated increasing levels of AQP8 and HIF-1α, a pattern in gene expression resulting in decreased follicle differentiation and oocyte quality (West-Farrell et al., 2009).

7.2.10 Membrane lipid unsaturation

In naturally occurring polyunsaturated fatty acids (PUFAs), the hydrogen atoms (called bis-allylic hydrogens) attached to the single bonded –C– atom that separate the double-bonded carbon units (–C=C–) have the lowest C–H bond-energies of the fatty acid chain. This makes them the most susceptible to attack by ROS produced during aerobic metabolism (Halliwell and Gutteridge, 2007). The more polyunsaturated the fatty acid, the more bis-allylic hydrogens it has and consequently the more prone it is to oxidative attack by metabolism-produced free radicals. Docosahexaenoic acid (DHA; 22:6), which has six double bonds and consequently five bis-allylic hydrogens per chain, is 320-times more susceptible to ROS attack than the common monounsaturated oleic acid (18:1) which has no bis-allylic hydrogens in its chain (Hulbert, 2010). PUFAs increase membrane proton permeability and the activity of membrane-associated metabolically active proteins (Brand et al., 1994; Porter et al., 1996; Brookes et al., 1998; Wu et al., 2001; 2004; Else and Hulbert, 2003; Turner N et al., 2003; 2005a; b) and these observations have led to the development of the ‘membrane pacemaker theory’ of metabolism (Hulbert and Else, 1999; 2000; 2004; 2005; Hulbert, 2003; 2007; 2008). It proposes that highly polyunsaturated acyl chains impart physical properties to membrane bilayers that enhance and speed up the molecular activity of membrane proteins and consequently the metabolic activity of cells and tissues (Hulbert, 2008). Vertebrate sperm are rich in PUFAs (Aitken et al., 1989a; b; Surai et al., 1998; Zalata et al., 1998; Lenzi et al., 2000; Bréque et al., 2003; Wathes et al., 2007) which renders their membranes particularly susceptible to free radical attack (Jones et al., 1979; Alvarez et al., 1987; Aitken et al., 1989a; b; Twigg et al., 1998; Lenzi et al., 2000). Unsaturated fatty acids not only are passive targets of oxidative stress. They may stimulate ROS generation and lipid peroxidation in spermatozoa and oocytes (Aitken et al., 2006; Wakefield et al., 2008; Koppers et al., 2010). A positive correlation was found between ROS production and lipid peroxidation (Rodrigues et al., 2010). Particularly, elevated PUFA levels enhance cellular ROS level in hypoxia, most likely by impairing the electron flux within the respiratory chain (Schönfeld et al., 2011). On the other hand, some PUFAs have a dual role in activating both UCP expression and uncoupling activity and have been shown to be the most potent UCP activators (Zackova et al., 2003; Beck et al., 2007; Rodrigues et al., 2010). Superoxide-induced lipid peroxidation leads to the production of reactive aldehydes, including 4-hydroxynonenal, the end-product of the lipoperoxidation cascade of omega-6 PUFAs (Echtay et al., 2003). These aldehydic lipid peroxidation products are in turn able to modify proteins such as mitochondrial uncoupling proteins and the adenine nucleotide translocase, converting them into active proton transporters. This activation induces mild uncoupling and so diminishes mitochondrial superoxide production (Echtay et al., 2005). Delta-6 desaturase-null mice (-/-) are unable to synthesize PUFAs. The -/- males exhibit infertility and arrest of spermatogenesis at late spermiogenesis (Stoffel et al., 2008; Stroud et al., 2009). DHA supplementation was capable of restoring all observed impairment in male reproduction (Roqueta-Rivera et al., 2010). ELOVL2 is a member of the mammalian microsomal ELOVL fatty acid enzyme family, involved in the elongation of very long-chain fatty acids including PUFAs required for various cellular functions in mammals. The lack of Elovl2 was associated with a complete arrest of spermatogenesis, with seminiferous tubules displaying only spermatogonia and primary spermatocytes without further germinal cells (Zadravec et al., 2011).

7.3 Male-driven mutagenesis

The testis has been appreciated as the engine of evolution (Agulnik et al., 1997; Short, 1997; Hales et al., 1999; Kleisner et al., 2010). In fish, birds and mammals, point mutations arise preferentially in the male gametes. Higher rates of mutation in the male germline, relative to the female germline, have been found in primates (Shimmin et al., 1993; Makova and Li, 2002), rodents (Chang et al., 1994), cats (Pecon Slattery and O’Brien, 1998), birds (Ellegren and Fridolfsson, 1997), fish (Ellegren and Fridolfsson, 2003), Arabidopsis (Whittle and Johnston, 2003) and gymnosperms (Whittle and Johnston, 2002). The male excess compared to mutations in female gametes has been estimated approximately five- to ten-fold in humans, six-fold in primates, two-fold in rodents, up to six-fold in birds (Shimmin et al., 1993; Chang et al., 1994, Redfield, 1994; Chang and Li, 1995, Crow, 1997b; 2000; Ellegren and Fridolfsson, 1997; 2003; Huang et al., 1997; Anagnostopoulos et al., 1999; Kahn and Quinn, 1999; Carmichael et al., 2000; Erlandsson et al., 2000; Fridolfsson and Ellegren, 2000; Li WH et al., 2002; Makova and Li, 2002; Bartosch-Härlid et al., 2003; Axelsson et al., 2004; Chimpanzee Sequencing and Analysis Consortium, 2005; Sandstedt and Tucker, 2005; Ellegren, 2007; Kong et al., 2012; Sun et al., 2012; Grégoire et al., 2013). These figures suggest that male bias may have accelerated along the mammal lineage. Data derived from studies in autosomal- and X-linked diseases, although variable, strongly support the notion of a male-driven mutagenesis (Risch et al., 1987; Bonaiti-Pellie et al., 1990; Rosendaal et al., 1990; Montandon et al., 1992; Oldenburg et al., 1993; Thompson and Chen, 1994; Tuchman et al., 1995; Becker et al., 1996; Thomas, 1996; Green et al., 1999; Ljung et al., 2001; Flodman and Hodge, 2003; Glaser and Jabs, 2004; Heron et al., 2010; O’Roak et al., 2012). Comparison of rates of evolution for X-linked and autosomal pseudogenes suggests that the human male mutation rate is 4 times the female mutation rate (Nachman and Crowell, 2000). However, there may be a substantial variance in human sex-specific mutation rates. In one study, 92% of germline de novo mutations were from the paternal germline, whereas, in contrast, in the other family, 64% of de novo mutations were from the maternal germline (Conrad et al., 2011). Recently, evidence for a male-to-female mutation rate of approximately 2 for the first invertebrate, Drosophila, has been presented, suggesting that DNA sequence evolution is male-driven in a wide variety of taxa (Bachtrog, 2008). Also, the ratio of partial to complete male compared to female sterility loads in Drosophila (0.5) is similar to that in the butterfly Bicyclus anynana (0.45) (Saccheri et al., 2005). Moreover, in dioecious plants (species with independent sexes) there is also evidence for higher rate of mutation on the Y than the X chromosome (Filatov and Charlesworth, 2002; Whittle and Johnston, 2002; 2003) and a higher recombination rate in male gametes (Robertson, 1984; Zhuchenko et al., 1989; Busso et al., 1995).

Following Haldane’s (1947) discovery that most mutations in human hemophilia are male-derived, point mutations leading to Lesch–Nyhan syndrome (Francke et al., 1976), hereditary retinoblastoma (Dryja et al., 1989; 1997; Zhu et al., 1989; Matsunaga et al., 1990; Kato MV et al., 1994; Munier et al., 1996), achondroplasia (Risch et al., 1987; Wilkin et al., 1998), Hutchinson–Gilford progeria syndrome (Eriksson et al., 2003; D’Apice et al., 2004), Muenke syndrome (craniosynostosis) (Rannan-Eliya et al., 2004), Noonan syndrome (multiple congenital anomaly syndrome) (Tartaglia et al., 2004), von Hippel-Lindau disease (familial cancer syndrome) (Richards et al., 1995), Rett syndrome (neurodevelopmental disorder) (Girard et al., 2001; Trappe et al., 2001; Zhu et al., 2010), multiple endocrine neoplasia Type B (Carlson et al., 1994; Schuffenecker et al., 1997), Apert syndrome (achrocephalosyndactyly) (Moloney et al., 1996), Crouzon syndrome and Pfeiffer syndrome (craniosynostotic disorders) (Glaser et al., 2000), Costello syndrome (mental retardation and predisposition to benign and malignant tumors) (Sol-Church et al., 2006), Townes–Brocks syndrome (malformation syndrome characterized by anal, renal, limb, and ear anomalies) (Böhm et al., 2006), Dravet syndrome (severe infantile epileptic encephalopathy) (Heron et al., 2010), and CHARGE syndrome (coloboma, heart defects, atresia of the choanae, retarded growth and development, genital hypoplasia, ear anomalies and deafness) (Pauli et al., 2012) were found to be either exclusively or largely paternal. Epidemiological data also suggest that paternally derived genetic damage may contribute significantly to the etiology of cancer in children and young adults (Aitken, 1999; Crow, 2000). Further evidence for the male mutation bias are findings that genes with detectable expression only in male reproductive tissues, evolve rapidly in many taxa, including Drosophila (Zhang Z et al., 2004a; Pröschel et al., 2006; Ellegren and Parsch, 2007; Haerty et al., 2007; Larracuente et al., 2008; Assis et al., 2012) and mammals (Clark and Swanson, 2005; Ellegren and Parsch, 2007).

It has been conventional interpretation that this male excess is owing to a difference in the number of germline replications (Penrose, 1955; Miyata et al., 1987; Crow, 1997a; Hurst and Ellegren 1998; Li WH et al., 2002; Ellegren, 2007). These differences have been accounted to the male-to-female ratio of the number of germ cell divisions per generation suggesting that errors in DNA replication are the primary source of mutations (Chang et al., 1994; Drost and Lee, 1995; Li WH et al., 1996; Hurst and Ellegren, 1998; Kahn and Quinn, 1999; Vogel and Motulsky, 2010). In the human male the number of cell divisions from zygote to a sperm is about 150 at age 20, 380 at age 30 and 610 at age 40, compared to only about 24 cell divisions from zygote to an ovum in human females (Vogel and Motulsky, 1997). More than 20 spontaneous congenital disorders have been reported to be associated with advanced paternal age (Jones et al., 1975; Glaser and Jabs, 2004; Arnheim and Calabrese, 2009; Sayres and Makova, 2011; Goriely and Wilkie, 2012; see also chapter 8.2.1) supporting the notion that the number of germ cell divisions is a key factor for the male mutation bias. Base substitutions are likely the most prominent type of male-biased mutations to arise from mitotic cell divisions because of the mis-incorporation of nucleotides by DNA polymerase (Johnson RE et al., 2000). Other types of mutations such as insertions and deletions were also found to be male biased from whole genome studies in rodents (Makova et al., 2004). Since insertions and deletions can occur from strand slippage during replication, this lent further support to the potential contribution of the higher number of cell divisions in the male mutation bias (Grégoire et al., 2013).

However, recent data indicate also a significant contribution of number of germline replication-independent mutagenic events in male germline cells (Agulnik et al., 1997; Crow, 1997b; Hurst and Ellegren, 1998; 2002; Huttley et al., 2000; Tiemann-Boege et al., 2002; Baker and Aitken, 2005; Blumenstiel, 2007; Pink et al., 2009; Yoon et al., 2009; Pink and Hurst, 2010; Grégoire et al., 2013). For instance, there was no age-dependent increase in the frequency (Kato et al., 2007) of a well-known de novo non-Robertsonian translocation, t(11;22)(q23;q11), of paternal origin (Ohye et al., 2010), suggesting that these translocations are independent of replication. Particularly, there is now substantial evidence that chromatin remodeling steps during post-meiotic events of spermatogenesis may not be genetically inert but could represent an evolutionary conserved, replication-independent, process that may act more specifically to introduce de novo mutations, insertions and deletions, or even chromosomal rearrangements such as translocations (Grégoire et al., 2013). An intriguing model presented by Blumenstiel (2007) concluded that assuming a tradeoff between producing large numbers of sperm and expending energetic resources in maintaining a lower mutation rate, sperm competition would select for males that produce larger numbers of sperm despite a higher resulting mutation rate. There is male-biased mutation for classes of mutations that are likely independent of DNA replication numbers (Huttley et al., 2000; Blumenstiel, 2007; Pink et al., 2009). For example, a male bias has been observed in Arabidopsis for the transmission of mutations that arise from UV irradiation (Whittle and Johnston, 2003). By adjusting for the fact that over evolutionary time an autosome is equally likely to reside in a male or female germline cell whereas an X chromosome experiences a female cell environment 66% of the time (33% in a male environment), the sequence divergence between the X chromosomes of human and chimp can be compared to estimate the autosomal divergence (X/A ratio). Increased expression of several polymerases (pol kappa, pol lambda, and pol iota) that support mutagenic translesion DNA synthesis is observed in male meiotic mid-pachytene cells and in postmeiotic (round spermatid) cells (McDonald et al., 1999; Garcia-Diaz et al., 2000; Gerlach et al., 2000; Yamada et al., 2000; Velasco-Miguel et al., 2003; Bavoux et al., 2005; van der Laan, 2005). Conceivably, these specialized DNA polymerases generate genetic variability in the male germline (Friedberg et al., 2002). Moreover, cumulative evidence indicates that polymerases kappa, lambda, and iota may be involved in somatic hypermutation (Poltoratsky et al., 2001; Friedberg et al., 2002; Bavoux et al., 2005). A strong male mutation bias for neutral nucleotide substitutions was detected at non-CpG sites: alpha (the male-to-female mutation rate ratio) in the X-autosome comparison was ~6–7, which was similar to the male-to-female ratio in the number of germline cell divisions (Taylor et al., 2006). Male mutation bias was equally strong at CpG and non-CpG sites, suggesting the replication-dependent origin of these mutations (Taylor et al., 2006). In contrast, mutations at CpG sites, where mutations typically result from replication-independent deamination of methylated cytosines, exhibited weak male mutation bias: alpha in the X-autosome comparison was only ~2–3. The study also indicated weak male mutation bias for transversions at CpG sites, implying a spontaneous mechanism largely not associated with replication. Moreover, a weak male-biased mutation rate has also been found in Drosophila: mutagenic transposons are often transmitted paternally in an active state but maternally in a repressed state (Bingham et al., 1982; Bucheton et al., 1984; Yannopoulos et al., 1987; Evgenev et al., 1997; Vieira et al., 1998). Computer simulations (Redfield, 1994) showed that the cost of excess male mutations can easily exceed the benefit of recombination, costs that would add to the various costs of sex listed in chapter 1.

In testes, a signalling role of hydrogen peroxide has been identified (Fujii and Tsunoda, 2011). Not all H2O2 produced within the mitochondrial matrix will survive to efflux from the mitochondria, owing to matrix peroxidases that consume H2O2 (Zoccarato et al., 2004; Andreyev et al., 2005; Rhee et al., 2005a; Murphy, 2009). These include peroxiredoxins (Rhee et al., 2001), catalase (Radi et al., 1991; Salvi et al., 2007) and glutathione peroxidases (Imai and Nakagawa, 2003). A striking feature of the rat germ cell antioxidant system is a very high SOD activity associated with low levels of glutathione peroxidase (GPx), glutathione S-transferase (GST), and glutathione reductase (GR) activity (Bauché et al., 1994), with GPx, GST and GR playing key roles in cellular defence against oxidative damage (Raes et al., 1987; Baker et al., 1988; Simmons and Jamall, 1988; Ketterer and Meyer, 1989; Mirault et al., 1991; Lavoie et al., 1992; Miller and Blakely, 1992). Moreover, no catalase activity that may be able to metabolize H2O2, was detected in testicular germ cells (Bauché et al., 1994). Intriguingly, the prooxidant actions of p53 appear to be dependent on its p53-inducible gene 3 (PIG3)-mediated downregulation of catalase activity (Kang MY et al., 2013, see chapter 8.2). A new family of antioxidative proteins, collectively referred to as peroxiredoxins (PRDX), have been identified (Rhee et al., 2005b). Six distinct gene products are known for the PRDX family in mammals. the most well-characterized function of Prx family members is the ability to modulate hydrogen peroxide signalling in response to various stimuli (Rhee et al., 2005b). PRDX2, the fastest regenerating redox protein (Chevallet et al., 2003), is more efficient in neutralizing H2O2 than catalase and GPx (Peskin et al., 2007) and is more effective in protecting cells from H2O2 damage than GPx (Berggren et al., 2001). PRDX2 transcripts and proteins occur in Leydig cells and Sertoli cells of mouse testis, while spermatogonia and spermatocytes apparently do not express them (Lee K et al., 2002), but spermatids and mature spermatozoa (Manandhar et al., 2009). PRDX4 plays a role in inhibition of NF-kappaB function as a cytosolic factor (Jin et al., 1997), but it also can activate NF-kappaB as an extracellular factor (Haridas et al., 1998). Based on observations in rat testes at puberty, a supportive function of the membrane-bound form of Prx4 in acrosome formation during spermiogenesis was proposed (Sasagawa et al., 2001a). PRDX4 knockout results in elevated spermatogenic cell death via oxidative stress (Iuchi et al., 2009). As compared to the liver, whole rat testes express equivalent SOD activity but only 5% of the GPx liver levels and 2% of the liver catalase activity (Peltola et al., 1992). Testicular guinea pig SOD levels were found to be about twice as high as liver SOD in the same species, whereas GPx, GR and catalase were about 60, 5, and 300 times less, respectively (Kukucka and Misra, 1993). Intriguingly, catalase expression is downregulated by ROS via methylation of a CpG island in the Oct-1 promoter (Quan et al., 2011). In contrast to its role as an antioxidant, overexpression of SOD, unbalanced by enzymes that can metabolize H2O2, has been shown to increase hydroxyl radical formation and to elevate steady-state hydrogen peroxide levels (Peled-Kamar et al., 1995; 1997). SOD has also been shown to catalyze the formation of hydroxyl radical from hydrogen peroxide (Yim et al., 1990; 1993). Increased MnSOD expression leads to increased H2O2 production (Rodriguez et al., 2000; Buettner et al., 2006; Sarsour et al., 2008). A number of reports indicate that either an increase or decrease in CuZnSOD/MnSOD activity may cause oxidative damage, apoptosis and cell cycle arrest (Oberley et al., 1981; Elroy-Stein et al., 1986; Avraham et al., 1988; 1991; Elroy-Stein and Groner, 1988; Norris and Hornsby, 1990; Amstad et al., 1991; Nelson SK et al., 1994; Rothstein et al., 1994; Troy and Shelanski, 1994; Kelner et al., 1995; Peled-Kamar et al., 1995; 1997; Bernard et al., 2001; Kim A et al., 2004; 2005; 2010; Sarsour et al., 2008). In accordance with these findings, transgenic male mice expressing higher levels of mitochondrial MnSOD have reduced fertility (Raineri et al., 2001). The copper-zinc superoxide dismutase (CuZnSOD) gene resides on chromosome 21 and is overexpressed in Down syndrome patients. Down syndrome patients show a multitude of abnormalities related to increased oxidative stress (Peled-Kamar et al., 1995; de Haan et al., 1997; Jovanovic et al., 1998; Iannello and Kola, 2001; Pallardó et al., 2006; Perluigi and Butterfield, 2011) caused by the imbalance in the SOD/Gpx ratio (de Haan et al., 1997; 2003). These findings indicate that mammal spermatogonia, pachytene spermatocytes and round spermatids are able to convert O2-• to H2O2 by SOD whereas they may encounter major difficulties to further metabolize hydrogen peroxide, as well as organic peroxides, to unreactive molecules. The difference between SOD and glutathione-dependent enzyme activity in germ cells may lead to saturation of the protective systems against peroxides. As a result, the H2O2 formed may be available for conversion into highly toxic hydroxyl radicals via Fenton-type reactions (see chapter 4.1) (Bauché et al., 1994). As has been pointed out by Bauché et al. (1994), the consequences can be especially dramatic in the testis, with the risk of germ cell mutation and heritable mutation. Intriguingly, high levels of antioxidants had adverse effects on the progression of normal germ cell differentiation and male fertility (Ten et al., 1997; Shalini and Bansal, 2005; 2007; Brigelius-Flohe, 2006; Puglisi et al., 2007), arguing for the physiological role of a certain amount of oxidative stress in spermatogenesis.

The proteomic analysis of sperm differentiation reveals that male germ cells express a multitude of genes that are involved in stress responses including heat shock proteins, hypoxia-inducible factor, and DNA repair (Aguilar-Mahecha et al., 2001; Gupta, 2005; Martínez-Heredia et al., 2006). Circumstantial evidence indicates that at least part of the excess male germ cell mutation rate is due to, compared to female, a high level of male developmental oxidative stress: during spermatogenesis germ cells massively undergo apoptosis (Rodriguez I et al., 1997; Blanco-Rodríguez, 1998; Print and Loveland, 2000) induced by oxidative stress (Erkkilä et al., 1998; 1999). Metabolic shifts from mitochondria-produced ATP to glycolysis affecting ROS generation occur at several stages of male gametogenesis (Ramalho-Santos et al., 2009). In addition to mitochondrial sources, an enzymatic system for ROS generation located in the mammalian sperm plasma membrane that utilizes the reduced adenine dinucleotides (NAD(P)H) as a substrate via an NAD(P)H-dependent oxidase has been suggested as one mechanism for ROS-mediated signaling in spermatozoa (Aitken et al., 1992; 1995; 1997; Ball et al., 2001; Armstrong et al., 2002; Sabeur and Ball, 2006; 2007). NADPH oxidases are the only known enzyme family with the sole function to produce ROS (Altenhöfer et al., 2012). NADPH oxidase 5 (NOX5) is highly expressed in pachytene spermatocytes, round and elongating spermatids (Bánfi et al., 2001; Sabeur and Ball, 2007) indicating a role of this isoform in meiosis and sperm maturation.

Vertebrate sperm are rich in polyunsaturated fatty acids (Aitken et al., 1989a; b; Surai et al., 1998; Zalata et al., 1998; Lenzi et al., 2000; Bréque et al., 2003; Wathes et al., 2007) which renders plasma membranes particularly susceptible to free radical attack (Alvarez et al., 1987; Aitken et al., 1989a; b; Twigg et al., 1998; Lenzi et al., 2000). Moreover, the fact that germ cells are not well equipped to combat oxidative stress or xenobiotic-mediated injury probably explains the high sensitivity of these cells to oxidative stress (Gomes, 1970; Le Grande, 1970; David et al., 1971; 2005; Saini, 1997; Rockett et al., 2001; Sikka, 2001; Chakir et al., 2002; Saleh and Agarwal, 2002; Araripe et al., 2004; Rohmer et al., 2004; Vollmer et al., 2004; Jensen et al., 2006; Jørgensen et al., 2006; Gupta et al., 2007; Podrabsky et al., 2008; Sakata and Higashitani, 2008; Hansen, 2009; Crespo and Shivaprasad, 2010; Hales and Robaire, 2010; Prasad et al., 2011) and the extremely limited viability of isolated spermatogenetic cells in culture (Chapin and Phelps, 1990). Several groups have demonstrated that, in contrast to spermatogonia, pachytene spermatocytes and round spermatids, elongated spermatids and spermatozoa have a reduced capability or are even unable to repair DNA damage (Ono and Okada, 1977; Van Loon et al., 1991). A variety of thioredoxins and glutathione peroxidases, oxidative stress-regulating systems, and Hsps are uniquely expressed in male spermatogenesis (Miranda-Vizuete et al., 2004; Gupta et al., 2007; Chabory et al., 2010). Particularly during meiotic synaptonemal complex formation when recombination occurs (Loidl, 1994; Egel, 1995; McKim et al., 1998), Hsps play an essential role (Allen et al., 1996; Dix et al., 1996; Dix, 1997; Iliopoulos et al., 1997; Berruti et al., 1998; Eddy, 1999; Son et al., 1999). Hsp70-2 knockout leads to failed meiosis, germ cell apoptosis and infertility in male but not female mice (Dix et al., 1996; Eddy, 1999) arguing for a specific spermatocyte vs. oocyte vulnerability. Although eleven per cent of proteins found in human spermatozoa are involved in protection against oxidative damage, apoptosis and cell cycling (Martinez-Heredia et al., 2006), they possess a limited arsenal of cytosolic antioxidant defences relative to somatic cells (Dowling and Simmons, 2009; Zini et al., 2009). Although, this shortfall seems to be counterbalanced by a high antioxidant capacity of the seminal fluid (Aitken, 1994; Smith R et al., 1996; Baker and Aitken, 2004), together with a range of seminal proteins (Menella and Jones, 1980; Jeulin et al., 1989; Nanogaki et al., 1992; Schöneck et al., 1996; Collins et al., 2004) that probably play a role in protecting spermatozoa.

In mice, a testes-specific form of cytochrome c is able to catalyse the reduction of H2O2 three times faster than its somatic counterpart, and a testes-specific form of cytochrome c is also more resistant to degradation by H2O2 than its somatic form (Liu Z et al., 2006). On the other hand, the apoptotic activity of testes-type cytochrome c is three to five times greater than the somatic type, thus playing a much more stringent role in apoptotic ‘‘quality control’’ (Liu Z et al., 2006). Testicular-type cytochrome c null mice are fertile but present with highly atrophied testes, with a reduced number of spermatocytes, spermatids and spermatozoa (Narisawa et al., 2002) possibly indicating that testicular-type cytochrome c is required to detoxify the increased level of H2O2 and/or to enforce a stringent ‘‘quality control’’. The increased oxidative stress exposure of spermatozoa in comparison to oocytes is also epitomized by their defective maintenance of mtDNA (Reynier et al., 1998) which may underlie the almost exclusive maternal inheritance of mtDNA (Ohno, 1997a). 8-hydroxydeoxyguanosine (8-OHdG) is the most abundant oxidative DNA adduct and a key biomarker of oxidative DNA damage (Ames et al., 1993) and was found highest in 6-week-old male rat testes (Nakae et al., 2000), particularly in leptotene, zygotene, and early pachytene spermatocytes (Sakai et al., 2010).

The Fenton reaction (see chapter 4.1) may play a role in male germ cell mutagenesis (Wellejus et al., 2000). Cellular DNA damage under prooxidant conditions has been shown to be mediated by iron. Conversely, oxidative stress itself influences iron metabolism and iron proteins. Intracellular iron levels are increased in response to oxidative stress (Udipi et al., 2012). In fact, iron is an important element in the establishment of a prooxidant status in the cell (Meneghini, 1997; Kruszewski, 2003). Sertoli cells make transferrin which is an iron transport protein as part of a proposed shuttle system that effectively transports iron around the tight junction complexes to the developing germ cells (Sylvester and Griswold, 1994). Iron metabolism is compartmentalized and closely regulated to protect developing male germ cells from iron fluctuations (Leichtmann-Bardoogo et al., 2012). Seminal plasma transferrin levels have been proposed as a functional parameter of Sertoli cells (Barthelemy et al., 1988; Cek et al., 1992; Irisawa et al, 1993). In humans, many studies show that the transferrin level in the seminal plasma is correlated with sperm yield (Sueldo et al., 1984; Orlando et al., 1985; Barthelemy et al., 1988; Zalata et al., 1996). A low concentration of transferrin in the seminal plasma has been correlated with severity of oligospermia (Orlando et al., 1985; Zalata et al., 1996). On the other hand, iron was found to accumulate in the sperm and other testes cells (Hoyes et al., 1995; Lucesoli et al., 1999; Doreswamy and Muralidhara, 2005). Sperm-associated iron is constantly lost from the testicular iron pool; consequently there must be continuous transfer of iron from the general circulation to adluminal germ cells. In adult male rats, approximately 0.25% of injected 114mIndium, a transferrin-binding radionuclide, localised within the testis by 48 h postinjection and remained constant for up to 63 d. In neonates, 0.06% of the activity was in the testis by 48 h, and this declined such that by 63 d only 0.03% remained (Hoyes et al., 1995). Various models of iron overload, e.g. beta-thalassemia, hemochromatosis (that is associated with increased risk of cancer), and a variety of animal models, predispose sperm to oxidative injury and impair fertility (Siemons and Mahler, 1987; Lucesoli et al., 1999; Perera et al., 2002; Lourdes de Pereira and Garcia e Costa, 2003; Doreswamy and Muralidhara, 2005; Aitken and Roman, 2008). Female patients with beta-thalassemia major and hypogonadotrophic hypogonadism with diminished ovarian reserve responded favorably to gonadotrophins with 80% success rate while males responded less favorably than females (Bajoria and Chatterjee, 2011). The mitoferrin gene product and other iron metabolism proteins show enriched expression in the testes of mammals and insects, suggesting special roles for mitochondrial iron metabolism in spermatogenesis (Hales, 2010). Studies in vertebrates and yeast demonstrate a role for mitoferrin family members in import of iron from the cytosol into mitochondria (Zhang Y et al., 2006; Paradkar et al., 2009). Mutations in the Drosophila mitoferrin gene, resulted in male sterility, but did not display other gross abnormalities and no female sterility (Metzendorf and Lind, 2010). A role of the free iron pool in oxidative DNA damage and genetic instability has been suggested (Gackowski et al., 2002; Olinski et al., 2003). Iron metabolism is regulated in response to oxidative stress coordinated with oxidative stress defenses (Pantopoulos and Hentze, 1995; Hanson and Leibold, 1998; Zheng et al., 1999). Iron accumulation dependent on metabolic and oxidative stress (Romslo, 1975; Fujimoto et al., 1982; Ceccarelli et al., 1995; Wang et al., 1995) may sensitize cells to oxidative stress (Lipinski et al., 2000). Translocated to the nucleus, iron and other trace metals may replace zinc in zinc finger domains of transcription factors, generate free radicals and induce transcription-dependent mutations (Omichinski et al., 1993; Sarkar, 1995; Conte et al., 1996). Further regulatory influences may stem from the type of metal ion catalyzing the Fenton reaction, protection of DNA by histones and chromatin structure (Chiu et al., 1993; Oleinick et al., 1994). Epidemiological data also indicate that elevation of the body iron level may increase the risk of cancer (Nelson RL et al., 1994; Stevens et al., 1994; Huang, 2003).

In nature there is a striking inverse relationship between stress resistance and metabolic and proliferating activity. Well known examples are survival forms like spores, cysts and seeds, which display high stress resistance and virtually no metabolic activity. In yeast cells there is a clear inverse correlation between stress resistance and growth and metabolic activity (Schenberg-Frascino and Moustacchi, 1972; Plesset et al., 1987; Van Dijck et al., 1995; Lu et al., 2009), indicating that there might be an incompatibility at the molecular level between high stress resistance and high metabolic activity (Van Dijck et al., 2000). In invertebrates, e.g. insects, the spermatogenic stress manifests by a, compared to females, higher susceptibility of male gametogenesis to temperature stress (David et al., 2005). The high level of germ cell proliferational and maturational stress may also explain the conspicuous susceptibility of mammalian male germ cells to additional stressors, e.g. heat and, for instance, their lower temperature threshold for Hsp activation (Precht et al., 1955; Sarge, 1995; Rockett et al., 2001; Banks et al., 2005; Setchell, 2006) and susceptibility to abdominal temperature in cryptorchidism that leads to infertility (Shikone et al., 1994; Yin et al., 1997; Lue et al., 1999). In the human scrotum the temperature is approximately five degrees lower than body temperature. The poor vascularization of the mammalian testes is requisite for the testicular lower temperature as can be inferred from varicocele, or dilation of the spermatic vein. Varicocele typically occurs on the left side only and is associated with an increase in male infertility (Fretz and Sandlow, 2002). Experimental left varicocele bilaterally increases testicular temperature in lab animals and causes a reduction in testicular sperm output (Turner, 2001). The unilateral lesion in humans also bilaterally increases testicular temperature (Goldstein and Eid, 1989) and establishes a trend toward increased blood flow (Ross et al., 1994). Pathogenesis of cryptorchidism infertility is mainly attributable to high testes temperature because in situ cooling of abdominal testes in pigs results in normal spermatogenesis (Frankenhuis and Wensing, 1979). Moreover, elevated lipid peroxidation has been demonstrated in a mouse model of experimentally induced cryptorchidism (Peltola et al., 1995). Oxidative stress is a major cause for thermal damage of mouse spermatogenic cells and leads to apoptosis and DNA strand breaks (Banks et al., 2005; Pérez-Crespo et al., 2008; Paul et al., 2008a; 2009). Thus, chronic heat stress such as in cryptorchidism (Ishii et al., 2005) causes ROS levels to rise beyond what can be managed by the antioxidative systems in the epididymis, resulting in spermatogenic cell death. Intuitively, it may be expected that testicular stress should upregulate sperm DNA repair mechanisms. However, following heat stress at 43° C, murine expression of a number of DNA repair genes such Ogg1 (involved in base excision repair), Xpg (involved in nucleotide excision repair) and Rad54 (involved in double-strand break repair) were all down-regulated (Rockett et al., 2001; Paul et al., 2008b). In response to heat stress, decreased expression of polyADP-ribose polymerase (PARP) in the rat testis (Tramontano et al., 2000) was reported; PARP proteins are involved in detection of strand breaks and signalling in both the base excision repair and nucleotide excision repair pathways (Schreiber et al., 2002; Flohr et al., 2003). In addition, a decreased expression of oxidative stress-induced antioxidants (Rockett et al., 2001) may leave the germ cells more susceptible to oxidative damage during hyperthermia (Paul et al., 2008b). Heat stress induced by cryptorchidism appears to result in decreased expression of DNA polymerase beta and DNA ligase III both of which are involved in the final stages of DNA repair, for example, in both base and nucleotide excision repair (Tramontano et al., 2000). Intriguingly, Lupu et al. (2004) found a relationship between DNA repair efficiency and thermotolerance in isofemale lines of Drosophila melanogaster originating from ‘Evolution Canyon’ (Mt Carmel, Israel), with thermotolerant lines tending to repair DNA more efficiently than thermosensitive ones. Likewise, thermosensitivity of cancer cells is thought to be associated with their genetic instability (Wheldon, 1977).

Male spermatogenesis in a wide variety of taxa is highly temperature-sensitive (Precht et al., 1955; Cowles, 1965). Paul et al. (2008a) reported that in vitro fertilization with sperm recovered from male mice in which the scrotum was heated to 42oC resulted in embryos with reduced ability to complete development. Females mated to males exposed to scrotal heating had conceptuses with smaller fetal and placental weights compared with controls (Jannes et al., 1998; Paul et al., 2008a). In modern mammals, the most sensitive steps with regard to temperature appear to be the maintenance of spermatogonial stem cells and the survival of gametes through meiosis and spermatid differentiation (Setchell, 1998; Ivell, 2007). Most mammals need 2 to 10°C below core body temperatures for viable sperm production and maturation (Moore, 1926; Cowles, 1958; 1965; VanDemark and Free, 1970; Waites, 1970; Bedford, 1977; Carrick and Setchell, 1977; Knobil and Neill, 1995). For spermatogenesis to proceed, a lower temperature achieved by the migration of the testes into the scrotum, is necessary. In addition, the pampiniform plexus is a highly efficient countercurrent heat exchanger in which the arterial blood is precooled before it reaches the testis, while venous blood is warmed to body temperature before it returns to the abdomen (Rommel et al., 1998; Brackett, 2004).

Part of the high reproductive success of mammals has been attributed to the fact that their immature offspring are protected inside the mother’s womb. On the other hand, there is no adaptive interpretation for the phenomenon of mammalian testicular descent (David et al., 2005). Indeed, an external location of the testes makes them prone to a diversity of wounds or accidents and may expose them to ionizing, mutagenic radiation (Oakberg, 1959; Ash, 1980; Otala et al., 2004; Kleisner et al., 2010; Wright, 2010; Merrifield, 2011), and it is easy to argue that internal protection should be favored by natural selection (David et al., 2005). It has been speculated that evolution of the scrotum occurred because of the need for low temperatures either for spermatogenesis, sperm storage or to minimize mutations in gamete DNA (Moore, 1926; Short, 1997; Werdelin and Nilsonne, 1999; Bedford, 2004). In agreement with this idea, in several mammalian species testes and epididymides remain in the body cavity and descend to the scrotum, i.e. to a lower temperature, only for the duration of the breeding season (Precht et al., 1955). Such seasonal testicular migrations are seen in chiropterans (Krutzsch, 1955; Marshall and Corbet, 1959; Krutzsch and Crichton, 1987; Jolly and Blackshaw, 1988), insectivores (Marshall, 1911), rodents (Rasmussen, 1917; Moore et al., 1934), carnivores (Koudele, 1986) and primates (Ramakrishna and Prasad, 1967). In some species, such as dogs, the scrotum can become hairless and acquire a dark coloration to aid heat radiation, emphasizing again the physiological importance of a cool scrotum (Ivell, 2007). Rhinos and tapirs have subcutaneous testes which lie in a scrotal-like sac external to the abdominal wall (Ottow, 1955). Importantly, subterranean mammals underwent a convergent loss of scrotum (Burda, 2003; Nevo, 1999). Mammals that inhabit an aquatic environment (cetacean, hippos, seals and walruses) lack a scrotum (Kleisner et al., 2010). Since spermatogenesis is just as threatened by testicular cooling as it is by testicular heating (Zhang Z et al., 2004b), aquatic mammals need to have intra-abdominal testes in order to keep them warm at all times. Dolphins, seals and Florida manatees possess a vascular countercurrent or anastomotic heat exchanger that functions to cool their intra- or para-abdominal testes (Rommel et al. 1992; 1995; 2001; Pabst et al., 1995). Elephants have abdominal testes, considered as a primitive trait (Kleisner et al., 2010). There is palaeontological, anatomic and genetic evidence suggesting that they may have had an aquatic origin in the distant past (de Jong, 1998; Gaeth et al., 1999; Inuzuka, 2000; Glickman et al., 2005). The elephant’s ancestors probably had intra-abdominal testes for a long time since there is no vestige of the countercurrent heat exchange mechanism – the pampiniform plexus – found in all scrotal mammals and no evidence of testicular migration or descent into a scrotum (Short et al., 1967; Gaeth et al., 1999). This is in marked contrast to the situation in cetaceans (whales and dolphins), which are thought to have evolved from a terrestrial (presumably scrotal) artiodactyl ancestor about 60 M years ago, and, although they have retracted their testes back into the abdomen, they have retained a well-developed pampiniform plexus (Glickman et al., 2005). In female mammals temperature gradients between mature Graafian follicles and ovarian stroma for human, rabbit, pig, and cow generally fell in the range of 1.3–1.7°C: follicles were always cooler than stroma. Temperature gradients are maintained locally by counter-current heat exchange mechanisms (Hunter et al., 2006).

Birds, for aerodynamic streamlining possess intra-abdominal testes. Cooling of the testes occurs via the adjacent air-sacs (into which inspired air is first drawn) (Cowles, 1965). However, the thoracic air sacs do not appear to be required for normal spermatogenesis (Williams, 1958; Herin et al., 1960). In breeding males of some species of birds, particularly passerine birds, the caudal sperm-storage region of the vas deferens is expanded into a cloacal protuberance which contains the convoluted portion of the vas deferens outside the abdomen (Salt, 1954; Wolfson, 1954). In sparrows and juncos the seminal vesicles protrude, during the sexual season, into the cloaca where the temperature is about 4°C lower than that of the body cavity (Wolfson, 1954; Hafez, 1964). Only sperm obtained from these evaginated vesicles are motile. There is also evidence that avian spermatogenesis may proceed mainly at night when body temperature is lowest and periods of hyperthermia are not likely to be experienced (Riley, 1937; 1940; Miller, 1938; Murton et al., 1970a; b; Greenwood and Wheeler, 1985). The domestic fowl, however, may have a unique system of heat shock tolerance mediated by both the amount and polyadenylation of Hsp70 and ubiquitin transcripts (Mezquita et al., 1998) that allows spermatogenesis at temperatures deleterious e.g. to mammals (Beaupré et al., 1997). Development of spermatogenesis in young fowl appears to be speeded up at temperatures lower than the abdominal temperature, but the process does not seem to be favored by very low temperatures (Williams, 1958). Species that need a high aerobic capacity, such as flighted birds could not get airborne at all if they did not have an aerobic capacity several-fold higher than even fast runners like the cheetah. In birds, the apoptotic threshold is low: they are sensitive to ROS leak from mosaic respiratory chains and quickly trigger apoptosis, translating into infertility and low fecundity (Lane, 2011a). Compared to mammals, the mitochondrial ROS leak of birds is nearly 10-fold lower (Barja, 2007). Taking into account that mitochondrial ROS generation is temperature-dependent (Zar and Lancaster, 2000; Abele et al., 2002; Heise et al., 2003; 2006; Mujahid et al., 2005; 2007), these features may explain the higher temperature tolerance of avian spermatogenesis.

Lizards kept in warm temperatures quickly exhibit testicular collapse (Cowles and Burleson, 1945). In male lizards, the range of temperatures preferred by each species appears to be critically adjusted to a level compatible with the maximum tolerable temperature for the testes, for 3 weeks 10 hr/day exposure to temperatures 1-2 °C higher resulting in marked spermatogenic damage (Licht, 1965).

Intriguingly, poikilotherm fishes display a male lower-temperature comfort zone. Male guppies prefer a significantly lower temperature (24.5°C) than females (28.2°C) or juveniles (28.1°C). Treatment of juveniles and females with testosterone lowers their preferred temperature to that of males (Johansen and Cross, 1980). This gender-differential temperature preference, at least during the reproductive period, has been observed in other fish species as well (Hagen, 1964; Baker et al., 1970; Swain and Morgan, 2001; Hernández-Rodríguez et al., 2002; Podrabsky et al., 2008). In rainbow trout, the maximal heat shock response of male germ cells, that are located in the same body compartment like the other organs, occurs at a significantly lower temperature (22°C) than for somatic cells (28°C) (Le Goff and Michel, 1999). These findings strongly suggest the existence of a particular mode of heat shock susceptibility of male germ cells that is not restricted to homeotherms (Le Goff and Michel, 1999).

Taken together, there is a clear trend from lower to higher taxa for an increasing male gametogenic stress that renders male gametogenesis susceptible to additional stressors, e.g. temperature stress (see chapter 14.1).

7.3.1 Chromatin remodeling in elongating spermatids

In this chapter, I will largely adhere to the excellent review of Grégoire et al. (2013). During spermiogenesis, the round haploid spermatids undergo a major morphological differentiation program characterized by one of the most dramatic changes in chromatin structure known to the eukaryotic world. Most of the histones are replaced by protamines providing both mechanical and chemical stability to the mature sperm chromatin (Ward, 2011). The molecular mechanism leading to such a striking nuclear transition is yet poorly understood but relies on histone variants, key post-translational modifications and general degradation of histones (Laberge and Boissonneault, 2005; Govin et al., 2006, 2007; Awe and Renkawitz-Pohl, 2010; Grégoire et al., 2011). The chromatin remodeling steps are specifically associated with transient, endogenous DNA strand breaks that are detected in the whole population of both mouse and human spermatids (Sakkas et al., 1995; Marcon and Boissonneault, 2004; Leduc et al., 2011a). Comet assays (Collins, 2004) in neutral conditions showed a clear accumulation of DSBs specifically in the nuclear DNA of spermatids throughout elongation. The origin of these DSBs is not clear but they may be created by several, mutually non-exclusive, possibly synergistic, mechanisms: enzymatically, for instance by type II topoisomerases (McPherson and Longo, 1993; Laberge and Boissonneault, 2005; Meyer-Ficca et al., 2011), from the activity of ROS (Muratori et al., 2006; Sakkas and Alvarez, 2010), or simply from the mechanical stress induced by the change in chromatin structure (Boissonneault, 2002; Muratori et al., 2006; Sakkas and Alvarez, 2010). Topoisomerase II appears as the unique enzyme responsible for the transient double-stranded breaks in elongating spermatids but depends on histone hyperacetylation for its activity (Laberge and Boissonneault, 2005). Chromatin remodeling in spermatids involves massive withdrawal and degradation of histones that should leave transient free DNA supercoils (Boissonneault, 2002). Such a high degree of free superhelical density is likely to generate non B-DNA structures which can be responsible for breakpoint hotspots and chromosomal rearrangements (Wang G et al., 2008). For instance, Z-DNA, characterized by a left-handed instead of a typical right-handed double helical structure, is generated within regions of high negative supercoiling and may serve as a recognition signal for DSB formation (Kha et al., 2010). High density of free supercoils independent of replication can produce cruciform extrusion that may also act to signal breakpoints involved in translocations (Inagaki et al., 2009). Transient DSBs were observed in spermatids of both human, rat and mouse (Marcon and Boissonneault, 2004; Meyer-Ficca et al., 2005; Leduc et al., 2008), but evidence of transient DSBs in spermatids was also reported in fruit flies (Drosophila melanogaster) (Rathke et al., 2007), in grasshopper (Eyprepocnemis plorans) (Cabrero et al., 2007) as well as in algae (Chara vulgaris) (Wojtczak et al., 2008). Taken together, these observations point to an evolutionarily conserved, physiological mechanism. These physiological DSBs trigger a repair response, based on the detection of the phosphorylated H2AX histone variant (gammaH2AX) and in situ detection of DNA polymerase activity in elongating spermatids (Leduc et al., 2008, 2011b). The presence of DSBs in this haploid context would necessarily prevent homologous recombination to be used as a reliable, template DNA repair mechanism that depends on sister chromatids as this is the case during the S phase in somatic cells (Grégoire et al., 2013). NHEJ processes must therefore be used in order to repair DSBs in spermatids but, based on studies in somatic cells, these mechanisms are associated with limited insertions or deletions at the repair site which alter the DNA sequence, although the structural integrity of the DNA is restored. Even from a homogeneous set of starting DNA ends as substrate, NHEJ creates important variations in the non-templated addition at the two DNA ends (Lieber, 2010). There are two different NHEJ pathways known to date that nevertheless display a similar potential to induce mutations. The canonical pathway, known as DNA-PKcs-dependent NHEJ uses DNA ligase IV, KU70, KU80 and XRCC4 to complete the DNA repair. NHEJ may proceed without some of the canonical factors using PARP1, DNA ligase III and XRCC1 as the alternative ‘back-up’ mechanism known as B-NHEJ (Iliakis, 2009). PARP1 is involved in oxidative-stress response pathways (Virág, 2005). In round spermatids both NHEJ pathways are active (Ahmed et al., 2008; 2010b; Rübe et al., 2011; Grégoire et al., 2013). In addition to having to rely on error-prone DNA repair systems, the chromatin remodeling context is likely to create an impediment to the repair process and the overall DNA repair capacity was found to decrease as spermatids progress through their differentiation program (Olsen et al., 2005; Marchetti and Wyrobek, 2008; Ahmed et al., 2010b). The spermatids that are formed after meiotic division II are increasingly transcriptionally and translationally silenced during the condensation of the chromatin. Thus these post-meiotic spermatids have a minimal capacity for DNA repair (Sega, 1974; 1979; Sotomayor et al., 1978, 1979; Tanaka and Katoh, 1979; Zbinden, 1980; Sega and Sotomayor, 1982; Working and Butterworth, 1984; Sotomayor and Sega, 2000) and therefore their DNA could be damaged in a cumulative manner, as the sperm cannot respond by inducing either apoptosis or DNA repair, as they are transcriptionally silent.

7.3.2 The gametogenesis-cancerogenesis connection

Gamete formation has been likened to a benign, slowly advancing but in the end lethal sort of cancer that is fought against by the soma (De Loof, 2011). Germline cell differentiation is controlled by a specific set of genes whose expression is tightly locked into the repressed state in somatic cells. Large-scale epigenome alterations, now evidenced in nearly all cancers, lead to aberrant activation of these normally silenced genes, as attested by the many reports describing the expression of testis-specific factors, known as cancer/testis (CT) genes, in various cancer cells. In normal adult testis, expression of CT antigens is present in spermatogonia and, to a variable degree, in later stages of sperm cell maturation (Takahashi et al., 1995; Jungbluth et al., 2000a; b; 2001a; 2002; Gjerstorff et al., 2006). In fetal ovary, immature germ cells (oogonia/primary oocytes) express CT antigens, whereas oocytes in the resting primordial follicles do not (Jungbluth et al., 2001b; Nelson et al., 2007). CT genes comprise more than 240 members from 70 families, and can be subdivided into two broad categories based on chromosomal localization on the X chromosome and on autosomes. The observation of shared characteristics between germline cells and tumor cells has led to the concept that recapitulation of portions of the germline gene-expression programme might contribute characteristic features to the neoplastic phenotype, including immortality, invasiveness, hypomethylation and DNA instability (Old, 2001; Zendman et al., 2003; Kalejs and Erenpreisa, 2005; Simpson et al., 2005). One of the distinguishing and near-universal hallmarks of cancer growth is hypoxia. Unregulated cellular proliferation leads to formation of cellular masses that extend beyond the resting vasculature, resulting in oxygen and nutrient deprivation. The resulting hypoxia triggers a number of critical adaptations that enable cancer cell survival, including apoptosis suppression, altered glucose metabolism, and an angiogenic phenotype. Oxygen depletion stimulates mitochondria to elaborate increased ROS, with subsequent activation of signaling pathways, such as HIF1alpha, that promote cancer cell survival and tumor growth (Fruehauf and Meyskens, 2007). Elevated rates of ROS have been detected in almost all cancers, where they promote many aspects of tumor development and progression (Dreher and Junod, 1996; Klaunig et al., 2010; Liou and Storz, 2010; Sosa et al., 2013).

Off-context activity of some of the testis-specific epigenome regulators can reprogram the somatic cell epigenome toward a malignant state by favoring self-renewal and sustaining cell proliferation under stressful conditions, thereby constituting a major oncogenic mechanism (Cheng et al., 2011; Wang J et al., 2011). Intriguingly, the expression of CT genes in a variety of cancers correlates with and may be causally involved in the DNA instability of cancers and thus contribute to tumor progression (Bodey, 2002; Iwata et al., 2005). Moreover, CT gene expression is associated with hypomethylation and transcriptional activation (De Smet et al., 1999; Rosty et al., 2002; Sato et al., 2004; Lee YM et al., 2006; Ye et al., 2005; Ehrlich, 2009), hypoxic response activation (Aprelikova et al., 2009; Kuphal et al., 2010), microsatelite instability, defective DNA mismatch repair (Iwata et al., 2005; Okada et al., 2005) and anti-apoptosis (Cilensek et al., 2002; Monte et al., 2006). A link between DNA hypomethylation, genomic instability and carcinogenesis has been established (Chen et al., 1998; Saito et al., 2002; Fan et al., 2003; Kanai, 2010). Ectopic expression of germline-specific genes can drive tumor growth in Drosophila (Janic et al., 2010) and Hydractinia (Millane et al., 2011). It was proposed that the aberrant expression of these genes by cancer cells confers a range of phenotypic traits that are essential for the survival and function of gametes and their descendents. These gamete-specific products would be deleterious for the orderly requirements of normal somatic cells, but highly advantageous for the cancer cell (Simpson et al., 2005).

8. Sex:...and selection of the "pearls among the pebbles"


One is left with the feeling that some essential feature of the situation is being overlooked.

J. Maynard Smith, 1976a

Summary

The competitive advantage of organisms depends on their stress resilience and acquisition of limited resources. The selection principles that work at the level of individuals are also pervasively deployed at the cellular level during development, immune surveillance and cancerogenesis. Cell competition, modulated by redox balance, is an efficient mechanism for selecting cell quality and thereby ensuring that the requisite cellular tasks will be done by the most efficient and competitive cells. Population genetic calculations have predicted that germ cell competition-based selection, if existent, would act as a sieve eliminating deleterious mutations and increasing the frequency of beneficial ones, reducing the genetic load imposed on the population by several orders of magnitude. A general pattern emerges: sperm that are less costly and subjected to a harsher (epi-)mutagenic process also are subjected to a more stringent selection program. Overall, selection regimes can be distinguished at more or less discrete levels: germ cells, gametes, fertilization, embryos/offspring, nonrandom mating. Their high developmental stress damages sperm mitochondria and they are eliminated subsequently. In oocytes, mitochondria pass through a selective bottleneck that maintains mitonuclear coadaptation and rewinds Muller’s ratchet.

If, as outlined in chapter 10, sexual reproduction introduces (epi)genetic variation by a plethora of (epi)mutations, the evolution of sex had to resolve a major dilemma: (epi)mutagenesis not only creates potential beneficial mutations but carries the much higher hazard of detrimental effects (Morgan, 1903; Sturtevant, 1937; Kibota and Lynch, 1996; Thatcher et al., 1998; Boe et al., 2000; Imhoff and Schlotterer, 2001) on the viability of germline cells. For its evolutionary success, sexual reproduction had to remove the excess genetic load. In microbial colonies with their billion-fold individuals (of which each can be viewed as a germ cell) chances are good that by the stochastic mutagenic process single organisms may have acquired beneficial mutation(s) that favor their survival in the selective environment of their harsh habitat. With the advent of multicellular organisms and the segregation of a germline, evolution “faced” the fundamental problem of evolvability. Kirschner and Gerhart (1998) defined evolvability as “the capacity to generate nonlethal phenotypic variation” (my bold type). In the light of the stochasticity of mutagenesis the key issue is the viability of the mutants. Importantly, no evidence suggests so far that sexually or asexually reproducing organisms have the capacity to direct or choose which genetic variants will arise (Sniegowski and Lenski, 1995) (but see the concept of interpretive mutations as advanced by Jablonka and Lamb, 2005). The ecological conditions decide whether a specific genetic change is beneficial or detrimental. Reproductive organs have no sensors and feedback control (but see chapter 12) to sense the environmental conditions and to control the fit of the mutation to these environmental conditions. To create large numbers of mutants and let natural selection decide over their viability is the unicellular strategy. In multicellular organisms this would mean a large investment into possibly poorly viable and reproductively less fit organisms. In resource-limited environments (see chapter 3) this wouldn’t have been an optimal strategy (and would hardly have established the evolutionary success of sexual reproduction). A targeted approach would be to invest as little as possible into small, mutated, organisms and have them selected by natural selection or select them internally for their quality. In fact, the former approach is realized by external fertilizations of a variety of marine organisms and plants (Serrão et al. 1996; Yund, 2000) while the latter approach is realized by internal fertilizations in a multitude of terrestrial and aquatic organisms. The censor to differentiate between beneficial and deleterious mutations and the gametes carrying them is cellular function and thus eventual survival. Cell survival is regulated by redox signaling (Heininger, 2001; Trachootham et al., 2008; Groeger et al., 2009; Fulda et al., 2010). Thus, oxidative and nitrosative stress serves the dual purpose to effect gametogenetic mutagenesis and gamete quality assurance. And exposing the gametes to a stressful environment introduces a selective bottleneck selecting the most stress resistant and resilient of them. In contrast to asexual reproduction with its direct reproductive lineage, sex evolution had to introduce an additional selection process which subjects particularly the sperm to a rigid ‘quality control’ (a feature which has received little attention in all sex evolution theories) (Sutovsky et al., 2001; Sutovsky, 2003). In fact, spontaneous death of germ cells is a widespread phenomenon in the testes of many metazoan species (Roosen-Runge, 1973; Sasso-Cerri et al., 2006; Riesgo et al., 2008). The degree and the modes of the cell elimination are species-specific, but in general the abnormalities appear to originate during gametic mitosis and meiosis (Roosen-Runge, 1973). To exert a rigid quality control sexual reproduction “relies” on five levels of selection: germ cell-, gamete-, fertilization-, embryo/offspring- and sexual-selection.

8.1 Selection: the pervasive phenomenon in evolution

In a system like the immune system, moulded by cellular selection, precision of the end product is made possible by introducing the highest possible degree of randomness at the earliest stages.

Michaelson, 1987

Populations have the potential to grow exponentially, but this is confronted with the limited nature of resources (Heininger, 2012). That populations outgrow the available resources is the central idea of Malthus’s An Essay on the Principle of Population (1798). The study of this work led Darwin to the conclusion that the pressure exerted by limited resources is a natural form of selection, analogous to breeder’s artificial selection (Ruse, 2009). On all levels of biological organization, competition for scarce resources is a pervasive driver of evolution (Loreau, 1998; Fisher and Hoekstra, 2010; Heininger, 2012). Natural selection susequently rewards those individuals who compete best for the scarce resources and can use them most economically (Grover, 1997). Thus, natural selection is an outcome of this competition (Fisher 1930; Endler, 1986; Bock, 2003; 2010).

Gamete overproduction requires a gamete bottleneck: a stochastic bottleneck inevitably would drive Muller’s ratchet and result in mutational meltdown; only a selection-associated bottleneck ensures long-term viability of populations. Competition is pervasive at every level of life—in ecology, economics, between countries and states, and in families—and helps to determine order, status, and survival (Johnston, 2009). The possibility that cells of multicellular organisms may also compete with one another has been postulated several times over the past two centuries (Díaz and Moreno, 2005). In 1881, Wilhelm Roux proposed the idea of a cellular struggle for survival during development (Roux, 1881; Heams, 2012). Wilhelm Roux transferred Charles Darwin’s theory of the struggle for existence to the fight among cells and ‘‘parts’’ of the organism in the process of ontogenesis. As evidence for the conflict between cell types, he referred to pathological processes in which cells of one tissue start to invade another (Roux, 1881). His idea received no acclaim, since cells within multicellular organisms were thought to display conflict mediation/repression between, and cooperation of, the different cell types because cooperation increases the fitness of the group (Michod, 1996; 2005; Frank, 2003b). A common hypothesis is that the unicellular bottleneck of the germ cell acts as a conflict mediator, by increasing the kinship among cells in the organism, thereby aligning the interests of cells with the interests of the organism (Bell and Koufopanou, 1991; Maynard Smith and Szathmáry, 1995; Grosberg and Strathmann, 1998). Repression of competition within social groups has been suggested as a key mechanism driving the evolution of cooperation and the major evolutionary transitions (Leigh, 1977; Alexander, 1979; 1987; Buss, 1987; Maynard Smith, 1988; Maynard Smith and Szathmáry, 1995; Szathmáry and Maynard Smith, 1995; Frank, 1995; 2003; 2009; Ratnieks et al., 2006; Gardner and Grafen, 2009). As shown in bacterial communities, repression of competition per se, as opposed to increased relatedness, is driving the observed increase in cooperation (Kümmerli et al., 2010). In bacteria, hypermutability accelerates the breakdown of cooperation due to increased sampling of genotypic space, allowing mutator lineages to generate non-cooperative genotypes (Harrison and Buckling, 2005; 2011), and cheat on the others (Vulic and Kolter, 2001), a phenomen that also may underlie cancerogenesis (Heininger, 2001). However, competition is also a strong coevolutionary force resulting in selection of fitter individuals (Zambrano et al., 1993; Heininger, 2012)

Tilman (1982) defined competition as ‘an interaction between individuals brought about by a shared requirement for a resource in limited supply leading to a reduction in the survivorship, growth, and/or reproduction rates of the competing individuals concerned’. According to Welden and Slauson (1986) “competition is the induction of strain in one organism as a direct result of the use of resource items by another organism”. Darwin imagined, in his last paragraph of the Origin of Species, a tangled bank of competing organisms, and it now seems that we can stretch his analogy to the dynamic interactions of cells that populate niches during development and repair (Green, 2010). Thus, Darwinian principles of variation and selection can be extended to sub-organismal entities, e.g. organelles, cells and the germ-soma competition (Edelman, 1987; Stoner et al., 1999; Heininger, 2001; 2012; Weiss, 2006). There are several established criteria accepted as evidence of competition among populations (McLean et al., 1997; Gaudin et al., 2004). For example: (i) The presence of competitors should modify the equilibrium size of a population. (ii) The presence of competitors should alter the dynamics of a population, e.g. the life expectancy of the individuals of the population. (iii) It should be possible to modify the equilibrium dynamics of two competing populations through the manipulation of the available resources.

Cell competition was discovered in the imaginal discs of D. melanogaster almost 40 years ago (Morata and Ripoll, 1975). It initially described a situation in which slowly dividing cells were eliminated by apoptosis from a population of more rapidly dividing cells (Morata and Ripoll, 1975; Simpson, 1979; Simpson and Morata, 1981; Moreno et al., 2002; Lolo et al., 2012), despite the fact that they would have been viable on their own (Morata and Ripoll, 1975; Simpson, 1979; Simpson and Morata, 1981; Moreno et al., 2002; de la Cova et al., 2004; Li and Baker, 2007; Moreno, 2008). Thus, competition is context-dependent—cells acquire “winner” or “loser” identity only when in confrontation; each is viable in a homotypic environment (Johnston, 2009; Baker, 2011; Lolo et al., 2012). Cellular competition also occurs in Drosophila tracheal branching morphogenesis (Ghabrial and Krasnow, 2006) and in postmitotic epithelial tissue repair (Tamori and Deng, 2013). It seems unlikely that such an effective mechanism to select for cell fitness should be confined to flies (Díaz and Moreno, 2005). In fact, cell competition has now been firmly established in a variety of taxa, including mammals (Oliver et al., 2004; Oertel et al., 2006; Sansom et al., 2007; Bondar and Medzhitov, 2010; Marusyk et al., 2010; Tamori et al., 2010; Baker, 2011; Kim et al., 2011; Krueger et al., 2011; Merlo et al., 2011; Petrova et al., 2011; de Beco et al., 2012; Norman et al., 2012). In vitro findings suggest that cell competition outcome is modulated by redox balance (Merlo et al., 2011) and activation of the Jun N-terminal kinase (JNK) stress-response pathway (Moreno et al., 2002; de la Cova et al., 2004; Moreno and Basler, 2004). Consistent with the role of these signaling pathways, competition intensity increases in high-intensity competitive environments (Chesson and Huntly, 1997; Violle et al., 2010; Miller et al., 2011). Competition among cells provides an efficient mechanism for selecting cell quality and thereby ensuring that the requisite cellular tasks will be done by the most efficient cells (Abrams, 2002; Díaz and Moreno, 2005; Khare and Shaulsky, 2006; Morata and Martin, 2007; Johnston, 2009; Green, 2010; Baker, 2011; de Beco et al., 2012; Vivarelli et al., 2012). Like in the ecological context, competition for limited resources underlies the cellular selection regime (McLean et al., 1997; De Boer et al., 2001; Gaudin et al., 2004). Soluble, growth-inhibitory factors, possibly including miRNA (Kosaka et al., 2012), mediate the competitive interactions between stressed cells (Komarova et al., 1998; Yu X et al., 2006; Senoo-Matsuda and Johnston, 2007; Kosaka et al., 2012). Here, it is proposed that these secreted factors are the cellular arms of competitive wars that cause the bystander effects observed in multiple cellular stress models (Sowa Resat and Morgan, 2004; Yu X et al., 2006; Di et al., 2008; Asur et al., 2009; 2010; Ilnytskyy and Kovalchuk, 2011).

Caporale (2009) posited that “selection must act on the mechanisms that generate variation, much as it does on beaks and bones”. On the other hand, it is variation that gives selection the raw material to work on. Variation can be caused both by genetic, epigenetic or stochastic processes. Stochasticity in gene expression gives rise to cell-to-cell variability in protein concentrations and individual cells differ widely in responsiveness to uniform physiological stimuli (see Heininger, 2012). Cellular oxidative stress-dependent responses, although undoubtedly programmed, are also highly variable (Heininger, 2012), at least in part based on the stochasticity of mitochondrial bioenergetic/oxidative events (Hüser et al., 1998; Genova et al., 2003; Passos et al., 2007; Wang W et al., 2008). Cells employ a variety of quality surveillance and assurance systems including molecular chaperones (Esser et al., 2004; McClellan et al., 2005; Bukau et al., 2006; Buchberger et al., 2010; Arias and Cuervo, 2011), the ubiquitin/proteasome pathway (Sutovsky et al., 2001; 2002; Kostova and Wolf, 2003; Sutovsky, 2003; Thompson et al., 2003; Kwon et al., 2005; Taylor and Rutter, 2011), autophagy (Jin and White, 2007; Yorimitsu and Klionsky, 2007; Lee JY et al., 2010; Lee and Yao, 2010; Arias and Cuervo, 2011; Murrow and Debnath, 2013), mitochondrial turnover (Tatsuta and Langer, 2008; Twig et al., 2008; Dagda and Chu, 2009; Luce et al., 2010), the endoplasmic reticulum (Ellgaard and Helenius, 2003; Jørgensen et al., 2003; Kostova and Wolf, 2003; Kleizen and Braakman, 2004; Groenendyk and Michalak, 2005; Buchberger et al., 2010), and apoptosis (Yin et al., 1998; 2002; Meier et al., 2000; Groenendyk and Michalak, 2005; Jin and White, 2007; Igaki, 2009). Quality control implies that it selects for performance in cellular functions and eliminates inferior units. Cellular selection is the ultimate consequence when repair systems fail or are overwhelmed. Cellular selection within multicellular organisms does not only occur within the immune system and between cancer cells (Nowell, 1976; Kisielow and von Boehmer, 1995; McLean et al., 1997; Breivik and Gaudernack, 1999; von Boehmer et al., 2003; Vineis, 2003; Frank and Nowak, 2004; Merlo et al., 2006; Moreno, 2008; Vermeulen et al., 2008; Kim et al., 2011; Tamori and Deng, 2011; Thomas et al., 2013) but, as predicted by Roux (1881), is ubiquitous during development (Purves, 1980; Kupiec, 1986; 1996; 1997; Edelman, 1987; Michaelson, 1987; 1993; Otto and Orive, 1995; Møller and Pagel, 1998; Otto and Hastings, 1998; Deppmann et al., 2008; Tamori and Deng, 2011; de Beco et al., 2012). At least two major theories are based on selectionism, even if at different levels. The clonal selection theory that was elaborated by Jerne (1955) and Burnet (1957), and later confirmed by Tonegawa (1976; 1983), states that the diversity of the antibody repertoire in dedicated immune cells is achieved by random gene recombination events, leading to a huge number of small cellular lineages (Heams, 2012). Another major selectionist theory is the ‘selective stabilization of synapses’ (Changeux et al., 1973; Changeux and Danchin, 1976), later confirmed and even explicitly named ‘neural Darwinism’ (Edelman and Mountcastle, 1978; Edelman, 1987).

Male-biased mutagenesis requires that genetic variation created by this bias undergoes stronger selective scrutiny. Male-biased genes are evolving more rapidly than female-biased genes (Zhang Z et al., 2004a; Davis JC et al., 2005; Hambuch and Parsch, 2005; Eads et al., 2007), providing evidence that males experience stronger positive selection than females (Ranz et al., 2003; Connallon and Knowles, 2005; Eads et al., 2007; Ellegren and Parsch, 2007; Mallet and Chippindale, 2011).

An early and massive wave of apoptosis in the testis occurs among germinal cells during the first round of spermatogenesis (Russell, 1977; Huckins, 1978; Kerr, 1992; Blottner et al., 1995; Rodríguez I et al., 1997; Matsui, 1998; Kimura et al., 2003; Strbenc et al., 2003; Dadhich et al., 2010; Aitken et al., 2011). Massive apoptotic death occurs at multiple sites within the testes (Russell, 1977; Kerr, 1992; Baum et al., 2005; Kwon et al., 2005; Zheng et al., 2006; Codelia et al., 2008; Vergara et al., 2011). Mammalian germ cell apoptosis during normal spermatogenesis has been estimated to result in the loss of up to 75% of the potential numbers of mature sperm cells in the adult testis (Oakland, 1956; Huckins, 1978; De Rooij and Lok, 1987; Bartke, 1995; Dunkel et al., 1997). Recent evidence indicates that the extent of gamete apoptosis may even have been underestimated (Mitchell et al., 2011).

In mammals, oogenesis begins with the formation of primordial germ cells and encompasses a series of cellular differentiation events, from primordial germ cells to oogonia, from oogonia to oocytes, and from oocytes to eggs. Extensive degeneration of germ cells has been described during embryonic, fetal, and early postnatal stages of oogenesis before follicle formation. In the mouse embryo, early morphological studies have shown that cell death may occur in primordial germ cells or oogonia (12–13 days post coitum, dpc), but mainly in oocytes at the zygotene/pachytene stage of meiotic prophase I (MPI; from 16.5 dpc through birth; Bakken and McClanahan, 1978).

Following such episodes, the number of oocytes decreases from w20,000 at 13.5 dpc to about 6000–10,000 after 6 days, at birth (Burgoyne and Baker, 1981; Tam and Snow, 1981). More recently, McClellan et al. (2003) performed a careful study of the number of oocytes throughout MPI in the embryonic mouse ovaries and reached the conclusion that the decrease of oocyte population during this period (about 65% loss) is a continuous process without apparent peaks of degeneration.

A variety of theories were put forward to explain the “waste” (Tilly, 2001): death by neglect, death by defect and death by self-sacrifice. Before discussing these theories, I would like to recall a couple of fundamental evolutionary principles.

In a world of limited resources (Heininger, 2012) metabolic efficiency and economic utilization of resources are fitness traits under directional selection (Fitter, 1986; Zotin, 1990; Boggs, 1992; Parsons, 1997; 2005; 2007; Stelling et al., 2002; Hunt et al., 2004; Demetrius, 2005; MacLean, 2008; Frank, 2010; Heininger, 2012). For instance, if two amino acids at a given position on the protein can do the same job, then selection might favor the retention of the one for which synthesis requires less energy. The amino-acid compositions in the proteomes of Escherichia coli and Bacillus subtilis reflect the action of such selection pressure (Akashi and Gojobori, 2002). The amino acid sequences of highly abundant proteins in E. coli, Saccharomyces cerevisiae, and Schizosaccharomyces pombe have to compromise between optimization for their biological functions and reducing the consumption of limiting resources for their synthesis. By contrast, the amino acid sequences of weakly expressed proteins are more likely to be optimized for their biological functions (Li N et al., 2009). In a wide range of taxa, highly and broadly expressed genes, such as housekeeping genes, are shorter in both their intronic and coding sequences than genes expressed at low level or in a few tissues (Hurst et al., 1996; Castillo-Davis et al., 2002; Urrutia and Hurst, 2003; Eisenberg and Levanon, 2003; Vinogradov, 2004; De Ferrari and Aitken, 2006; Tu et al., 2006; Rao et al., 2010). Because transcription and translation are energetically costly, this shortness has been interpreted as a result of selection for economy (Castillo-Davis et al., 2002; Urrutia and Hurst, 2003; Eisenberg and Levanon, 2003; Seoighe et al., 2005; Rao et al., 2010). According to MacArthur and Wilson (1967, p. 149): “Evolution … favours efficiency of conversion of food into offspring”. In a similar ecological context, an energetic definition of fitness was put forward. According to the concept of Brown et al. (1993; 2004), reproductive power is composed of two component processes: acquisition (acquiring resources and storing them in reproductive biomass) and conversion (converting reproductive biomass into offspring) (Loreau, 1998; Allen C et al., 2006). In a world of limited resources, waste production should be strongly selected against. In fact, a significant proportion of the transcriptome that has been regarded as ‘junk’ and ‘transcriptional noise’ (Brosius, 2003; 2005; Volff, 2006; Kapranov et al., 2007; Ponjavic et al., 2007; Struhl, 2007) appears increasingly to be functional and selected for (Muotri et al., 2007; Ponjavic et al., 2007; Nordström et al., 2009; Ponting et al., 2009; Khachane and Harrison, 2010; Lebenthal and Unger, 2010; Managadze et al., 2011; Atkinson et al., 2012; Barry and Mattick, 2012; Hu et al., 2012; Lee, 2012; Rinn and Chang, 2012). Thus, it makes little evolutionary sense that organisms that exert such economy with regard to other energetically costly functions should senselessly engage in immense waste production related to their most fundamental biological function of reproduction. In a world of limited resources, such organisms should be easily outcompeted by organisms that exert a higher efficiency when converting resources into offspring (MacArthur and Wilson, 1967; Brown et al., 1993; 2004).

Outlining the traditional theories, about the rationale behind the gamete overproduction, I use the arguments brought forward by Tilly (2001).

Death by neglect. Death by neglect is used to conceptualize developmental cell death such as the high loss of neurons during the formation of the central nervous system (Giehl, 2001). The competition between cells for a limiting amount of growth (‘survival’) factors has been proposed as a key component of organogenesis (Jacobson et al., 1997). Cells that receive insufficient trophic support from their environment simply wither and die. Evidence for ‘death by neglect’ in the developing female germline comes mainly from studies of gametogenic failure in mutant female mice that lack germ-cell survival factors, such as stem-cell factor (Mintz and Russell, 1957) or interleukin-1α/β (Morita et al., 2001). Furthermore, in vitro culture of fetal mouse ovaries in the absence of serum or cytokines leads to a rapid induction of germ-cell apoptosis, and germline death in this artificial situation of acute neglect can be prevented by adding exogenous survival factors (Morita et al., 1999). In addition, fetal ovarian germ-cell death both in vivo and ex vivo is attenuated in female mice by deletion of either a key stress sensor (ceramide) (Morita et al., 2000) or a downstream executioner of apoptosis (caspase-2) (Bergeron et al., 1998). In fact, the increased numbers of oocytes that are lost from the ovaries of mice that are deficient in germ cell survival factors can be rescued from death by the inactivation of the caspase-2 gene (Morita et al., 2001). Likewise, withdrawal of hormonal support enhances apoptosis in male germ cells (Sinha Hikim et al., 1995; 1997; Tesarik et al., 1998; 2002; Woolveridge et al., 1999).

Death by defect. A second theory for why so many germ cells are lost during fetal ovarian development is that apoptosis eliminates oocytes with meiotic pairing or recombination anomalies. The idea is that a surveillance (or quality-control) mechanism detects and removes defective oocytes and retains meiotically competent oocytes for the formation of primordial follicles. Evidence for ‘death by defect’ comes mainly from studies of impaired germline development in mice with genetic mutations that cause abnormalities in either chromosomal recombination or pairing during meiosis. For example, inactivation of the ataxia telangiectasia-mutated (Atm) gene causes massive germ-cell apoptosis in both sexes at, or shortly after, prophase I of the first meiotic division. As a consequence, Atm-deficient female mice are born with ovaries that lack oocytes (Barlow et al., 1998). Various testicular injuries, all of them known to induce DNA damage, including heat (Lue et al., 1999) and toxicant exposure (Ku et al., 1995; Li LH et al., 1996), radiation (Meistrich, 1993), freezing and thawing (Tesarik et al., 2000) have been shown to enhance the apoptotic process as compared with the physiological condition. In the human testis, apoptosis appears to be the final result of various testicular and systemic pathologies (Gandini et al., 2000).

Interestingly, this model of death by defect is unaffected by the simultaneous inactivation of the gene that encodes either caspase-2 or its upstream activator Bax (Morita et al., 2001). This indicates that fetal oocyte apoptosis occurs by more than one pathway, and that death by defect might not be amenable to control by experimental manipulation or therapeutic intervention. Mutant mice with X-chromosome defects have also been used to support the death by defect hypothesis. Female mice that lack a second X chromosome (XO) or that harbour a large X-chromosome inversion (InX/X) show gametogenic failure (Burgoyne and Baker, 1985), presumably due to failed chromosome pairing, which leads to increased germline death during the development of the fetal ovaries. A similar situation occurs in Ullrich–Turner (XO) syndrome in humans (Zinn and Ross, 1998). Mutant mice with X-chromosome defects have also been used to support the death by defect hypothesis. Unfortunately, cytogenetic evidence from investigations of oocyte attrition in human fetal ovaries has not clearly shown whether meiotic defects are a cause or a consequence of germline apoptosis (Speed, 1988; Mittwoch and Mahadevaiah, 1992).

Therefore, the jury is still out on the contribution, if any, of death by defect to the overall number of oocytes that are eliminated by apoptosis before birth under normal physiological conditions.

It should be noted that both “Death by neglect” and “Death by defect” that are treated here as independent factors, in my theory constitute interdependent factors of the gamete quality control system by competition for trophic factors.

I will not dwell on the third theory discussed by Tilly (2001): “Death by self-sacrifice”. I have shown (Heininger, 2001; 2012) that “altruistic suicide” is evolutionary no man's land. How could a cell that “decides” voluntarily to commit suicide possibly transmit to its progeny the genes that program cell death? This evolutionary nonsense is still thoughtlessly reiterated in texbooks and scientific publications and certainly warrants another evolutionary blow (in preparation).

There is good evidence for both separate and synergistic roles for both testosterone and follicle-stimulating hormone (FSH) in achieving quantitatively normal spermatogenesis. Based on withdrawal and replacement studies, FSH has key roles in the progression of type A to B spermatogonia and, in synergy with testosterone, in regulating germ cell viability. Testosterone is an absolute requirement for spermatogenesis. In rats, it has been shown to promote the adhesion of round spermatids to Sertoli cells, without which they are sloughed from the epithelium and spermatid elongation fails (McLachlan et al., 2002). Withdrawal of trophic support results in a potentially stressful environment (Sinha Hikim and Swerdloff, 1999). Programmed cell death (apoptosis) is required for normal spermatogenesis in mammals (Knudson et al., 1995; Furuchi et al., 1996; Rodríguez I et al., 1997; Tres and Kierszenbaum, 1999). In some transgenic mice in which spermatogonial apoptosis was inhibited, there was an accumulation of spermatogonia and early spermatocytes that eventually all entered apoptosis (Knudson et al., 1995; Furuchi et al., 1996; Rodríguez I et al., 1997). Currently, it is thought that germ cell death serves to ensure cellular homeostasis and the fine balance between germ cells and Sertoli cells (Blanco-Rodríguez, 1998; de Rooij and Russell, 2000; Kierszenbaum, 2001). Huckins (1978) suggested that degeneration might be a mechanism to limit germ cells to that number which can be sustained by the Sertoli cell population, referred to as ‘‘density-dependent regulation’’ (de Rooij and Russell, 2000), which has become very popular (Kierszenbaum, 2001). In fact, the elongate spermatid/Sertoli cell ratio (a measure of the workload of the Sertoli cell and a prime factor determining their efficiency) in various mammalian species is approximately 10:1 (Russell and Peterson, 1984; Russell and Griswold, 1993), but significantly lower in primates (Johnson et al., 1984b; Russell and Peterson, 1984; Russell and Griswold, 1993; Johnson et al., 2008). Yet, meta-analyses demonstrated that the efficiency of spermatogenesis in several nonhuman primate species is comparable to that of rodents which are considered as species with highly efficient germ cell production. Comparative data from various primates revealed that Sertoli cell work load is species-specific but has no impact on germ cell numbers and on the efficiency of spermatogenesis (Luetjens et al., 2005). The relative inefficiency of human spermatogenesis has been well documented, with the human testis producing comparatively low numbers of sperm cells per unit weight of testis compared to various animal species (Brinkworth et al., 1997). However, the germ cell/Seroli cell ratio has been taken as given and the question for the adaptive role of this ratio, particularly the “understaffing” of Sertoli cells, has not been asked. Cell kinetic and radiobiological data indicate that Sertoli cells more resemble arrested proliferating cells than the classic postmitotic and terminally differentiated somatic cells that they have always been assumed to be (Ahmed et al., 2009; Tarulli et al., 2012). For instance, unlike terminally differentiated cells, Sertoli cells express a puzzling mixture of proliferation inducers and inhibitors (Ahmed et al., 2009). Sertoli cells proliferate in seasonal breeders, in which season-dependent variations in Sertoli cell numbers per testis occur (Hochereau-de Reviers and Lincoln, 1978; de Reviers et al., 1980; Johnson and Thompson, 1983; Johnson and Nguyen, 1986; Hötzel et al., 1998; Sinha Hikim et al., 1988; Johnson et al., 1991; Tarulli et al., 2006). The regulation of testicular development and proliferation of adult Sertoli cells is influenced by follicle-stimulating hormone (Orth et al., 1988; Singh and Handelsman, 1996; McLachlan et al., 2002; Johnston et al., 2004), testosterone (McLachlan et al., 2002; De Gendt et al., 2004; Johnston et al., 2004), estrogen (Fisher et al., 1998; Shetty et al., 1998; Pentikäinen et al., 2000; Kula et al., 2001; Oliveira et al., 2001) and thyroid hormone (Jannini et al., 1993; Van Haaster et al., 1993; Cooke et al., 1994; Holsberger and Cooke, 2005). Reducing estrogen synthesis in developing boars by inhibition of aromatase resulted in delayed lumen formation, lower testicular weight, fewer detergent-resistant spermatids, and fewer Sertoli cells, but by 7 to 8 months, these boars had recovered and had larger testes, more detergent-resistant spermatids per testis, and more Sertoli cells (At-Taras et al., 2006). It can be concluded that a delay in testicular maturation/puberty allows for a longer window for the proliferation of Sertoli cells and maturation of Leydig cells, leading to larger testes and higher spermatid production (Kula et al., 2001; At-Taras et al., 2006). In adult mice deficient for inducible nitric oxide synthase, testis weights were approx. 31% higher and testicular sperm count was 65% higher than in control animals (Lue et al., 2003). Associated with a reduced incidence of spontaneous germ cell apoptosis and unchanged rate of germ cell proliferation, mutant mice had a ~40% increase in the number of pachytene spermatocytes and ~34% in round spermatids, with no apparent changes in the number of preleptotene spermatocytes and spermatogonia (Lue et al., 2003). Neonatal hypothyroidism also lengthens the period of Sertoli cell proliferation, with consecutive increases in Sertoli cell number, testis weight, and daily sperm production when euthyroidism is re-established (Cooke et al., 1991; Cooke and Meisami, 1991; Hess et al., 1993; De França et al., 1995; Matta et al., 2002; Holsberger and Cooke, 2005). Meachem et al. (1996) extended the normal period of Sertoli cell proliferation, resulting in increased Sertoli cell numbers that persisted into adulthood; the number of Sertoli cells per testis was 118% and 149% greater in 90-day-old rats that had been treated with follicle-stimulating hormone during the first 10 or 15 days of life, respectively. In vertebrates, number of Sertoli cells and amount of smooth endoplasmic reticulum of Leydig cells (but not Leydig cell number) are related to relative testis mass and efficiency of spermatogenesis (De França et al., 1995; Johnson, 1995; Matta et al., 2002; Sharpe et al., 2003; Atanassova et al., 2005; Ford et al., 2006; Petersen and Söder, 2006). Testis mass and sperm output are increased in diverse taxa under the co-evolutive pressure of sperm competition risk (Harcourt et al., 1981; Billard, 1986; Møller 1988a; b; 1989; 1991; Jennions and Passmore, 1993; Gage, 1994; Møller and Briskie, 1995; Hosken, 1997; Stockley et al., 1997; Hosken and Ward, 2001; Hosken et al., 2001; Garamszegi et al., 2005). Thus, there are physiological processes that could augment Sertoli cell numbers but there seem to exist adaptive reasons that keep Sertoli cell numbers relatively low. A clue to these adaptive reasons is provided by the finding that, compared with normospermic counterparts, teratospermic cats have a higher sperm output achieved by more sperm-producing tissue, more germ cells per Sertoli cell, and reduced germ cell loss during spermatogenesis. However, gains in sperm quantity are produced at the expense of sperm quality (Neubauer et al., 2004). The germ cell quality signal, TNFalpha (Pentikäinen et al., 2001), mediates a negative feedback loop to Leydig cells that have been shown to respond to TNFalpha by decreasing their biosynthesis of testosterone (Li et al., 1995; Mauduit et al., 1998; Budnik et al., 1999). In rodent testis, the demise of spermatogonia (Allan et al., 1992), spermatocytes, and spermatids (Billig et al., 1995; Henriksen et al., 1995; Sinha Hikim et al., 1995) occurs through an apoptotic mechanism regulated by gonadotropins and androgens. Intriguingly, testosterone has been shown to both inhibit (Tapanainen et al., 1993; Henriksen et al., 1995; Erkkilä et al., 1997) and induce (Henriksen et al., 1995; Lue et al., 2006; Jia et al., 2007; Wang et al., 2007) germ cell apoptosis in a stage-specific manner.

Reproduction biology failed to outline an evolutionary rationale for the waste production of gametes. Taking into account the legendary “twofold cost of sex” (Maynard Smith, 1978a), the millionfold “costs of waste gamete production”, if having no adaptive function, should definitely deter any organism from investing into sexual reproduction. The fitness function w(x) describes how resources affect individual survival and reproductive success and its value is the expected number of offspring born to individuals with x units of resource (Rogers, 1992). According to MacArthur and Wilson (1967, p. 149): “Evolution … favours efficiency of conversion of food into offspring”. In this regard, sexual reproduction should be highly uneconomic. Yet ‘‘..unnecessary but costly structures or activities should be lost in evolution.’’ (Michod, 1999a). Organisms have to bear heavy costs of reproduction (Dewsbury, 1982; Nakatsuru and Kramer, 1982; Heininger, 2012) that are paid in terms of reduced growth, immunocompetence, stress resistance and survival (Richner et al., 1995; Møller et al., 1998; 1999; Siva-Jothy et al., 1998; Brown et al., 2000; Rigby and Jokela, 2000; Hosken, 2001; Heininger, 2012). Organisms that would have been able to economize these costs, e.g. by minimizing the huge“waste” gamete production, would have a substantial competitive advantage. An adaptive explanation of these counterintuitive features is urgently warranted. It is often argued that further improvements in replication fidelity and DNA repair become too costly (André and Godelle, 2006), reflecting the combined metabolic and temporal costs of perfection in replication and transcription fidelity (Kimura, 1967; Sniegowski et al., 2000). But wouldn’t it be much more resource-efficient and resource-saving to invest into the perfection of replication and transcription fidelity than to produce millions and billions of “waste/poor-quality” gametes? Particularly, since the huge energetic investment into this “waste” production results in gonadal functional hypoxia (at least in mammals), and is associated with increased metabolic/oxidative stress and genetic instability that further deteriorates gamete quality. In fact, improving genome replication fidelity is feasible in General Purpose Genotypes, as has been shown e.g. in asexual Darwinula stevensoni (Rossi et al., 1998; 2004; Schön et al., 1998; 2000, 2003; 2009; Gandolfi et al., 2001) and an invasive grass weed (Le Roux et al., 2007). Although evolutionarily highly successful, these organisms rather appear to be evolutionary dead-ends. That evolution did not reward this economized, but non-innovative, conversion of reproductive biomass into offspring (Brown et al., 1993; 2004) but rather favored the increase of this “waste” investment along the phylogenetic axis from invertebrates to mammals clearly emphasizes the adaptive role of this gamete overproduction. In fact, theoretical models show that mate and gamete selection are more efficient in the use of biomass, energy and time, than natural selection at the level of organisms, helping to make sexual reproduction an evolutionary success (Jaffe, 2004).

The disposable soma theory (DST) predicts that aging occurs due to the accumulation of damage during life and that failures of defensive or repair mechanisms contribute to aging (Kirkwood, 1977; Kirkwood and Holliday, 1979; Kirkwood and Austad, 2000). It postulated a special class of gene mutations with antagonistic pleiotropic effects in which hypothetical mutations save energy for reproduction (positive effect) by partially disabling molecular proofreading and other accuracy promoting devices in somatic cells (negative effect). In other words, given finite resources, the more resources an animal spends on bodily maintenance, the less it can expend on reproduction, and vice versa. This theory has a variety of implausibilities (see Heininger, 2012). According to the DST, any resources not used for reproduction should benefit the somatic maintenance and repair and delay aging. Therefore waste of resources during reproduction should be selected against.

Importantly, the adaptive explanation integrates both the arguments of the death by neglect and death by defect theories into a coherent concept. I will present compelling evidence that there are several levels of selection that unfold during sexual reproduction. I would have preferred to call these levels summarily as “sexual selection”. However, a narrow use of the term “sexual selection” in the sense of pre-mating and postmating selection is currently deeply entrenched in the scientific literature. The future will show whether a broad use of the term can prevail. But for the time being, I will use the term “sexual selection” in its narrow sense and use the term “sexual mutagenesis-selection-cascade” (SMSC) for the entire sexual reproduction-related bet-hedging strategy.

8.1.1 Cell competition and Myc

Myc family transcription factors are phylogenetically conserved and arose before the divergence of the choanoflagellate and metazoan lineages (Walker et al., 1992; Atchley and Fitch, 1995; Gallant, 2006; Hartl et al., 2010; Young et al., 2011). The Myc target gene network is estimated to comprise about 15% of all genes from flies to humans. Both genomic and functional analyses of c-Myc targets suggest that while c-Myc behaves as a global regulator of transcription, groups of genes involved in cell cycle regulation, metabolism, ribosome biogenesis, protein synthesis, and mitochondrial function are over-represented in the c-Myc target gene network (Dang et al., 2006). Thus expression of Myc family transcription factors is closely tied to cell growth and proliferation as well as inhibition of terminal differentiation and induction of apoptosis (Grandori et al., 2000; Zhou and Hurlin, 2001; Pelengaris et al., 2002; de la Cova and Johnston, 2006). c-Myc also affects the stability of the whole genome and triggers the initiation of a complex network of genomic instability via the induction of ROS (Kuttler and Mai, 2006; Prochownik and Li, 2007). On the other hand, Myc facilitates replication under stress (Herold et al., 2009) and regulates both proliferation and apoptosis (Evan et al., 1994; Secombe et al., 2004; Montero et al., 2008). MYC is a potent oncogene that can promote tumorigenesis in a wide range of tissues (Verbeek et al., 1991; Felsher and Bishop, 1999; Pelengaris et al., 1999; Jain et al., 2002; Flores et al., 2004; Soucek et al., 2008; Sodir et al., 2011). The ability of Myc-overexpressing cells to abrogate apoptosis and maintain proliferation in a cell autonomous manner is an important step in tumor progression (Evan and Vousden, 2001). MYC is the most frequently amplified oncogene and the elevated expression of its gene product, the transcription factor c-Myc, correlates with tumor aggression and poor clinical outcome (Nesbit et al., 1999; Beroukhim et al., 2010; Lin et al., 2012). Elevated expression of c-Myc occurs through multiple mechanisms in tumor cells, including gene amplification, chromosomal translocation, single nucleotide polymorphism in regulatory regions, mutation of upstream signaling pathways, and mutations that enhance the stability of the protein (Eilers and Eisenman, 2008; Meyer and Penn, 2008; Pomerantz et al., 2009; Wright et al., 2010). Rather than binding and regulating a new set of genes, c-Myc amplifies the output of the existing gene expression program (Lin et al., 2012; Littlewood et al., 2012; Nie et al., 2012). The relative level of Myc between cells is a marker of competitiveness. Cells that express more Myc become ‘‘supercompetitors’’ that outcompete even wild-type cells (Abrams, 2002; de la Cova et al., 2004; Moreno and Basler, 2004; Secombe et al., 2004; Rhiner et al., 2009). Loss of completely normal, wild-type cells in the presence of triplo-myc or tetraplo-myc ‘supercompetitor’ cells reinforces the conclusion that competition does not reflect any intrinsic defect in the outcompeted population but is a response to relative competitiveness (Baker, 2011). In addition to MYC, cell competition is regulated by a variety of signaling pathways (Moreno et al., 2002; Bondar and Medzhitov, 2010; Marusyk et al., 2010; Rhiner et al., 2010; Tamori et al., 2010; Baker, 2011; Vincent et al., 2011; Chen CL et al., 2012; Norman et al., 2012; Rodrigues et al., 2012; Verghese et al., 2012; Vivarelli et al., 2012), including p53 (Bondar and Medzhitov, 2010; Green, 2010; de Beco et al., 2012).

Secreted protein acidic, rich in cysteine (SPARC)/osteonectin is a secreted multifunctional glycoprotein and belongs to the family of matricellular proteins, which modulate cell-cell and cell-matrix interactions and are induced during morphogenesis, development, tissue injury, and tissue remodeling (Lane and Sage, 1994; Clark and Sage, 2008). Mammalian SPARC is known to bind several extracellular matrix (ECM) proteins and modulate the activity of various growth factors/chemokines (TGF-beta, VEGF, etc.). It has been mainly found to function in de-adhesion, anti-proliferation and regulation of ECM production (Framson and Sage, 2004). In Drosophila development, SPARC is transcriptionally upregulated in loser cells during cell competition that provides a transient self-protective signal by inhibiting caspase activation in outcompeted cells (Portela et al., 2010). As a self-protecting signal, SPARC has also been implicated in cancerogenesis (Brekken et al., 2003; Framson and Sage, 2004; Clark and Sage, 2008; Arnold and Brekken, 2009; Petrova et al., 2011). The expression of SPARC in human tumors is also consistent with its role during cell competition (Petrova et al., 2011). SPARC is expressed in mouse spermatogonia, pachytene spermatocytes and less in round spermatids (Pang et al., 2006). During fetal testis development, SPARC is internalized in Sertoli, Leydig, and germ cells suggesting an intracellular regulatory role in these cell types (Wilson MJ et al., 2006). Non-transformed cells are anchorage dependent for the execution of the mitotic program (Chiarugi and Fiaschi, 2007). Hence, survival and development of germ cells depends on their continuous and close contact to Sertoli cells (McLachlan et al., 2002; Schulz et al., 2010). In this context, SPARC-mediated de-adhesion (Murphy-Ullrich et al., 1995; Greenwood and Murphy-Ullrich, 1998; Framson and Sage, 2004; Sweetwyne et al., 2004; Liu A et al., 2009) may be a mechanism mediating the fate of germ cells during cell competition.

8.1.2 Germ cell competition

Germ cell selection occurs when cells that differ genetically (because of mutation, crossing over, or gene conversion) also differ in their propensity to proliferate or survive during development (Hastings, 1989; Otto and Hastings, 1998). Although, the possibility of germline selection as a mechanism contributing to natural selection and organismal evolution was raised as early as in the 1930s (Shapiro, 1936; Haldane, 1937), selection among cells during the development of an individual has played only a cameo role in population genetics theory (Otto and Hastings, 1998). Population genetic calculations have predicted that such selection, if existent (it was held that conflict between cells are a potent threat to the integrity of multicellular individuals and needed to be reduced), would have significant effects on the frequencies and types of mutations and alleles in a population (Hastings, 1989; Otto and Orive, 1995; Otto and Hastings, 1998). Because selection within the individual acts as a sieve eliminating deleterious mutations and increasing the frequency of beneficial ones, mutations observed among progeny are pre-selected, hindering the spread of deleterious mutations, and reducing the genetic load imposed on the population by several orders of magnitude (Otto and Orive, 1995; Otto and Hastings, 1998). These effects increase with the number of cells within the germline and with the number of cell divisions (Otto and Hastings, 1998).

Mutations at the dominant white-spotting (W) locus of the mouse are one of the heritable disorders causing sterility and have pleiotropic effects on growth and differentiation of germ cells, erythrocytes, melanocytes, and mast cells. The W locus has been determined to be allelic with the c-kit proto-oncogene (Chabot et al., 1988; Geissler et al., 1988; Tan et al., 1990). The c-kit gene plays a fundamental role during the establishment, the maintenance and the function of germ cells. In the embryonal gonad the c-kit tyrosine kinase receptor and its ligand Stem Cell Factor (SCF) are required for the migration, survival and proliferation of primordial germ cells. In the postnatal animal, c-kit/SCF are required for the production of the mature gametes in response to gonadotropic hormones. c-kit is required for the survival and/or proliferation of the differentiating type A spermatogonia and for the growth and maturation of oocytes, but the primitive (undifferentiated) type A spermatogonia or spermatogenic stem cells are independent from c-kit (Yoshinaga et al., 1991; Manova et al., 1993; Mauduit et al., 1999; Sette et al., 2000). Mice homozygous or double-dominant heterozygous for some, but not all, W mutant alleles, such as W/Wv, Wv/Wv, and W44/W44 mice, are known to have postnatal viability but to be impaired in their fertility, an impairment almost always due to lack of germ cells in the gonads (Coulombre and Russell, 1954; Geissler et al., 1981). In homozygotes for mutant alleles, the germ cells fail to increase after 9 days of gestation, a consequence of a deficiency in proliferative capacity of the PGC during migration from the hindgut region to the genital ridges (Mintz, 1957; Mintz and Russell, 1957). Heterozygotes for several severely impaired alleles, such as W35, W38, W40, W42, and W43 have smaller testes; and W39/W39 mice, which are viable and have limited fertility, showed reduced spermatogenic activity (Geissler et al., 1981). The Wsh and Wf mutant alleles have rather weak effect on proliferation and/or differentiation of germ cells (Guenet et al., 1979; Lyon et al., 1984); the homozygotes and the heterozygotes between them are fertile. However, regenerative differentiation after surgical reversal of cryptorchidism was impaired in W mutants, suggesting a functional significance of the W (c-kit) gene in postnatal gametogenesis in males (Koshimizu et al., 1991). Nakayama et al. (1990) used chimeras between W embryos and +/+ embryos to investigate germ cell competition. Cleavage-stage embryos from the mating between Wsh/Wf males and Wsh/Wf females were aggregated with +/+ embryos to produce chimeras. Three male (XYXY) and three female (XXXX) chimeras were mated with +/+ partners. Two male chimeras made only +/+ progenies. In the remaining one male, the proportion of +/+ progeny was larger than that of Wf/+ progeny, whereas the contribution of +/+ cells was considerably smaller in the fibroblast population. In contrast, such an apparent predominance of +/+ progenies was not observed in the three female chimeras examined (Nakayama et al., 1990). The results suggest a selective survival of +/+ germ cells in the male but not the female gonads arguing for a more stringent germ cell competition and quality control in males. The authors speculated that using W mutants with more severely impaired germ cell proliferative potential, may have allowed to demonstrate germ cell competition and selective progeny generation also in females (Nakayama et al., 1990).

In the Drosophila ovary, germ stem cells have a competitive relationship for niche occupancy, which may serve as a quality control mechanism to ensure that accidentally differentiated stem cells are rapidly removed from the niche and replaced by functional ones (Jin et al., 2008). dmyc, the homolog of the human c-myc oncogene, is known to be required for endoreplication of the Drosophila differentiating cysts (Maines et al., 2004) and its higher expression gives a competitive advantage to germline stem cells (Rhiner et al., 2009). c-myc expression is modulated by growth regulators and is correlated with the establishment and maintenance of the 'growing state' in somatic cells (Kelly et al., 1983; Armelin et al., 1984; Campisi et al., 1984). Intriguingly, however, in non-dividing Xenopus oocytes c-myc mRNA is present at a steady-state level 104 fold the myc content of proliferative somatic cells. This very high level of c-myc transcript was reached early in oogenesis and remained constant in cell cycle-arrested vitellogenic oocytes, suggesting a posttranscriptional control and that the function of the c-myc gene in oocytes may not be implicated directly in sustaining DNA synthesis or mitosis (Godeau et al., 1986; Taylor et al., 1986; Méchali et al., 1988). C-MYC is also expressed in growing and fully grown mouse oocytes (Suzuki T et al., 2009). Yet, it appears that in oogenesis of a variety of higher vertebrate taxa p53, as a sensor of cellular fitness (de Beco et al., 2012), rather than Myc took over the role of cell competition censor (Ghafari et al., 2009; Belyi et al., 2010; Levine et al., 2011). A possible reason for this switch in higher taxa may be that p53 is a tumor suppressor while MYC is an oncogene that is routinely downregulated in differentiating cells (Reitsma et al., 1983; Campisi et al., 1984; Dotto et al., 1986; Resnitzky et al., 1986) as a barrier to oncogenesis. Although physiological HIF1 responses can inhibit the activity of normal MYC, paradoxically, the deregulated expression of oncogenic MYC collaborates with HIF to confer the tumor metabolic phenotype that is described as the Warburg effect, or aerobic glycolysis (Dang, 2007; Dang et al., 2008).

In response to a Sertoli cell-derived growth factor, glial cell line derived neurotrophic factor, mouse spermatogonial stem cells proliferate and up-regulate N-myc expression (Braydich-Stolle et al., 2007). In various invertebrate and vertebrate seasonal breeders, c-myc is seasonally expressed in testicular tissue indicating peak expression during active spermatogenesis (Walker et al., 1992; Chieffi et al., 1995; 1997). c-Myc is specifically highly expressed in rodent spermatogonia and spermatocytes (Koji et al., 1988; Wolfes et al., 1989; Uetani et al., 1994; Teng and Vilagrasa, 1998) and in human sperm cells (Kumar et al., 1993). In erythroleukemia cells, in primary keratinocytes, in teratocarcinoma, in myeloid M cells, and in human promyelocytic cells, c-Myc protein down-regulation is required to reach the state of differentiation (Reitsma et al., 1983; Campisi et al., 1984; Dotto et al., 1986; Resnitzky et al., 1986). Likewise, a stage of c-Myc gene down-regulation during meiosis has been hypothesized to be necessary for triggering spermatogenic cells to differentiate (Stewart et al., 1984). Spermatocyte apoptosis induced by gossypol, an uncoupling agent to oxidative phosphorylation, is correlated with biphasic c-Myc protein expression (Teng and Vilagrasa, 1998). Within 0.5 to 2 h of the addition of gossypol to spermatocytes, levels of c-Myc proteins fall dramatically and remain at a low level for the next several hours. Between 3 and 5 h after exposure to gossypol, the c-Myc protein content returns to preexposure (or higher) levels, another 1.5–4 h before the apoptotic death of germ cells (Teng and Vilagrasa, 1998). The fall of c-Myc protein levels is consistent with the loss of competitive potential of metabolically stressed spermatocytes. Likewise, a rapid and constant induction of HL-60 cell apoptosis depends on a combination of down- and up-regulation of the c-Myc gene (Kimura et al., 1995). In mice (Suzuki et al., 1996) and rats (Kodaira et al., 1996), overexpression of a c-myc transgene in testis caused arrest of spermatogenesis during meiosis, massive apoptosis of primary spermatocytes and subsequent sterility.

8.2 Germ cell selection

As pointed out by Hastings (1991), individual genes in sexual organisms pass through numerous asexual mitotic cell divisions in the germline prior to meiosis and sexual recombination. The processes of mitotic recombination, mitotic crossing over, and mitotic gene conversion may create genotypic diversity between diploid cells in the germline (John and Miklos, 1988; Lankenau, 2007) and have been demonstrated in yeast (Malone and Esposito, 1980; Rudin and Haber, 1988; Lichten and Haber, 1989; Yuan and Keil, 1990; McGill et al., 1993), plants (Puchta et al., 1993), flies (Kennison and Ripoll, 1981; Gethmann, 1988; Lankenau, 2007) and mammals (Panthier et al., 1990; Rouet et al., 1994; Choulika et al., 1995; Sargent et al., 1997; Liang et al., 1998; Johnson and Jasin, 2001; Helleday, 2003; Moynahan and Jasin, 2010). The rate of mitotic crossing-over has been estimated to be within the range of 10-4 to 10-2 per individual generation in Drosophila (Gethmann, 1988), 10-5 to 10-4 per individual generation in plants (Evans and Paddock, 1979), and 10-7 to 10-5 per cell generation in yeast (Lichten and Haber, 1989; Yuan and Keil, 1990). Recombination repair is often referred to as an error-free repair pathway and advantageous to the cell. However, recombination is a double-edged sword and may also result in loss of heterozygocity (Sengstag, 1994; Tischfield, 1997). Particularly, copying of a donor sequence associated with gene conversion may be mutagenic (Pâques et al., 1998; 2001; Hicks et al., 2010), due to inefficient mismatch repair during gene conversion (McGill et al., 1998; Hicks et al., 2010).

It has been shown in vivo that at least two-thirds of rodent type A spermatogonia undergo programmed cell death (Oakberg, 1956; Clermont, 1962; Huckins and Oakberg, 1978; Allan et al., 1992; Dym, 1994). The extrinsic pathway of apoptosis that is characterized by the oligomerization of death receptors such as FAS or tumor necrosis factor followed by the activation of caspase-8 and caspase-3, plays an important role in germ cell apoptosis throughout the first wave of spermatogenesis in the rat under physiological conditions (Jahnukainen et al., 2004; Moreno et al., 2006; Zheng et al., 2006; Lizama et al., 2007; Codelia et al., 2008; Tripathi et al., 2009; Vergara et al., 2011). Particularly, spermatogonia and midpachytene spermatocytes are eliminated (Krishnamurthy et al., 1998; Oldereid et al., 2001; Weinbauer et al., 2001; Francavilla et al., 2002; Jahnukainen et al., 2004; Royere et al., 2004). Apoptosis both with and without caspase activation invariably requires oxidative stress (Smith R et al., 2006; Lysiak et al., 2007; Kalia and Bansal, 2009; Maheshwari et al., 2009). Intriguingly, spermatogonia have been thought to have a high tolerance to oxidative stress that is due to high levels of Zn and Cu/Zn superoxide dismutase (Celino et al., 2011). However, high levels of SODs, if unbalanced by enzymes that can metabolize H2O2, may have detrimental effects on cellular redox homeostasis as discussed in chapter 7.3.

Primitive type A (collected from 6-day-old mice) and type A spermatogonia (collected from 8-day-old mice) have the highest spontaneous mutant frequencies, followed by type B spermatogonia, preleptotene spermatocytes, leptotene and zygotene spermatocytes, pachytene spermatocytes, round spermatids, and epididymal spermatozoa, all of which display a similarly low mutant frequency (Walter et al., 1998). Ionizing radiation (IR) induces an increased mutant frequency, the amount of which declines as cells progress through spermatogenesis (Xu et al., 2008). Shapiro (1936) and Abrahamson et al. (1966) showed that premeiotic male germ cells undergo selection against X-linked lethals. These studies revealed mechanisms that exist to eliminate premeiotic cells with elevated mutation frequencies during spermatogenesis (Walter et al., 1998; Xu et al., 2008; 2010). In fact, apoptosis functions in male germ cells to mediate a decline in mutant frequency during spermatogenesis by removing cells with a high mutant frequency (Walter et al., 1998; Xu et al., 2008; 2010). Thus, apoptosis occurs extensively in the first wave of spermatogenesis in rodents and is critical as quality control tool eliminating less competitive germ cells (see chapter 8.1).

Genes expressed in the germline whose products affect cell viability (such as many “housekeeping” enzymes) may be subjected to selection acting on their variability resulting in a non-Mendelian output of gametes (Massicotte et al., 2006). Such genes will be governed by the population genetics of the sexual/asexual life cycle rather than the conventional sexual/Mendelian life cycle (Hastings, 1991). Importanly, many housekeeping genes are in general highly expressed (Vinogradov, 2004; Zhu et al., 2008) and are involved in cellular metabolism, energy, stress responses, cell adhesion, signal transduction, cytoplasmic and nuclear functions, such as protein and intracellular transport (Dix, 1997; Lequarré et al., 1997; Rose-Hellekant et al., 1998; Krisher and Bavister, 1999; Khurana and Niemann, 2000; Neuer et al., 2000; Aguilar-Mahecha et al., 2001; Rockett et al., 2001; Dalbiès-Tran and Mermillod, 2003; Muramoto et al., 2003; Yu et al., 2003; International Chicken Genome Sequencing Consortium, 2004; Krisher, 2004; Urner and Sakkas, 2005; Lowe et al., 2007), arguing for a role of these genes, and the cellular processes regulated by them, in the germline cell selection process. On the basis of publicly available expressed sequence tag data, a large fraction (40%) of currently-annotated human genes are universally expressed (Zhu et al., 2008). There is a hierarchy of purifying selection that is reflected by the pattern of molecular evolution. Comparisons of pufferfish (Takifugu rubripes), chicken and human genomes reveal around 7,000 genes that have 1:1 orthologs in all three species, suggesting a ‘core’ of genes that may have an essential role in all vertebrates (International Chicken Genome Sequencing Consortium, 2004; Furlong, 2005). The sequences in this core tend to be more conserved than other orthologs, indicating that strong purifying selection is acting upon them, stressing their functional importance.

When cells were subjected to hypoxia, HIF-1 preferentially bound to loci that were already transcriptionally active under normal growth conditions characterized by the presence of histone H3 lysine 4 methylation, the presence of RNA polymerase II, and basal production of mRNA. Cell type-specific differences in HIF-1 binding were largely attributable to differences in the basal gene expression patterns in the cells (Xia and Kung, 2009). Thus, the expression level is one of the major determinants of protein evolution (Sharp, 1991; Green et al., 1993; Duret and Mouchiroud, 2000; Pál et al., 2001; 2006; Krylov et al. 2003; Rocha and Danchin, 2004; Zhang and Li, 2004; Agrafioti et al., 2005; Drummond et al., 2005; 2006; Koonin and Wolf, 2006; Rocha, 2006; Drummond and Wilke, 2008; Wolf et al., 2008). Conserved patterns of simple covariation between sequence evolution, codon usage, and mRNA level in E. coli, yeast,worm, fly, mouse, and human suggest that selection against toxicity of misfolded proteins generated by ribosome errors suffices to create all of the observed covariation (Drummond and Wilke, 2008). Studies on the yeast Saccharomyces cerevisiae indicate that the strongest predictor of evolutionary rate is expression level of a protein that explains 30–50% of the variation in the rate of protein evolution (Pál et al., 2001; Drummond et al., 2005; 2006), much more than any other known variable. Furthermore, many amino-acid changes seem to be due to positive selection, often reflecting arms races or compensatory mutations rather than adaptation to changed environments (Pál et al., 2006). Broadly expressed proteins in mammals (Duret and Mouchiroud, 2000; Subramanian and Kumar, 2004), insects (Subramanian and Kumar, 2004) and plants (Wright et al., 2004), that can be expected to be expressed in germ cells evolve more slowly than tissue-specific proteins that can be expected to be not expressed in germ cells (Duret and Mouchiroud, 2000; Subramanian and Kumar, 2004; Zhang and Li, 2004; Liao and Zhang, 2006). Proteins functioning during different stages of development may be predisposed to having mutations of different selective effects. Thus, proteins expressed early in development and particularly during mid–late embryonic development evolve unusually slowly (Davis JC et al., 2005). Housekeeping genes evolve more slowly in protein sequence than tissue-specific genes (Hughes and Hughes, 1995; Hastings, 1996; Duret and Mouchiroud, 2000; Hirsh and Fraser, 2001; Jordan et al., 2002; Zhang and Li, 2004). Similarly, protein dispensability affects evolutionary rate (Hurst and Smith, 1999; Hirsh and Fraser, 2001; Jordan et al., 2002; Wall et al., 2005; Zhang and He, 2005; Liao et al., 2006; Larracuente et al., 2008). Genes expressed early in spermatogenesis had rates of divergence similar to the genome median, while genes expressed after the onset of meiosis were found to evolve much more quickly. Rates of protein evolution were fastest for genes expressed during the dramatic morphogenesis of round spermatids into spermatozoa. Late-expressed genes were also more likely to be specific to the male germline (Good and Nachman, 2005). Tissue expression tends to evolve rapidly for genes that are expressed in only a limited number of tissues, whereas tissue expression can be conserved for a long time for genes expressed in a large number of tissues (Duret and Mouchiroud, 2000; Zhang and Li, 2004; Yang J et al., 2005; Liao et al., 2006; Parmley et al., 2007; Park and Choi, 2010). A similar relationship can be seen when exons of a given gene are considered: alternatively spliced exons (which are generally also tissue specific) evolve at higher rates than constitutively spliced exons (Xing and Lee, 2005). However, there appears to be no reduction of mutation rates in genes expressed specifically in the germline (Duret and Mouchiroud, 2000). Expression pattern is an essential factor in determining the selective pressure on functional sites in both coding and noncoding regions. Conversely, silent substitution rates do not vary with expression pattern, even in ubiquitously expressed genes. This latter result thus suggests that synonymous codon usage is not constrained by selection in mammals (Duret and Mouchiroud, 2000). In addition to this high gene expression, high-evolutionary-rate, hotspots, another class of mutagenesis hotspots was identified in the genome characterized by reduced repair in hypermethylated lowly-transcribed genes (Zhao and Epstein, 2008; Tang and Epstein, 2010). These  processes result in regional mutation-rate variation within genomes (Wolfe et al., 1989; Ellegren et al., 2003).

To compensate for the stochastic nature of the (epi)mutation and recombination process, sexual reproduction relies on sperm and oocyte quality control. The cytokine TNFalpha appears to be a paracrine factor that signals germ cell quality (Pentikäinen et al., 2001). TNFalpha is known to be secreted by testicular germ cells (De et al., 1993). TNFalpha effectively and dose-dependently plays an anti-apoptotic role via the activation of NF-kappaB (Barkett and Gilmore, 1999; Pentikäinen et al., 2001; Suominen et al., 2004). Various studies have shown that NF-kappaB has both anti- and pro-apoptotic effects within cells (Kaltschmidt et al., 2000; Pentikäinen et al., 2002; Mathur et al., 2011). The expression of the TNF receptor protein in the seminiferous epithelium was predominantly found in Sertoli cells (De et al., 1993; Mauduit et al., 1996; 1998; De Cesaris et al., 1999; Pentikäinen et al., 2001). In Sertoli cells, TNFalpha regulates the production of germ cell trophic factors, e.g. lactate (Nehar et al., 1997; Boussouar et al., 1999; Erkkilä et al., 2002), transferrin (Sigillo et al., 1999; Lécureuil et al., 2004), glutathione (Meroni et al., 2000; Rahman, 2000; Yang H et al., 2005; Reuter et al., 2009), and IGF-binding protein (Besset et al., 1996). TNF-alpha stimulated NF-kappaB binding to the androgen receptor promoter, induced androgen receptor promoter activity, and increased endogenous androgen receptor expression in primary cultures of Sertoli cells (Delfino et al., 2003). In addition, cultured Leydig cells have been shown to respond to TNFalpha by decreasing their biosynthesis of testosterone (Li et al., 1995; Mauduit et al., 1998; Budnik et al., 1999), possibly a negative feedback loop that tightens the supply of the trophic factor, securing the maintenance of the quality assurance system.

Studies of mouse, rat, and human testicular apoptosis have shown that the Fas-Fas ligand system (Nagata and Golstein, 1995; Nagata, 1997) is a powerful mediator of male germ cell death (Lee J et al., 1997; 1999b; Boekelheide et al., 1998; Pentikäinen et al., 1999; Yin et al., 2002) and possibly involved in the quality control mechanism of the produced gametes (Braun, 1998; Ross et al., 1998; Odorisio et al., 1998; Francavilla et al., 2000; Neubauer et al., 2004). The Fas system is a mechanism through which cells expressing Fas ligand (FasL) induce apoptosis of Fas expressing cells. The testis is the only organ that constitutively expresses abundant amounts of FasL mRNA (D’Alessio et al., 2001). p53 has been reported to target the Fas gene for transcription (Owen-Schaub et al, 1995; Schilling et al., 2009), and up-regulation of Fas appears to be p53-dependent (Müller et al., 1997; Yin et al., 2002). FasL and Fas, expressed by Sertoli cells and germ cells, respectively, respond to environmental conditions and initiate germ cell death. In the rat seminiferous tubules, Sertoli cells express FasL (Suda et al., 1993; French et al., 1996; Richburg et al., 2000; D'Abrizio et al., 2004). FasL associates with the Fas receptor, which is a type I transmembrane receptor, located on testicular germ cells (Lee J et al., 1997; 1999b; Boekelheide et al., 1998; Pentikäinen et al., 1999). Up-regulation of Fas is a common and critical step for initiating germ cell death in vivo. Moreover, if Sertoli cells are injured, they up-regulate FasL to eliminate Fas-positive germ cells (Lee et al., 1997; 1999). In cultured mouse Sertoli cells, TNFalpha regulates the expression and function of the Fas system, suggesting a role for TNF in testicular apoptosis (Barkett and Gilmore, 1999; Riccioli et al., 2000). Germ cell-derived TNFalpha appears to down-regulate the level of the Fas ligand and thereby reduce physiological germ cell apoptosis (Pentikäinen et al., 2001; Suominen et al., 2004). Another study reported that FasL mRNA is strongly expressed in differentiating germ cells and that the protein is thereafter displayed on the surface only when the gamete is fully mature and, as spermatozoon, leaves the testis (D’Alessio et al., 2001). The authors proposed that FasL protein displayed on the surface of mature spermatozoa may represent a self-defense mechanism against lymphocytes present in the female genital tract (Riccioli et al., 2003).

Spermatogonia bearing relatively higher mutation frequencies are more likely to be included among the apoptotic cells than are those with low mutation frequencies (Walter et al., 1998). Cell competition (see chapter 8.1) that selects for the least damaged cells is controlled by p53 (Bondar and Medzhitov, 2010; Green, 2010; Marusyk et al., 2010), possibly via p53 C-terminal regulatory domain signaling (Wang YV et al., 2011). In invertebrates and vertebrates, p53 family proteins are activated by oxidative stress (Jayaraman et al., 1997; Xie et al., 2001; Martindale and Holbrook, 2002; Chen K et al., 2003; Vousden and Lane, 2007; Kotinas et al., 2012; Gambino et al., 2013), monitor the genomic quality of germ cells, protect them against stress and, in an ambivalent role, eliminate inferior ones (Norimura et al., 1996; Ollmann et al., 2000; Derry et al., 2001; Suh et al., 2006; Yamada et al., 2008; Gonfloni et al., 2009; Hu, 2009; Belyi et al., 2010; Levine et al., 2011). With regard to its protective role, p53 appears to play a central role in repressing ROS-induced stresses by upregulating antioxidant genes (Tan et al., 1999; Budanov et al., 2004; Sablina et al., 2005; Bensaad et al., 2006; Matoba et al., 2006; Tomko et al., 2006; Meiller et al., 2007; Cano et al., 2009; Chen et al., 2009; Olovnikov et al., 2009; Hu W et al., 2010; Pallepati and Averill-Bates, 2010; Popowich et al., 2010; Budanov, 2011; Nam and Sabapathy, 2011; Nii et al., 2012; Vurusaner et al., 2012; Kang MY et al., 2013). p53 couples energy metabolism and ROS formation by modulating the transcription of target genes that control the fluxes through mitochondrial respiration, glycolysis, or the pentose phosphate shunt (Liu B et al., 2008). On the other hand, high p53 expression appears to be a marker of “losers” in cell competition (Bondar and Medzhitov, 2010; Marusyk et al., 2010) that increases ROS production and activates apoptosis pathways (Johnson et al., 1996; Polyak et al., 1997; Donald et al., 2001; Liu and Chen, 2002; Martindale and Holbrook, 2002; Macip et al., 2003; Hussain et al., 2004; Sablina et al., 2005; Valko et al., 2007; Mai et al., 2010; Pani and Galeotti, 2011; Vurusaner et al., 2012; Kang MY et al., 2013). One of the cell fate decisions mediated by p53 may be via the repression of Hsp70 expression (Agoff et al., 1993). Intriguingly, both the antioxidant and prooxidant actions of p53 may be mediated by its concentration-dependent effects on catalase activity (Kang MY et al., 2013). The p53-inducible gene 3 (PIG3)-mediated downregulation of catalase activity (Kang MY et al., 2013) may explain the low level of catalase activity in rat testicular germ cells (Bauché et al., 1994). Throughout metazoan phylogenesis, p53 and its homologues mediate spontaneous germ cell apoptosis and failure to remove defective germ cells by this mechanism results in increased percentages of abnormal gametes and reduced fertility (Beumer et al., 1998; Yin et al., 1998; Sinha Hikim and Swerdloff, 1999; Gartner et al., 2000; Schumacher et al., 2001; Lettre et al., 2004; Baum et al., 2005; Kwon et al., 2005; McKee et al., 2006; Pankow and Bamberger, 2007; Codelia et al., 2010; Xu et al., 2010). Many invertebrates, such as C. elegans and Drosophila melanogaster, have only one p53 family member, which resembles more closely p63 and p73 than p53 with regard to both structural and functional aspects. The p53 members from C. elegans, CEP-1 (Derry et al., 2001; Ross et al., 2011), and from D. melanogaster, Dmp53 (Ollmann et al., 2000), are both exclusively required for germline fidelity. Unique isoforms of the p53 homolog p63, the most ancient member of the p53 gene family, are highly and specifically expressed in human testicular germ cells (Amelio et al., 2012). Phylogenetically, this expression was supported by insertion of an endogenous retrovirus with its requisite LTR with strong promoter activity that occurred 10 to 15 million years ago during primate evolution at the branching point to long-lived Hominidae. Upon DNA damage, the resulting germ cell-associated, transcriptionally active p63 suppresses proliferation and induces apoptosis in mice (Beyer et al., 2011) like in the sea anemone Nematostella vectensis (Pankow and Bamberger, 2007) selecting against germ cells that evolved genomic instability.

Human females are provided with approximately 1-2 x 106 oocytes at birth, a starting reserve that is further decimated by apoptosis to fewer than 4 x 105 at puberty (Baker, 1963). Although definitive proof has been lacking, the purge of oocytes that takes place before and after birth is widely suspected of being a mechanism of ridding the germline of genetically inferior eggs (Jansen, 2000; Jansen and Burton, 2004). Routinely in the range of 35-45 years there is a well-known decrease in oocyte number (to ~ 25000; Faddy et al, 1992) and reproductive competence (Serhal and Craft, 1989; Jansen, 1995; Faber et al, 1997) Thereafter, there can be an exponential acceleration in numeric decline (Faddy etal., 1992) to fewer than 1000 in the years immediately preceding ovarian senescence (on average, around age 50) (Richardson et al., 1987; Faddy et al., 1992; Wise et al., 1996). In the span of 4 weeks (every ovarian cycle during the reproductive years), tens or hundreds of follicles start their growth from the resting primordial state. Yet in most circumstances just one follicle each month presents its oocyte to be fertilized. The observer could reasonably suspect that he or she is witnessing a competition: waves of folliculogenesis and waves of atresia, with a very small number of winners (Medvedev, 1981; Krakauer and Mira, 1999; Jansen, 2000; Jansen and Burton, 2004). In fact, oocyte selection, at least in part, is due to competition of follicles for trophic factors such as growth hormone, IGFs, basic fibroblast growth factor and VEGF that ensure proper vascularization and provision of nutrients (Neeman et al., 1997; Schams et al., 1999; Berisha et al., 2000; Silva et al., 2009). The VEGF signaling pathway also operates as evolutionarily conserved selective agent in cellular competition during angiogenic sprouting (Jakobsson et al., 2010; Krueger et al., 2011; Blanco and Gerhardt, 2013). Bidirectional communication between oocytes and granulosa and cumulus cells is responsible for nurturing oocyte growth, the gradual acquisition of oocyte developmental competence and oocyte selection (Gilchrist et al., 2008). p63 is constitutively expressed in female germ cells during meiotic arrest and, monitoring the integrity of the female germline, is essential in a process of DNA damage-induced oocyte death (Suh et al., 2006; Gonfloni et al., 2009; Deutsch et al., 2011a; b; Amelio et al., 2012). Together with its control of recombination (Stürzbecher et al., 1996; Lu et al., 2010), p53 family members are involved in the complete cascade of gametogenic events, assessing and assuring quality and stress resilience of germ cells (Amelio et al., 2012).

The ubiquitin-proteasome system is a universal cellular quality control system (Hershko and Ciechanover, 1998; Taylor and Rutter, 2011). In spermatogenesis, the ubiquitin-proteasome system is required for the degradation of numerous proteins throughout the mitotic, meiotic, and postmeiotic developmental phases (Wilkinson, 1997; Baarends et al., 1999; 2000). Ubiquitin C-terminal hydrolase (UCH) L1 is expressed at high levels in both testis and epididymis and may play an important role in the regulation of spermatogenesis (Martin R et al., 1995; Fraile et al., 1996; Kon et al., 1999; Kwon et al., 2003). Furthermore, it has been suggested that UCHL1 also functions as a regulator of apoptosis (Harada et al., 2004b). UCHL1-deficient testes of gad mice have reduced ubiquitin levels and are resistant to cryptorchid injury-mediated germ cell apoptosis (Kwon et al., 2004). Testicular germ cells in the immature testes of gad mice are resistant to the early apoptotic wave that occurs during the first round of spermatogenesis. In adult gad mice, cauda epididymidis weight, sperm number in the epididymis, and sperm motility were reduced. Moreover, the number of defective spermatozoa is significantly increased; however, complete infertility was not detected. These data indicated that UCHL1 is required for normal spermatogenesis and sperm quality control and demonstrated the importance of UCHL1-dependent apoptosis in spermatogonial cell and sperm maturation (Kwon et al., 2005).

The heat shock transcription factor (HSF) family regulates expression of heat shock genes via a heat shock element. HSF1 is also involved in quality-control mechanisms, eliminating injured male mouse germ cells when these cells are exposed to stress (Nakai et al., 2000; Izu et al., 2004; Widlak et al., 2007). Male mice deficient in both HSF1 and HSF2 are sterile with severe defects in spermatogenesis (Wang G et al., 2004) while female mice deficient in HSF1 are infertile as well (Xiao et al., 1999). In Drosophila, a single HSF is necessary for oogenesis (Jedlicka et al, 1997).

In Hydra, oocyte determination involves a mechanism that establishes a subset of precursor interstitial cells competent to differentiate into oocytes. The oocyte is singled out from this subset and the competence of the remaining cells to become oocytes dramatically decreases as they adopt the alternative nurse cell fate. Nurse cells differentiate and enter an unusual apoptosis program where they are phagocytosed by the oocyte (Miller et al., 2000). Spermatogenesis in Hydra is associated with apoptosis of pre-meiotic spermatocytes (Kuznetsov et al., 2001). Significant apoptosis occurs premeiotically during normal spermatogenesis in teleost and cartilaginous fishes (Billard, 1969; Callard et al., 1995; 1998; Cinquetti and Dramis, 2003; Prisco et al., 2003; McClusky, 2005; 2011; Corriero et al., 2007; 2009; Leal et al., 2009; McClusky et al., 2009), amphibians (Yazawa T et al., 1999, 2000, 2003; Ricote et al. 2002; Sasso-Cerri et al., 2006) and reptiles (Comitato et al., 2006; Zhang L et al., 2008). However, spermatogonia loss may also be as low as 5% as in cod (Almeida et al., 2008).

Programmed cell death is implemented during germarium, mid- and late-oogenesis of a variety of Diptera and is absolutely required for the normal maturation of the developing egg chambers (Nezis et al., 2000; 2001; 2002; 2003; 2006a; b; c; McCall, 2004; Baum et al., 2005). Each Drosophila ovary is composed of approximately fifteen ovarioles, chains of developing egg chambers. Egg chambers are sixteen-cell germline cysts surrounded by up to a thousand somatic follicle cells (King, 1970; Spradling, 1993; Wu et al., 2008). In Drosphila oogenesis, three waves of apoptosis occur (Pritchett et al., 2009): PGC cell death (Coffman, 2003), mid-oogenesis cell death (McCall, 2004) and nurse cell death (Velentzas et al., 2007). In addition to germline cell death, cell death occurs in the somatic follicle cells (Pritchett et al., 2009). Approximately 50% of initially formed PGCs successfully migrate and are incorporated into the gonads. The remaining PGCs do not transdifferentiate but are eliminated (Sonnenblick, 1950; Underwood et al., 1980; Technau and Campos-Ortega, 1986). Removal of a phospholipid survival factor seems to be responsible for guiding PGC away from the midline and to eliminate those PGCs unable to follow (Zhang N et al., 1996; 1997; Starz-Gaiano et al., 2001; Hanyu-Nakamura et al., 2004; Renault et al., 2004; Boldajipour and Raz, 2007). Extavour and García-Bellido (2001) investigated whether mutations in heterozygosis are subject to pregametic selection in the germline and showed that cell selection that precedes and conditions subsequent zygotic selection takes place in mosaic germ-line populations. This apoptosis is mediated by Drosophila p53 pathways in which signaling between p53 and cellular metabolism are integrated to regulate programmed cell death decisions (Yamada et al., 2008). In Drosophila oogenesis, somatic follicle cells surround the egg chamber providing the proper environment for the development of the 16 germline cells that are derived from a single germline precursor cell by four mitotic divisions to generate interconnected 16-cell groups known as germline cysts. During these cell divisions, the cysts elaborate a cytoskeletal polarity that ultimately causes one cell to develop as an oocyte, while the others become nurse cells (Spradling, 1993; Johnstone and Lasko, 2001; Reichmann and Ephrussi, 2001). The nurse cells support the development of the oocyte providing organelles, proteins and maternal RNAs to the oocyte through the ring canals (Trougakos and Margaritis, 2002). The development of each mature egg is always accompanied by the apoptosis of its 15 sister nurse cells at the end of oogenesis (McCall and Steller, 1998; Buszczak and Cooley, 2000; McCall, 2004; Velentzas et al., 2007; Pritchett et al., 2009).

Like in Drosophila, a quality control system with downregulation of a survival factor seems to be responsible for guiding mouse PGCs away from the midline towards the genital ridges and to eliminate those PGCs unable to follow (Mahakali Zama et al., 2005; Runyan et al., 2006; Boldajipour and Raz, 2007).

Interestingly, in the developing Xenopus ovary containing oogonia and stage I oocytes, the apoptosis seems to be limited to the somatic cells, and there is no indication of apoptosis in the germ cells (Kloc et al., 2004b). In Xenopus oocytes, pentose phosphate pathway generation of NADPH is critical for oocyte survival. Pentose phosphate pathway-mediated inhibition of cell death resulted from the inhibitory phosphorylation of caspase 2 (Nutt et al., 2005). p53 regulates the blockade of glycolysis, directing the pathway into the pentose phosphate shunt to produce NADPH (Bensaad et al., 2006; Green and Chipuk, 2006), increasing glutathione levels which promotes the scavenging of RONS.

It has been estimated that roughly one half of all female germ cells die in the adult C. elegans hermaphrodite gonad (Gumienny et al., 1999; Gartner et al., 2000; 2008; Navarro et al., 2001; Jaramillo-Lambert et al., 2007). More recent estimates of germ cell proliferation rates indicate that approximately 20 germ cells are produced every hour, while only ~3 oocytes are laid (Fox et al., 2011; Bailly and Gartner, 2013). Apoptosis only occurs in the female and not in the male germline and is thus only present in hermaphrodites (Gumienny et al., 1999; Gartner et al., 2000; Jaramillo-Lambert et al., 2010). The core cell death machinery is expressed in female and male germlines; however, CED-3 caspase is not activated in the male germline. Intriguingly, recombination checkpoint functions in male germ cells to promote repair of meiotic recombination intermediates, thereby improving the fidelity of chromosome transmission in the absence of apoptosis (Jaramillo-Lambert et al., 2010). Oocyte cell death occurs exclusively during the adult stage and primarily in the loop region containing pachytene stage meiotic germ cells.

8.2.1 Case study: paternal age effect disorders

Advanced paternal age has been associated with an increased risk for spontaneous congenital disorders, including Apert syndrome (caused by FGFR2 mutations), achondroplasia, and thanatophoric dysplasia (FGFR3), and Costello syndrome (HRAS), that were collectively termed ‘‘paternal age effect’’ (PAE) disorders (Arnheim and Calabrese, 2009; Goriely and Wilkie, 2012). All are caused by a small number of dominantly-acting point mutations in key developmental regulators, which cluster within the growth factor receptor-RAS signaling pathway; moreover, the causative point mutations originate almost exclusively from the unaffected fathers, indicating that the original mutational events are taking place during spermatogenesis. Extensive analyses of the birth incidence of achondroplasia showed that unaffected fathers in their fifties are 10-fold more likely to have offspring with a de novo achondroplasia mutation compared with unaffected fathers in their twenties (Risch et al., 1987). Individuals born with Apert syndrome exhibit prematurely fused cranial sutures and fused fingers and toes. Two nucleotide substitutions in the human FGFR2 gene (C755G or C758G) are responsible for virtually all sporadic cases of Apert syndrome. The birth frequency of individuals with new mutations at either of these two nucleotide sites suggests that the mutation frequency at either site is 100- to 1,000-fold greater than expected based on what is known about transversion mutations since humans and chimpanzees last had a common ancestor (Nachman and Crowell, 2000) and mutation data at many human disease loci (Kondrashov, 2003). (Qin et al., 2007; Choi et al., 2008). In one case (Tiemann-Boege et al., 2002), the magnitude of the sex bias for the C755G Apert syndrome was at least 99-fold greater in the male germline than in the female germline (Wilkin et al., 1998; Glaser et al., 2000), whereas estimates of male bias (male-driven evolution) using data on neutral mutations at many different sites would indicate only an approx. five- to ten-fold male bias (see chapter 7.3).

The hotspot model, the idea that the nucleotide has a higher-than average chance of undergoing a base substitutions at both sites (C755G and C758G) has been ruled out (Qin et al., 2007; Choi et al., 2008). Various mutations associated with PAEs in testes of unaffected men were not uniformly distributed across each testis as would be expected for a mutation hotspot but were highly clustered and showed an age-dependent germline mosaicism (Qin et al., 2007; Choi et al., 2008; 2012; Shinde et al., 2013; Yoon et al., 2013). An alternate hypothesis to explain the high mutation frequencies argues that diploid premeiotic cells carrying the mutations have a selective advantage over wild-type cells (Tiemann-Boege et al., 2002; Goriely et al., 2003; 2005; Crow, 2006; Qin et al., 2007; Choi et al., 2008; 2012; Shinde et al., 2013; Yoon et al., 2013). This selection takes place on self-renewing Ap spermatogonial stem cells (SrAp) carrying the mutations that arise at approximately the frequency expected from the existing data on neutral mutations (Nachman and Crowell, 2000; Kondrashov, 2003). The selection model proposes that mutant adult SrAp occasionally divide symmetrically (inferred rate 1 out of 100 divisions on average, or approximately once every 4 y), whereas wildtype SrAp always undergo asymmetric self-renewal divisions. Rare patients with multiple mutations in the FGFR2 gene that leads to Apert syndrome were also cited as support for a germline selection model (Goriely et al., 2005). In another study (Goriely et al., 2003), the authors exploited a nearby single nucleotide polymorphism (SNP) to argue that selection acted on the C755G mutation.

In other PAE disorders, there is evidence for positive selection of mutant premeiotic spermatogonia. First, 97–99% of the de novo mutations leading to achondroplasia (the most common form of dwarfism) result from a G-to-A transition mutation at base pair 1138 (G1138A) in exon 10 of fibroblast growth factor receptor 3 (FGFR3) (Shiang et al., 1994; Rousseau et al., 1996). The cytosine at base pair 1138 is part of a CpG dinucleotide and, if methylated, is highly susceptible to mutation caused by spontaneous deamination (see chapter 10.3.1.1). Direct measurement of the G1138A mutation frequency in sperm suggested that the high frequency may be explained by selection (Tiemann-Boege et al., 2002; Crow, 2006). The observed G1138A mutation distribution in human testes fits a model, where mutant spermatogonial stem cells have a proliferative advantage over unmutated cells (Dakouane Giudicelli et al., 2008; Shinde et al., 2013).

The Noonan syndrome (NS), whose features include characteristic craniofacial abnormalities, short stature, heart defects, intellectual disability and delay, and a variety of other anomalies, as well as a predisposition to certain cancers, is among the most common Mendelian genetic diseases (~1/2,000 live births) (Tartaglia and Gelb, 2005; Allanson and Roberts, 2011). Most cases (50%–84%) are sporadic, and new mutations are virtually always paternally derived (Tartaglia et al., 2004). More than 47 different sites of NS de novo missense mutations are known in the PTPN11 gene that codes for the protein tyrosine phosphatase SHP-2. Surprisingly, many of these mutations are recurrent with nucleotide substitution rates substantially greater than the genome average; the most common mutation, A922G, exceeds the genome average A>G mutation frequency by more than 2,400 fold (Yoon et al., 2013). Data of the spatial distribution of the A922G mutation in testes from unaffected men were inconsistent with hypermutation, but consistent with germline selection: mutated spermatogonial stem cells gained an advantage that allowed them to increase in frequency (Yoon et al., 2013).

Multiple endocrine neoplasia type 2 (MEN2) is characterized by thyroid cancer, variable penetrance of tumors or hyperplasia in other endocrine organs. Half of all new cases result from sporadic mutations, the vast majority (>95%) of which arise in the male germline (Carlson et al., 1994; Kitamura et al., 1995). The average age of the males who transmit a new mutation to their children is greater than that of the average age of all fathers (Carlson et al., 1994). Since almost all sporadic cases are caused by the same nucleotide substitution in the RET proto-oncogene, the calculated disease incidence is 100–200 times greater than would be expected based on the genome average mutation frequency. The spatial distribution of the mutation in testes of unaffected men suggested that the MEN2B mutation provides an advantage for spematogonial clonal expansion by altering signaling pathways (Choi et al., 2012).

I offer an alternative interpretation: All published work missed the fact that spermatogonia undergo a massive wave of apoptosis that selects against low-quality germ cells (see chapter 8.2). Thus the observed birth frequency may not only be due to positive selection of mutated gametes but may simply reflect the lack of negative selection. Although, a variety of PAE-mutations occur in proto-oncogenes (Choi et al., 2012; Yoon et al., 2013) that may confer a competitive advange to the gametes carrying them (see chapter 8.1.1). Compared to putatively neutral mutations, the mutation rate per cell division was found not elevated at the nucleotides under study in the Apert and the heritable cancer syndromes (Qin et al., 2007; Choi et al., 2008; 2012). Moreover, the C755G mutations arise at approximately the frequency expected from the existing data on neutral mutations (Nachman and Crowell, 2000; Kondrashov, 2003) that per definition cannot be selected against during the germ cell selection process because they are not “seen” by evolution. Both the clonal expansion of mutated spermatogonia and the lack of negative selection may jointly contribute to the observed PAE. The fibroblast growth factor receptor 3 (FGFR3) gene does not appear to be a housekeeping gene since it is highly methylated both in mature sperm and oocytes (El-Maarri et al., 1998). This finding does not preclude that the gene is expressed during gametogenesis. However, it is thought that only housekeeping genes that are expressed during gametogenesis should give rise to a selective advantage or disadvantage. At any rate, the evidence from PAEs supports the notion of a germ cell selection process.

8.3 Selective mitochondrial bottleneck

Another selective process during sexual reproduction unfolds in cell organelles. Here, only the selection of mitochondria is considered, but it can be expected that a similar process of organelle competition occurs between varieties of chloroplasts (Eberhard, 1980). Mitochondrial functionality is dependent on a harmonious and concerted interplay of approximately 1,000 genes with the vast majority encoded in the nuclear genome, and usually only 37 encoded by mitochondrial DNA (mtDNA) (Mootha et al., 2003; Sickmann et al., 2003; Gregersen et al., 2012). Localization of fundamental components of mitochondrial respiration in two different compartments of the cell requires extensive anterograde (nucleus to organelle) and retrograde (organelle to nucleus) signaling to maintain respiratory efficiency (Woodson and Chory, 2008; Chacinska et al., 2009). A complex signaling network regulates the replication and transcription of mtDNA, with ATP balance and oxidative phosphorylation activity of individual mitochondria serving as potential signals (Allen, 1993). The proximity of mtDNA to the sites of oxidative phosphorylation and ROS production, the fact that mtDNA is not associated with histones, an error-prone polymerase and limited DNA repair are primary explanations for high mutation pressure on mtDNA (Shigenaga et al., 1994; Bogenhagen, 1999; de Grey, 1999). Increased mtDNA rearrangements and deletions in human gametes (Brenner et al., 1998; Reynier et al., 1998; Barritt et al., 1999) witness the high mutation rate. The mitochondrial mutation rate is much higher than the nuclear one (Brown et al., 1979; 1982; Ingman et al. 2000; Mishmar et al. 2003; Ballard and Whitlock, 2004; Kivisild et al. 2006; Lynch et al., 2006). On the other hand, ROS have been implicated in stimulating mtDNA replication in mammalian and yeast cells (Moreno-Loshuertos et al., 2006; Hori et al., 2009).

With few exceptions, mtDNA is maternally inherited (Birky, 1995; White et al., 2008). As many as 100,000 copies are passed on in mammalian oocytes. Taking into account the high developmental oxidative stress to which sperm mitochondrial DNA is exposed, it should have inferior quality, compared to oocyte mitochondrial DNA, and its elimination should be an evolutionary necessity. Various theories have been proposed to explain the uniparental inheritance of mitochondrial DNA (Birky, 1995). In Caenorhabditis elegans, the paternal mitochondria are eliminated after fertilization (Rawi et al., 2011; Sato and Sato, 2011). In Drosophila, mitochondrial DNA is removed from developing spermatids in a process requiring the mitochondrial nuclease, Endonuclease G (DeLuca and O’Farrell, 2012). In the Japanese medaka fish, paternal mtDNA vanishes within sperm mitochondria after fertilization, and in bovine embryos, paternal mitochondria are eliminated during the first two zygotic cell divisions (Nishimura et al., 2006; Sutovsky et al., 1996). Before fertilization, sperm mitochondria commonly harbor multiple DNA deletions (Reynier et al., 1998). Poorly motile sperm have higher numbers of mtDNA per cell than progressively motile sperm do. Such specimens also have a higher abundance of mtDNA species with deletions (Kao et al., 1995). In a man with the mtDNA A3243G mutation, decreasing sperm motility correlated with an increase in mutant mtDNA from 42 to 64% (Spiropoulos et al., 2002) implying that mitochondrial defects accentuate the degree of mtDNA damage that is commonly acquired during spermatogenesis and epididymal transit (Jansen and Burton, 2004). In contrast to the situation in which competing spermatozoa find themselves, the egg has for the entire time been in a relatively anaerobic and metabolically quiet environment, one suited for shielding the egg’s genomes from DNA damage and mutation (Jansen and Burton, 2004).

Here, I reproduce for the most part the excellent review of Nick Lane (2011a): “If large complex cells are not possible at all without tiny mtDNA genomes, then there is a necessary interaction between mtDNA and the nuclear genes encoding mitochondrial proteins. In other words, mosaic respiratory chains, whose protein subunits are encoded by two separate genomes, are a strictly necessary feature of eukaryotic cells; eukaryotes could not exist with any other arrangement. The trouble is that the proteins encoded by the two genomes must interact with nanoscopic precision, or electron flow down respiratory chains will be blocked. Any blockage of electron flow in an aerobic world leads to a high rate of ROS leak, a collapse in energy charge (which is to say, an irreversible fall in ATP levels), the oxidation of membrane lipids such as cardiolipin, and the release of cytochrome c. The surprising involvement of cytochrome c in apoptosis (Blackstone and Green, 1999), emerges as an explicit prediction of the hypothesis that eukaryotic cells must undergo functional selection for the compatibility of mtDNA and nuclear genes encoding adjoining respiratory chain subunits (Lane, 2011b). The speed of electron transfer down respiratory chains depends on the distance between redox centres, and slows down by about an order of magnitude per Ångstrom additional distance, for reasons that relate to the probability of transfer by quantum tunnelling (Moser et al., 2006). A likely consequence of even single nucleotide mutations or polymorphisms in mtDNA would be small misalignments in subunit juxtaposition, slowing electron transfer. Slower electron transfer increases the reduction state of respiratory complexes, making them more reactive with oxygen and therefore increasing ROS leak and susceptibility to apoptosis. Thus, any mismatch between mtDNA and nuclear genes encoding respiratory-chain subunits should increase the risk of apoptosis. Selection for mitonuclear coadaptation necessarily involves oxidative stress (Lane, 2011a; b). It has been argued that selection for mitonuclear coadaptation cannot take place during oocyte development, because at this time the new nuclear background is not known. Selection for mitonuclear coadaptation must therefore take place after fertilization, during development or after birth (Lane, 2011a). A variable apoptotic threshold has profound implications for fertility, fecundity, adaptability, fitness, aging and age-related disease. The reason is simple. Setting the apoptotic threshold high – meaning a high tolerance of ROS-leak before apoptosis is triggered – enables high fertility and fecundity. Poor mitonuclear match is overlooked, and embryos that would fail to develop in more discerning animals develop full term. Some degree of heteroplasmy (the presence of two or more mtDNA haplotypes in the same individual) is tolerated and indeed can be beneficial, as a range of mtDNA enables greater adaptability to changing environments. However, the offspring are less fit, and more likely to suffer from mitochondrial diseases. They will have lower aerobic capacity. Worst of all, they will leak ROS from their mitochondria at a faster rate, without triggering apoptosis. The outcome is a shorter lifespan, and a greater tendency to oxidative stress and chronic inflammatory conditions linked with aging, such as diabetes, cardiovascular disease and cancer. In short, there is a trade-off between fertility, fecundity and adaptability, on the one hand, and aerobic capacity, life-span and susceptibility to age-related disease on the other (Lane, 2009). The trade-off is mediated by sensitivity to oxidative stress.” (Lane, 2011a)

In mouse, the process in which many mtDNA molecules are discarded and only a few are transmitted to PGCs was thought to be random (Jenuth et al., 1996). However, the study of fitness effects of distinct mtDNA haplotypes (MacRae and Anderson, 1988; Clark and Lyckegaard, 1988; Nigro and Prout, 1990; Fos et al., 1990; Kambhampati et al., 1992; Hutter and Rand, 1995; Kilpatrick and Rand, 1995; García-Martínez et al., 1998; Rand et al., 2001) and statistical tests of neutral models of molecular evolution (Whittam et al., 1986; Excoffier, 1990; Ballard and Kreitman, 1994; Ballard, 2000a; b; Nachman et al., 1994; 1996; Nachman, 1998; Quesada et al., 1999; Rand et al., 1994; 2000; Rand and Kann, 1996; 1998; Templeton, 1996; Wise et al., 1998; Weinreich and Rand, 2000; Blouin, 2000) have identified multi-level selection of mtDNA (Birky, 1995; Bergstrom and Pritchard, 1998; Jacobs et al., 2000; Rand, 2001). The selection of mtDNA variants is likely to depend on the complex interplay of the fusion and fission apparatus, mitophagy regulation, ROS production and antioxidant defences (Skulachev, 1996b; Kanki et al., 2009; Malena et al., 2009; Zhou et al., 2010). Mitochondria from different strains and species of Drosophila have been transplanted in various combinations to construct experimental heteroplasmic lines, revealing clear evidence for selection in the cytoplasm (Matsuura et al., 1989; 1991; de Stordeur et al., 1989; de Stordeur, 1997). However, various nuclear backgrounds can have considerable effects on the selection coefficients of mtDNAs in heteroplasmic cells and germlines (Kilpatrick and Rand, 1995; Doi et al., 1999). It is evident that the mitochondrial genome is subject to strong selective pressure and to assure survival it had to develop strategies to slow down or halt the ratchet. Muller (1964) proposed that asexual genomes (like organellar genomes) will inevitably accumulate deleterious mutations, resulting in an increase of the mutational load, an inexorable, ratchet-like, loss of the least mutated class and extinction. At least two processes are thought to operate to rewind Muller’s ratchet in mtDNA (Saccone et al., 2002):

(i) the homoplasmy of mitochondrial DNA molecules during oocyte maturation (Bergstrom and Pritchard, 1998). In metazoans a severe bottleneck, restricting the number of mtDNA molecules passing through the germline, is taking place. This mechanism is essential in maintaining both mitochondrial genetic quality and mitonuclear match (Lane, 2011a; b) over evolutionary time, as it is able to restore genetic homoplasmy among descendants; moreover, the bottleneck improves mitochondrial performance and, over a longer time scale, acts to slow the progression of the ratchet.

(ii) female germ cell atresia (Krakauer and Mira, 1999). Another mechanism related to the mt genetic bottleneck is the female germ cell atresia, which reduces the population of female germ cells to a small fraction of that present in early fetal life; recent studies demonstrated that this process not only assures the selection of high-quality mitochondria, necessary for descendants, but it is also able to retard Muller’s ratchet. Indeed, there is a significant positive correlation between the number of offspring and the number of mitochondria in germline cells and a negative correlation between the number of mitochondria in oocytes and the fraction of follicles undergoing atresia. Species with small litters need to have high-quality mitochondria in the oocytes to be sure that descendants are fit for survival; for this reason the mitochondrial bottleneck needs to be more severe. Indeed, this mechanism decreases the number of mitochondria to a level that highlights the difference in functionality among cells: only cells with healthy mitochondria will be able to escape the apoptotic process occurring during atresia.

8.3.1 Mitochondrial homoplasmy

Again Nick Lane (2011a): “Mitonuclear mismatch is unavoidable. The only question is how much can or ‘should’ be tolerated. The basic problem is that the tempo and mode of evolution of the two genomes are quite distinct (see above). Species that need a high aerobic capacity, such as flighted birds or bats, could not get airborne at all if they did not have an aerobic capacity several-fold higher than even fast runners like the cheetah. On the other hand, rats do not require a high aerobic capacity, and so could presumably tolerate a poorer mitonuclear match. Put another way, there must be an adjustable threshold, above which ROS leak stimulates apoptosis and developmental failure, and below which ROS leak is tolerated, or might even be beneficial as a redox signal (Lane, 2009). In birds, the apoptotic threshold is low: they are sensitive to ROS leak from mosaic respiratory chains and quickly trigger apoptosis. A poor mitonuclear match leads to slow electron flow, high ROS leak and swift apoptosis, translating into infertility and low fecundity. An intolerance of heteroplasmy means a low incidence of mitochondrial disease but also a low adaptability to changing conditions. On the positive side, birds have a high aerobic capacity, a long lifespan and low susceptibility to the chronic inflammatory conditions characteristic of old age in mammals. The difference is not trivial. Pigeons and rats have a similar body size and similar metabolic rate, even a similar foraging lifestyle, to the point that pigeons are often dismissed as flying rats. Far from it. Rats live for 3 – 4 years, pigeons for 35, ten times longer. Their mitochondrial ROS leak is nearly 10-fold lower (Barja, 2007). While this difference makes no sense in terms of the efficiency of respiration (the proportion of ROS leak as a fraction of total oxygen consumption is too small) it makes a big difference in terms of functional selection for mitonuclear match, and it makes a big difference in terms of lifespan and healthspan.” (Lane, 2009; 2011a)

In the absence of external factors, heteroplasmy should, theoretically, be the default state for mtDNA under a simple mutation-drift scenario. The high mutation rate coupled with the large population size of mtDNA in mature oocytes would make a state of homoplasmy in any cell, tissue or individual surprising under a simple neutral drift model, in the absence of mechanisms that effectively remove genetic diversity from the mtDNA population (White et al., 2008). A state of heteroplasmy is also reached if paternal mtDNA enters the egg cytoplasm at fertilization, referred to as paternal leakage. In this case, a heteroplasmic state has been achieved in the fertilized oocyte, not via mutation, but due to the coexistence of mitochondria from two unique ancestral lineages. Until recently, paternal leakage was thought impossible as it was widely held that paternal mtDNA did not reach the egg cytoplasm (Ankel-Simons and Cummins, 1996). However, interspecific paternal leakage has been revealed in silkmoths, fruit flies, periodical cicadas and salmonids (Arunkumar et al., 2006; Sherengul et al., 2006; Ciborowski et al., 2007; Fontaine et al., 2007) and cases of intraspecific paternal leakage include scorpions, fruit flies, and lizards (Gantenbein et al., 2005; Sherengul et al., 2006; Ujvari et al., 2007). However, despite the high mtDNA polymorphism in populations, most individuals appear to be homoplasmic with respect to their mtDNA, i.e. a human individual’s mtDNA population is ca. 99.9% identical (Monnat and Loeb, 1985; Lutz et al., 2000). Mice harbouring heteroplasmy showed a significant reduction in metabolic rate, reduced activity and feeding, hyperexcitability and a decreased capacity to learn and remember, even when the two mtDNA types appear to function equally well against the same nuclear background (Jones, 2012; Lane, 2012; Sharpley et al., 2012). Mathematical modelling suggests that uncontrolled heteroplasmy can indeed lower fitness significantly, with approximately 90% of homoplasmic populations achieving 95% fitness, compared with barely 50% of heteroplasmic populations (Lane, 2011b). The existence of two sexes helps to optimize mitonuclear match by generally promoting homoplasmy (Lane, 2011b).

There are reports about individuals that are heteroplasmic with regard to mtDNA (Greenberg et al., 1983; Howell et al., 1992; 1996; Bendall and Sykes, 1995; Comas et al., 1995; Bendall et al., 1996; 1997; Jazin et al., 1996; Marchington et al., 1997; Parsons et al., 1997; Hühne et al., 1998; Lutz et al., 2000; Klütsch et al., 2011). Individuals affected by mtDNA diseases are usually heteroplasmic: most of their tissues and cells have a mixture of both normal and mutant mtDNAs (Poulton et al., 2010). There is also a threshold effect (tissues function normally unless the proportion of mutant mtDNA rises above a particular level) in most diseases. The level of this threshold varies with both tissue and mutation type, usually in the range 50 to 100% mutant mtDNA, but occasionally as low as 10% (Sacconi et al., 2008). Heteroplasmy is conceivable at three levels. These are: (i) The cell: a single cell contains only mitochondria that are homoplasmic, but different cells carry variants. (ii) The mitochondrion: one cell carries different mtDNA haplotypes, but each single mitochondrion is homoplasmic. (iii) The nucleic acid: an individual mitochondrion carries different mtDNA types (Lutz et al., 2000). Pedigree analyses of heteroplasmic individuals in cattle, mice and humans revealed that mtDNA genotypes shift rapidly among offspring and return to homoplasmy in some progeny within a few generations (Hauswirth and Laipis, 1982; Olivo et al., 1983; Ashley et al., 1989; Laipis et al., 1988; Gyllensten et al., 1991; Koehler et al., 1991; Larsson et al., 1992; Blok et al., 1997; Meirelles and Smith, 1997; Brown et al., 2001; Sharpley et al., 2012), suggesting that a mtDNA bottleneck accounts for the rapid segregation. However, other work suggests that mtDNA heteroplasmy may be stably maintained for multiple generations (Gemmell et al., 1996; Ivanov et al., 1996; Howell and Smejkal, 2000; Taylor and Breden, 2002). A rapid switch from heteroplasmy to homoplasmy, which is occasionally observed in animals, indicates that the mtDNA copy number in the primary oocyte (the size of the inheritance unit) may be as low as one or a few molecules (Koehler et al., 1991; Bendall et al., 1997). To determine the fate of mtDNA mutations, Fan et al. (2008) introduced mtDNAs containing two mutations that affect oxidative phosphorylation into the female mouse germline. The severe ND6 mutation was selectively eliminated during oogenesis within four generations, whereas the milder mutation was retained throughout multiple generations even though the offspring consistently developed mitochondrial myopathy and cardiomyopathy (Fan et al., 2008). On the other hand, in insects it may take 500 generations to return to homoplasy (Solignac et al., 1984; Harrison et al., 1985; Rand and Harrison, 1986), suggesting the mtDNA bottleneck is considerably less strong in insects compared to mammals (White et al., 2008). The switch to homoplasmy seems to be common only in animals while in plants it usually results in a different level of heteroplasmy (Kmiec et al., 2006).

Different authors proposed that the bottleneck exists at different sites and works through different mechanisms. The assumed stages include the earliest PGCs where the germ cell number is small and there is a low number of mitochondria in each PGC (Jansen and de Boer, 1998; Krakauer and Mira, 1999; Cree et al., 2008; Wai et al., 2008); during expansion of the oogonial population (Jenuth et al., 1996); during postnatal folliculogenesis, during which a clone of mtDNA replicates fast and populates the developing oocyte, diluting out preexisting mtDNA (Wai et al., 2008); during oocyte maturation, during which there is about a 100-fold increase of mtDNA, the amplification may use a limited number of template mtDNA molecules and yield one predominating genotype in the mature oocyte (Hauswirth and Laipis, 1982; Marchington et al., 1998); or during early embryonic development, cells that form the embryonic inner cell mass rather than extraembryonic tissues may receive very different ratios of heteroplasmic mtDNAs (Laipis et al., 1988; White et al., 2008). Most of the blastocyst forms extra-embryonic tissues; thus, only a subset of all cells (the inner cell mass, ICM) will contribute to the developing embryo (Hogan et al., 1986; Fleming et al., 1992). The apportionment of mitochondria to the ICM constitutes a numerical bottleneck, during which rare mtDNA haplotypes are prone to loss (Bergstrom and Pritchard, 1998). Wai et al. (2008) reported that PGCs possessed a mean of ~280 mtDNA copies and showed by directly tracking the evolution of mtDNA genotypic variance during oogenesis that a genetic bottleneck for the transmission of mtDNA occurs during postnatal oocyte maturation through replication of a subpopulation of genomes. However, another group presented evidence that the mitochondrial bottleneck occurs without reduction of mtDNA content in female mouse PGCs (Cao et al., 2007; 2009). Whether the bottleneck occurs during oogenesis (i.e. during the development of mature oocytes from PGCs), embryogenesis (i.e. in the cleaving embryo from the zygote to the establishment of the germ layers including the PGCs), or both remains uncertain (Jenuth et al., 1996; Smith et al., 2002; Cao et al., 2007; 2009; Cree et al., 2008; Wai et al., 2008). Wolff et al. (2011) measured the genetic variance in mtDNA heteroplasmy at three developmental stages (female, ova and fry) in chinook salmon. A mathematical model estimated the number of segregating units (Ne) of the mitochondrial bottleneck. Values for mtDNA Ne were 88.3 for oogenesis, and 80.3 for embryogenesis. The results suggest that the mechanism underlying the mtDNA bottleneck is conserved between fish and mammals, and showed that segregation of mtDNA variation is effectively complete by the end of oogenesis (Wolff et al., 2011).

The Balbiani body, or mitochondrial cloud, is a transient structure, containing mitochondria, Golgi, endoplasmic reticulum (ER) and RNA, that forms in the young (previtellogenic) oocytes of insects and vertebrates (Guraya, 1979; Cox and Spradling, 2003; Kloc et al., 2004a; Wilk et al., 2005; Pepling et al., 2007; Zhou et al., 2010; Voronina et al., 2011). It has been proposed that in the Balbiani body the best mitochondria for incorporation in the germline are selected (Guraya, 1979; Cox and Spradling, 2003; Kloc et al., 2004a; Pepling et al., 2007; Zhang YZ et al., 2008; Zhou RR et al., 2010; Castellana et al., 2011). Zhou RR et al. (2010) believe that a selective bottleneck occurs in the maturation of oocytes. Based on estimates of the mitochondrial inner membrane potential of mature zebra fish oocytes, they showed that the most efficient mitochondria with high-inner membrane potential tend to be recruited preferentially into the mitochondrial cloud (Zhang YZ et al., 2008; Zhou RR et al., 2010). Different distribution patterns of low- and high-inner membrane potential mitochondria in human and animal oocytes have been reported (Sun et al. 2001; Van Blerkom et al. 2003; Van Blerkom 2008; Zhang YZ et al., 2008).

8.3.2 Follicle atresia

For mitochondria, the processes of oogenesis, follicle formation and loss constitute a restriction/amplification/constraint event of the kind predicted by L. Chao (1997) for purification and refinement of a haploid genome. Maintaining the integrity of mitochondrial inheritance is such a strong evolutionary imperative that at least some features of ovarian follicular formation, function and loss can be expected to be primarily adapted to this specific purpose (Jansen and de Boer, 1998). In this vein of thought, cumulative evidence indicates that the removal of pathogenic mtDNA mutations occurs during female germline development by purifying selection (Nielsen and Weinreich, 1999, Nielsen, 2001; Fan et al., 2008; Stewart et al., 2008a; b; Poulton et al., 2010; Castellana et al., 2011; Pereira et al., 2011). Krakauer and Mira (1999) considered cellular atresia as a possible response to Muller’s ratchet, particularly in birds and mammals. In their view, only a very small number of all primordial follicles (and the primary oocytes inside them) reach maturation during the first stages of fetal life; the others undergo apoptosis. In this way, germ cells carrying ‘‘less functional’’ mitochondrial genomes are removed. In particular, species which produce small numbers of offspring tend to undergo a more severe bottleneck event, in order to guarantee the survival and viability of the (few) members of the future generation. Tracking mtDNA mutations between generations in the murine model, it has been shown that most deleterious mtDNA mutations are removed during germline development, while less detrimental mtDNA mutations are more likely to pass through the female germline and are transmitted to the next generation (Stewart et al., 2008a). Mouse proto-oocytes containing a higher proportion of frameshift mutant mtDNA that severely affected oxidative phosphorylation were eliminated by selection before ovulation (Fan et al., 2008). The authors concluded that taking into account that apoptosis in preovulatory follicles is thought to be induced by oxidative stress (Tsai-Turton and Luderer, 2006), it is conceivable that the proto-oocytes with the highest percentage of severe mtDNA mutations produce the most ROS and thus are preferentially eliminated by apoptosis.

These selection mechanisms are far from being faultless, as human mitochondrial diseases clearly demonstrate. The higher tolerance to mitonuclear mismatch in mammals, compared to birds, appears to play a role (Lane, 2011 a; b). Recent studies have suggested the presumed asexuality of mitochondrial genomes as responsible for their high mutational load (Jansen and de Boer, 1998; Stewart et al., 2008b). However, asexuality, per se, may not be the primary determinant of the high mutation load in mtDNA. Very little sex and recombination is required to counter mutation accumulation (Pamilo et al., 1987; Charlesworth et al., 1993; Green and Noakes, 1995; Hurst and Peck, 1996; Chasnov, 2000; Keightley and Eyre-Walker, 2000; Bengtsson, 2003; D’Souza et al., 2006; Beukeboom, 2007; Haddrill et al., 2007; D'Souza and Michiels, 2010; Lampert and Schartl, 2010), and recent evidence suggests that mitochondrial genomes do experience occasional recombination (Lunt and Hyman, 1997; Ladoukakis and Zouros, 2001; Burzynski et al., 2003; Rokas et al., 2003; Kraytsberg et al., 2004; Piganeau et al., 2004; Barr et al., 2005; Tsaousis et al., 2005; White et al., 2008; Neiman and Taylor, 2009). Instead, a high rate of accumulation of mildly deleterious mutations in mtDNA may result from the small effective population size associated with effectively haploid inheritance (Neiman and Taylor, 2009). The asymmetry in mtDNA inheritance becomes problematic in the case of traits that affect exclusively males and shared traits that, if compromised, have a disproportionally greater effect on males than females (Frank and Hurst, 1996; Zeh and Zeh, 2005). In this instance mutations that harm males but leave females unaffected will escape purifying selection and lead to the accumulation of a mutational load in the mitochondrial genome detrimental to male-specific traits; a scenario described as “mother’s curse” (Ruiz-Pesini et al., 2000b; Gemmell et al., 2004; Wolff and Gemmell, 2012).

Conversely, the ‘‘internal quality control’’ of mtDNA takes place in specific physiological conditions (maturation of oocytes) and does not act in all possible situations of energy requirements of the organism (embryo, fetus, adult, restricted diet, cold environment, etc.). Energy demands do differ and are contradictory, imposing specific ‘‘trade-offs’’ for each functional state of the oxidative phosphorylation machinery (Das, 2006).

In Drosophila, a selective transfer of a subgroup of mitochondria with some specific feature from nurse cells to a specific position of the oocyte does occur (Cox and Spradling, 2006; Zhou et al., 2010). The colossal purge of female germ cells in humans, mice and Drosophila may play an important role in eliminating deleterious mtDNAs if wild-type mtDNA or mtDNA with advantageous characteristics has been selectively transported into the destined oocytes before the purge takes place (Zhou et al., 2010).

The competition between follicles to deliver the oocyte that will be fertilized has a mitochondrial component (Jansen and Burton, 2004). Mitochondria play an important role in nuclear and cytoplasmic oocyte maturation since they provide the main supply of ATP (Krisher and Bavister 1998; Stojkovic et al., 2001). Dysfunctional mitochondria and subsequent low ATP production is one of the major factors that compromise oocyte quality and fertilization (Van Blerkom et al., 1995; Stojkovic et al., 2001; Dumollard et al., 2004; 2006; 2007b; Tamassia et al., 2004; Van Blerkom, 2004; Brevini et al., 2005; Takeuchi et al., 2005; Zeng et al., 2007; Yu et al., 2010). Levels of mtDNA deletions are greater in unselected preovulatory human oocytes than in early embryos (Brenner et al., 1998; Perez et al., 2000), arguing for selection at the ovulation or/and early embryo stage (Jansen and Burton, 2004). Intracytoplasmatic injection of ‘normal’ mitochondria can overcome mitochondrial dysfunctions (Nagai et al., 2004) and inhibit oocyte apoptosis (Perez et al., 2000), while injection of abnormal mitochondria induces oocyte apoptosis (Perez et al., 2007). Selection of functional mtDNA genomes likely involves a mechanism for functional testing to prevent transmission of mutated genomes to the offspring (Stewart et al., 2008a; b). Low mitochondrial DNA content, due to inadequate mitochondrial biogenesis or cytoplasmic maturation, may also adversely affect oocyte fertilizability (Reynier et al., 2001).

The existence of a female germline filter for severely deleterious mtDNA mutations makes evolutionary sense. Assuming that mtDNA variation is pivotal to species adaptation to changing environments and that the uniparental mtDNA cannot generate diversity by recombination, then mtDNA diversity must be generated through a high mutation rate (Ruiz-Pesini et al., 2004; Ruiz-Pesini and Wallace, 2006; Wallace, 2007). However, a high mutation rate would generate many highly deleterious mutations that could create an excessive genetic load and endanger species fitness. This dilemma can be resolved by the addition of a graded filter in the female germline that eliminates the most severe mutations before fertilization (Fan et al., 2008; Stewart et al., 2008a; b; Zhou RR et al., 2010). For such a filter to succeed, multiple cell divisions resulting in a large population of proto-oocytes would be required to segregate out the new mtDNA mutations. Mitochondrial function may also represent a quality control system in the early embryo that will determine whether the embryo proceeds further into development or is quickly eliminated (Dumollard et al., 2007b; Lane, 2011b).

8.4 Gametic selection

According to Lewontin (1970), gametic (or gamete) selection means specifically the differential motility, viability, and probability of fertilization of gametes that arises from their own haploid genotype, independent of the genotype of the parents which formed them. Haploid selection gives the unique opportunity to expose haploid genes to selection in organisms with a predominant diploid stage in the life cycle. In haploid individuals, composed of a single set of chromosomes, all novel adaptive mutations are immediately ‘seen’ by evolution, and selection is very efficient. In diploids (composed of two sets) or polyploids (multiple sets), mutations generally arise in a single copy that can be partially or completely masked by wild-type alleles. The efficacy of selection depends on the dominance properties of the mutation in question (Orr and Otto, 1994). Selection may act quickly in the case of a fully dominant allele, or slowly if the mutation is recessive and must appear in the same genome in multiple copies before its effects are exposed. This may reduce the rate of adaptation in diploid populations (Gerstein and Otto, 2011) and may prevent the establishment of recessive or partially recessive adaptive mutations (Anderson and Sirjusingh, 2004). Comparing the effect size of 20 adaptive mutations in haploids and homozygous diploids of the budding yeast Saccharomyces cerevisiae, Gerstein (2012) found that the same mutations often had a much larger effect in haploids than homozygous diploids. Further, haploidy together with periods of epigenetic stripping during reprogramming (see chapter 10.3.4) can be expected to unmask alleles with deleterious effects. Opportunity for haploid gene expression is limited in eggs given that they spend little or no time in the haploid phase: in many animals the final meiotic division of egg development takes place after fertilization (McCarter et al.,1999), whereas in others it happens just before fertilization (Erickson, 1990). Selection on genes expressed in haploid phases may occur at quite different quantities: in pollen, up to 65% of the genome might be expressed and in spermatids of mammals just 1.3–3.8% (Joseph and Kirkpatrick, 2004). However, expression during the haploid phase is not necessary for haploid selection. Some parts of the genome that are silent in the haploid phase can nevertheless experience haploid selection. The extreme compaction of DNA in sperm is accomplished by protamine 1 and 2 (Fuentes-Mascorro et al., 2000), which interact with specific recognition sites scattered throughout the sperm genome (Gatewood et al., 1987). The number, location and sequence of these sites affects the density and conformation of the DNA in sperm and are likely to be under haploid selection, even though they are transcriptionally silent (Joseph and Kirkpatrick, 2004). Sperm nuclear chromatin condensation during spermiogenesis is achieved through ROS-induced formation of disulfide bonds between cysteine residues in protamines (Aitken et al., 2004).

During yeast gametogenesis widespread apoptotic-like DNA fragmentation in coordination with an unusual form of autophagy occurs that is most similar to mammalian lysosomal membrane permeabilization and plant autophagic cell death (Eastwood et al., 2012). A remarkable feature of megasporogenesis, as seen in the heterosporous fern Marsilea and almost all seed plants, is the regular apoptotic abortion of three of the products of meiosis often shortly after meiosis (Bell, 1996; Wu and Cheung, 2000; Papini et al., 2011).

In Hydra, spermatogenesis is associated with massive apoptosis of pre-meiotic spermatocytes and post-meiotic spermatozoa (Kuznetsov et al., 2001). In teleosts, a certain level of apoptosis is normal and although fish spermatogenesis is comparatively efficient in this regard, still 30–40% of all germ cells that could be produced theoretically become apoptotic before differentiating to spermatozoa (Billard, 1969; Vilela et al., 2003; Schulz et al., 2010). In teleosts, apoptotic germ cells have been observed mainly during the spermiogenic phase (Billard, 1969; Vilela et al., 2003). Sertoli cells remove and recycle this material very efficiently, so that cellular debris rarely appears in spermatogenic tubules (Scott and Sumpter, 1989). Indeed, even when a loss-of-function of a protein required during meiosis leads to apoptotic loss of all spermatocytes in zebrafish (Feitsma et al., 2007), the fraction of apoptotic spermatocytes visible in histological sections from mutant males is only ~10% of the total number of spermatocytes, suggesting a rapid clearance of material phagocytosed by Sertoli cells (Schulz et al., 2010). In addition, in many seasonal breeders, Sertoli cells play an important role in the phagocytosis of residual sperm after the spawning season (Billard and Takashima, 1983; Besseau and Faliex, 1994; Grier and Taylor, 1998; Almeida et al., 2008).

The occurrence of spermatogenic cell apoptosis at various stages of spermatogenesis has been reported (Roosen-Runge, 1973; Clermont, 1972; de Kretser and Kerr, 1988; Allan et al., 1992; Kerr, 1992; Shikone et al., 1994; Brinkworth et al., 1995; Callard et al., 1995; Sinha Hikim and Swerdloff, 1999). Degeneration of different types of spermatogonia (A2–A4) and meiotic cells in rats has been reported to result in the depletion of up to 75% of the mature sperm pool (Kerr, 1992; Tapanainen et al., 1993). Overall, during mammalian spermatogenic differentiation, more than half of the differentiating spermatogenic cells die by apoptosis before maturation into spermatozoa (Huckins, 1978; Dunkel et al., 1997; Braun, 1998; Sinha Hikim and Swerdloff, 1999). Only a limited number of apoptotic spermatogenic cells, however, are detectable when testis sections are histochemically examined. This is probably due to the rapid elimination of apoptotic cells by phagocytosis. In fact, electron microscopic studies with rodent testis sections reveal that degenerating spermatogenic cells are engulfed by Sertoli cells (Russell and Clermont, 1977; Chemes, 1986; Pineau et al., 1991; Miething, 1992; Francavilla et al., 2002). Studies have demonstrated the existence of a ‘pachytene checkpoint’, the triggering of which at late meiotic prophase leads to the elimination of defective germ cells by apoptosis (Roeder and Bailis, 2000). Surprisingly, there is evidence that this checkpoint, compared to spermatogenesis, is missing or less stringent in mammalian oogenesis (Hunt and Hassold, 2002; Pacchierotti et al., 2007), another sign of a relaxed quality control in oogenesis. Gamete phosphatidylserine translocation is a sensitive marker for increased oxidative stress and DNA damage (Limoli et al., 1998; Barroso et al., 2000; Shamsi et al., 2009). Catalysis of selective phosphatidylserine oxidation during apoptosis by cytosolic cytochrome c precedes its translocation from the inner to the outer leaflet of the plasma membrane (Vermes et al., 1995; Kagan et al., 2000; 2002; Tyurina et al., 2000; 2004; Jiang J et al., 2003). This translocation of phosphatidylserine is one of the earliest features of cells undergoing apoptosis (Martin SJ et al., 1995; Rimon et al., 1997). Sertoli cells recognize apoptotic spermatogenic cells through the binding of their surface receptor, class B scavenger receptor type I, to phosphatidylserine that is expressed on the surface of spermatogenic cells during apoptosis (Mizuno et al., 1996; Shiratsuchi et al., 1997; 1999; Kawasaki et al., 2002). These relationships link germ cell redox homeostasis with mutagenesis/DNA damage and ensure the quality control aspect of apoptotic removal that is mediated by ROS-dependent phosphatidylserine externalization. The limitation of trophic testosterone plays an essential role in spermatid apoptosis (Russell and Clermont, 1977; O’Donnell et al., 1994; 1996). The Bcl2 modifying factor (Bmf) is a pro-apoptotic member of the Bcl2 family of apoptosis-related proteins that has been shown to initiate apoptosis in response to the loss of attachment of cells from their basal lamina (anoikis). Experimental reduction in intratesticular testosterone concentration brings about the death of spermatids that express Bmf as a consequence of their sloughing from Sertoli cells (Show et al., 2004). The ubiquitin-dependent sperm quality control, residing in the epididymal epithelium, has the ability to detect spermatozoa with apoptotic or necrotic DNA, while spermatozoa with defects other than DNA fragmentation are also recognized and ubiquitinated (Sutovsky et al., 2001; 2002).

Mammalian oocytes exhibit increased abnormalities in fertilization, chromosome segregation, cleavage divisions, and stress response with age (Tarin, 1996; Blondin et al., 1997; Goud et al., 1999; te Velde and Pearson, 2002; ESHRE Capri Workshop Group, 2005; Jones, 2008; Tamura et al., 2008). In humans, reproductive cessation occurs a decade before menopause, suggesting that declining oocyte quality, rather than quantity, is the major cause of maternal age-associated infertility and birth defects (ESHRE Capri Workshop Group, 2005). Similarly, several C. elegans studies show that oocyte quality is the limiting factor for aging-related reproductive capacity decline (Andux and Ellis, 2008; Luo S et al., 2010; Luo and Murphy, 2011).

The inhibition of phagocytosis in live animals resulted in a decrease in the number of epididymal sperm, indicating that phagocytosis of apoptotic spermatogenic cells by Sertoli cells is required for the efficient production of sperm (Maeda et al., 2002; Nakanishi and Shiratsuchi, 2004).

The differing autosomal recombination rates in males and females, known as the Haldane-Huxley rule (Lenormand and Dutheil, 2005), may be a corollary of the differing mutation rates in male and female gametes keeping mutation and recombination effects separate and exposing the more mutated haploid to different selection pressures than the more recombined haploid. In more than 15% of animal species, most notably in Hymenoptera, normal males are haploid and arise from unfertilized eggs, while females are diploid. In these species the principles of haploid selection concerning beneficial and deleterious mutations extend to the adult period.

8.5 Fertilization selection

Sperm are the most diverse cell type known: varying not only among- and within-species, but also among- and within-ejaculates of a single male (Crean et al., 2012). While the aspects of between-male sperm competition have been widely covered (Short, 1979; Harcourt et al., 1981; Smith, 1984; Birkhead and Møller, 1998; Birkhead, 2000; Simmons, 2001; 2005; Anderson and Dixson, 2002), the issue of sperm competition within a single male has been rarely taken into account (but see Sivinski, 1984; Manning and Chamberlain, 1994). Already the definition of sperm competition: “competition between the sperm from two or more males for the fertilization of the ova” (Parker GA, 1998) highlights this limited view. However, in the vast majority of taxa, males ejaculate much more sperm than would be required to fertilize each egg. Within-ejaculate variation in sperm phenotype is always present, and often substantial (Immler et al., 2008; Crean et al., 2012) but the adaptive value of this variation so far has remained unclear (Higginson and Pitnick, 2011). Within-ejaculate variation in sperm phenotype is generally attributed to developmental errors during spermatogenesis and/or poor quality control by the male (Hunter and Birkhead, 2002). Sperm numbers will be increased by: (i) the number of loci which affect sperm competitiveness in the haploid state; (ii) the mutation rate; and (iii) the recombination rate (Manning and Chamberlain, 1994). A correlation between recombination rates and sperm numbers is therefore to be expected. Thus, even when a single male mates with a female, sperm competition for fertilizations of the limited number of ova takes place. Sperm of an ejaculate bear a coefficient of relatedness of 0.5 to many or most of ejaculated siblings with quality variation as the basis for sperm selection. Given that sperm have undergone more or less extensive mutagenesis and consecutive quality control, there is harsh selection for the most viable sperm. Female spermicidal reproductive tracts are common in birds, mammals and invertebrates, and might permit females to bias fertilization in favor of robust or preferred sperm (Birkhead et al., 1993b; Holman and Snook, 2006; 2008). In birds, sperm selection occurs in the vagina soon after insemination, with only a few percent of inseminated sperm being retained by the female (Bakst et al., 1994; Birkhead and Brillard, 2007; Calhim et al., 2007). In the promiscuous fowl Gallus gallus, where social status determines copulation success, dominant males produce more sperm than subordinates but the quality of dominant males’ sperm decreases over successive copulations, whereas that of subordinates remains constant. Experimentally manipulating male social status demonstrated that ejaculate quality is a response to the social environment rather than the result of intrinsic differences between dominant and subordinate males (Cornwallis and Birkhead, 2007). Dominant males increased ejaculate quality in response to female sexual ornamentation, which signals reproductive quality, by adjusting the number and quality of sperm they transferred, whereas subordinate males did not (Cornwallis and Birkhead, 2007). There may also be selection at the level of the sperm storage structures since the length of sperm and length of the sperm storage structures positively covary across bird species (Briskie and Montgomerie, 1992; 1993; Briskie et al., 1997). Similar patterns have been reported in insects (Dybas and Dybas, 1981; Pesgraves et al., 1999; Morrow and Gage, 2000; Minder et al., 2005) and mammals (Gomendio and Roldan, 1993; Anderson et al., 2006) and have been interpreted as an example of male-female coevolution, possibly mediated by sexual conflict over fertilization.

In mammals, sperm have to survive the physical and chemical stresses within the fairly inimical environment of the female reproductive tract that ensures that only sperm with normal morphology and vigorous motility will be the ones to succeed (Suarez and Pacey, 2006). Basic mammalian biology dictates that survival of a species depends on the ability of sperm to fertilize eggs, which is influenced by many factors. At least in humans, one of these factors is the quality of mitochondria, as they determine the speed of sperm on its way towards the egg (Moore and Reijo-Pera, 2000; Ruiz-Pesini et al., 2000a; b; Jansen and Burton, 2004). This appears to be a race against time, since (i) sperm may be killed by mild acidity and natural osmotic decrease, (ii) a healthy human vagina has a pH of 4.0 to 4.5 and (iii) ejaculate acts as an alkaline buffer for several hours only (Olmsted et al., 2000; Chen and Duan, 2011). On the other hand, the physical contact to the female epithelial cells and the stress to which the spermatozoa are exposed is required for their capacitation (Kervancioglu et al., 1994; Smith, 1998; Ford, 2004). The female spermicidal reproductive tracts require a high sperm/egg ratio so that in hamster a consistent fertilization occurs only when the relative gamete ratio is above 103.5 sperm/egg × ml, due to reduced sperm survival at low sperm concentrations (Cummins and Yanagimachi, 1982; Stewart-Savage and Bavister, 1988). Intriguingly, spermatozoa may form, at least temporary, alliances (Moore et al., 2002; Immler et al., 2007; Pizzari and Foster, 2008; Pitnick et al., 2009; Fisher and Hoekstra, 2010; Higginson and Pitnick, 2011). Apparently, it is advantageous for sperm to cooperate either for faster locomotion (Moore et al., 2002) and/or to withstand the harsh conditions in the female reproductive tract. Male domestic cats appear to use two different reproductive strategies. Compared with normospermic counterparts, teratospermic cats have a higher sperm output achieved by more sperm-producing tissue, more germ cells per Sertoli cell, and reduced germ cell loss during spermatogenesis. Gains in sperm quantity are produced at the expense of sperm quality (Neubauer et al., 2004). From a 50- to 500-million spermatozoa in a human male ejaculate less than 1000 will find their way to the neighborhood of a competent oocyte in the oviduct of a fertile partner and these should be the ‘best of the best’ (Suarez and Pacey, 2006; Anand-Ivell and Ivell, 2011). Similarly, outcome of assisted reproductive techniques is predicted by sperm DNA quality (Duran et al., 2002; Morris et al., 2002; Agarwal and Allamaneni, 2004; Sharma et al., 2004).

Another fertilization selection concerns sperm mitochondria. Mammals inherit mitochondria from the mother only, even though the sperm contributes nearly one hundred mitochondria to the fertilized egg. This strictly maternal inheritance of mitochondrial DNA arises from the selective destruction of sperm mitochondria (Hutchinson et al., 1974; Giles et al., 1980) that are tagged by the recycling marker protein ubiquitin (Ciechanover, 1994). This imprint is a death sentence that is written during spermatogenesis and executed after the sperm mitochondria encounter the egg’s cytoplasmic destruction machinery (Sutovsky et al., 1999).

Polyandry is widespread, with females of most taxa mating with more than one male (Birkhead and Møller, 1998). When females copulate with multiple males during a single reproductive episode, sperm from these males compete to fertilize the female’s ova (Parker, 1970; Dziuk, 1996; Rowe and Pruett-Jones, 2011). Sperm competition offense occurs when a male copulates with a previously-mated female, and can be quantified as the proportion of offspring sired by the second male to mate a female (P2) (Simmons, 2001). Under conditions of sperm competition, male fertilization success is commonly determined by the number of sperm, relative to rival males, transferred during copulation (Laskemoen et al., 2010; Martin et al., 1974; Parker, 1982; Birkhead and Pizzari, 2002; Boschetto et al., 2011). Drosophila males produce seminal fluid proteins that have substantial effects on sperm transfer, sperm storage, female receptivity, ovulation, and oogenesis (Wolfner, 1997). P2 appears to be condition-dependent, e.g. affected by rearing conditions of the rival males (Amitin and Pitnick, 2007; McGraw et al., 2007; Clark NL et al., 2012). Social stress and dominance hierarchy also determine the competitive success of inseminations (Montrose et al., 2008; Thomas and Simmons, 2009). Across a diverse range of taxa, comparative and experimental studies have demonstrated that a common evolutionary response to sperm competition is an increase in testes size (Møller, 1991; Harcourt et al., 1995; Hosken, 1997; Stockley et al., 1997; Hosken and Ward, 2001; Byrne et al., 2002; Simmons and García-González, 2008; Rowe and Pruett-Jones, 2011). In addition to testes size, sperm competition may select for increases in sperm production: species under higher sperm competition have a greater proportion of sperm-producing tissue within the testes (Schultz, 1938; Lüpold et al., 2009; Rowe and Pruett-Jones, 2011). Furthermore, inter- and intra-specific studies suggest that sperm competition is positively associated with greater numbers of sperm (i.e. sperm reserves or ejaculate size) (Stockley et al., 1997; Gage, 1991; Birkhead et al., 1993a; Firman and Simmons, 2010; Rowe and Pruett-Jones, 2011), at least partially because larger testes produce more sperm (Amann, 1970; de Reviers and Williams, 1984; Møller, 1988a; b; 1989).

Sperm swimming velocity is positively correlated with successful fertilizations in a number of vertebrate species. Sperm motility influences paternity success in birds (Birkhead et al., 1999; Donaghue et al., 1999; Burness et al., 2004; Denk et al., 2005; Pizzari et al., 2008), fish (Gage et al., 2004; Gasparini et al., 2010; Boschetto et al., 2011), and mammals (Holt et al., 1989; Moore and Akhondi, 1996; Malo et al., 2005). In 42 North American and European free-living passerine bird species, sperm swimming speed was positively related to the frequency of extrapair paternity (a proxy for the risk of sperm competition) and negatively associated with clutch size (a proxy for the duration of female sperm storage) suggesting both sperm competition and female sperm storage duration as evolutionary forces driving sperm swimming speed (Kleven et al., 2009).

Mollusks and teleosts employ external or internal fertilization. Sperm competition in the course of external fertilization is mainly provided through the sperm number (Yund, 2000). Yet, sperm quality may also play a role. In a marine invertebrate with external fertilization, Styela plicata, offspring sired by longer-lived sperm had higher performance compared to offspring sired by freshly-extracted sperm of the same ejaculate (Crean et al., 2012). Sperm longevity is negatively related to sperm length. Mean sperm length is shorter in free spawners compared to teleosts with internal fertilization (Ball and Parker, 1996). Across fish species, there is a positive association between the intensity of sperm competition and sperm swimming speed (Fitzpatrick et al., 2009). In the swordtail Xiphophorus nigrensis, an internally fertilized fish with alternative reproductive tactics, males with greater sperm viability sired more offspring than their rival, as predicted if the number of fertilization-capable sperm influences sperm competition in a numerical raffle (as tickets in a raffle = lottery). In contrast, males with faster swimming sperm sired fewer offspring, but only when sperm were stored prior to fertilization (Smith CC, 2012).

In insects, sperm viability has been shown to influence competitive fertilization success (García-González and Simmons, 2005), and polyandrous species have been shown to have a greater proportion of viable sperm available for ejaculates relative to monandrous species (Hunter and Birkhead, 2002). Female yellow dung flies (Scathophaga stercoraria) control the rate of sperm release from the spermatheca, utilizing sperm more efficiently when it is limited. This mechanism is likely to explain the greater female effects on paternity observed in second clutches compared with first clutches (Sbilordo et al., 2009). Similarly, in situ observation of the female reproductive tract in a polyandrous moth revealed that contractions of the spermatheca eject around 50% of sperm from the first mate, explaining the strong last male precedence (Xu and Wang, 2010). Up to 4000 sperm are transferred to a Drosophila female during mating but only 25–35% of sperm enter the sperm-storage organs. Before the first egg is ovulated, unstored uterine sperm are expelled (Bloch Qazi et al., 2003). Postcopulatory gametic selection in D. melanogaster includes the early release of resident sperm from specialized sperm-storage organs, the heterogeneous distribution of competing sperm proportions in different storage organs and fair raffle sperm use in the seminal receptacle (Manier et al., 2010; Schnakenberg et al., 2012). From a single mating, females can lay between 300 and 400 eggs and remain fertile for nearly two weeks (Schnakenberg et al., 2012). In C. elegans hermaphrodites, fertilization occurs in the spermatheca by the first sperm to contact the oocyte. Not every oocyte is fertilized because oocytes are made in excess (Ward and Carrel, 1979). Since apoptosis during gametogenesis only occurs in the female and not in the male germline (Gumienny et al., 1999; Gartner et al., 2000; Jaramillo-Lambert et al., 2010), at least in hermaphrodites, C. elegans gamete quality selection may be exerted only in the female germline. Hermaphrodites make smaller sperm than males (LaMunyon and Ward, 1998; 1999). When males copulate with hermaphrodites, they displace the hermaphrodite sperm from the spermathecal walls and preferentially fertilize the oocytes. This preferential fertilization appears to be due in part to inhibition of hermaphrodite sperm fertility by the male sperm (Ward and Carrel, 1979). Moreover, larger sperm outcompete smaller sperm and male sperm competition favors larger sperm (LaMunyon and Ward, 1998; 1999; 2002).

In most entelegyne spiders (e.g., orb web spiders, jumping spiders, and wolf spiders), female genitalia are bilaterally symmetrical with paired copulatory openings, each leading into separate sperm storage organ. The sperm of 2 males could be therefore stored separately, avoiding direct sperm competition (Herberstein et al., 2011). Genital damage in male spiders is surprisingly common in araneoid spiders, including in the genus Argiope (Miller JA, 2007). Males break off parts or the entire pedipalp, which then becomes lodged as a plug in the female genitalia, preventing another male from utilizing this copulatory opening (Nessler et al., 2007; Foellmer, 2008; Uhl et al., 2010). Storage in different sites provides the female with a relatively easy mechanism to favor ejaculates of one male over another by selectively storing, activating, or utilizing sperm of one or the other spermatheca (Hellriegel and Ward, 1998; Herberstein et al., 2011). Indeed A. bruennichi females favor males that provided longer courtship by allowing them greater paternity shares (Schneider and Lesmono, 2009). Furthermore, female A. lobata store less sperm from related males if they fill the second spermatheca (Welke and Schneider, 2009). Male spiders are able to produce secretion in the testis and/or deferent duct, with one or several types of seminal secretion present (Michalik and Uhl, 2005; Michalik and Huber, 2006) that may influence fertilization success (Aisenberg and Costa, 2005).

Polyandry remains one of the most controversial topics in evolutionary biology (Yasui, 1997). This is primarily because in most species, females derive no direct benefits from mating with many males, but frequently incur direct costs (Chapman et al., 1995; Blanckenhorn et al., 2002). Field studies of vertebrates suggest, and laboratory experiments on invertebrates confirm, that even when males provide no material benefits, polyandry can enhance offspring survival (Madsen et al., 1992; Tregenza and Wedell, 1998; 2000; Jennions and Petrie, 2000; Zeh and Zeh, 2001; Ivy and Sakaluk, 2005; Fisher et al., 2006). Superior sperm competitors sire higher-quality young (Hosken et al., 2003; Fisher et al., 2006).

Benthic marine animals also compete for fertilizations (Yund and McCartney, 1994). In angiosperms, the interaction between pollen and stigma is one of the most important stages in the life cycle of a flowering plant, because its outcome determines whether fertilization will occur and thus whether seed will be set. For fertilization to be achieved, pollen must establish molecular congruity/compatibility with the stigma and then, following production of a pollen tube, with the transmitting tissue of the style and ovary as the pollen tube grows through the pistil to deliver its two sperm cells to an ovule. Gametophytic competition among growing tubes may produce more vigorous or more competitive progeny, apparently because faster growing pollen genotypes, in a pollen-tube race, fertilize ovules first and transmit their metabolic superiority to the seeds (Mulcahy 1974; Lee, 1984; Mulcahy and Bergamini-Mulcahy, 1987; Thomson, 1989; Hormaza and Herrero, 1994; Mulcahy et al., 1996). This process has been linked to the evolutionary success of angiosperm (Mulcahy, 1979). In addition to pollen-tube growth rate selection, maternal sorting of donor pollen has been observed (Marshall and Ellstrand, 1988). Stigmas from various different angiosperms were found to accumulate ROS, predominantly H2O2, constitutively, possibly as a means to select for stress-resistant pollen. ROS/H2O2 amounts appeared reduced in stigmatic papillae to which pollen grains had adhered (McInnis et al., 2006). The measured outcrossing rate is frequently higher than that expected from the relative amounts of self versus cross pollen deposited on the stigma (Husband and Schemske, 1996). This selective filtering of pollen or zygotes can occur in a variety of forms, including the rejection of self pollen or selfed ovules with self-incompatibility (Seavey and Bawa, 1986; Dickinson, 1994; de Nettancourt, 1997; Lipow and Wyatt, 2000), and cryptic self-incompatibility caused by differential pollen-tube growth (Weller and Ornduff, 1977; Cruzan and Barrett, 1993; Jones, 1994; Eckert and Allen, 1997).

8.6 Embryo and offspring selection

Embryo and offspring selection is another level of quality selection.

A general pattern of angiosperm reproduction is that parental provisioning of offspring (seed filling) is delayed until after fertilization. The ability to assess offspring quality establishes the basis for the maternal sporophyte to provision only the highest quality offspring (Temme, 1986). Evidence suggests that higher quality offspring are preferentially provisioned and sired (Bertin, 1982; Bookman, 1984; Bertin and Peters, 1992), the differential provisioning of outcrossed versus selfed embryos or fruits (Stephenson, 1981; Marshall and Ellstrand, 1986; Stephenson and Winsor, 1986; Rigney, 1995; Korbecka et al., 2002), and the death of selfed embryos expressing lethal recessive alleles (Lande et al., 1994; Husband and Schemske, 1996). Offspring of lower quality are selectively aborted in a variety of plants (Andersson, 1993; Kärkkäinen et al., 1999; Ramsey and Vaughton, 2000; Melser and Klinkhamer, 2001; Korbecka et al., 2002; Berg, 2003). In gymnosperm such as cycads or conifers as the spruce there are several embryos that engage in an intense life and death competition during their development. Only one embryo reaches its full term of growth to become the seed embryo, while the weaker individuals are aborted in the earlier stages (Buchholz, 1922; Mock and Parker, 1997).

In plants and benthic aquatic invertebrates, a persistent seed bank may be an additional level of zygote selection. In the wild, plant species, particularly annual ones, may have strongly fluctuating population sizes due to poor survival or seed set and, to avoid extinction, have to rely on a persistent seed bank (Lunt, 1990; Thompson K et al., 1997; Bekker et al., 1998). A seed bank has the effect of overlapping the generations of the population, integrating the effects of selection over long periods of time (Templeton and Levin, 1979; Nunney, 2002). Plants are subject to strong interspecific competition that substantially limits seed recruitment (Harradine and Whalley, 1980; Morgan, 1995; Hitchmough et al., 1996; Miriti et al., 2001; Pettit and Froend, 2001; Silvertown et al., 2002; Fréville and Silvertown, 2005). Mortality selection on plants is probably strongest at the seed and seedling stages (Huey et al., 2002).

Because variation in maternal investment directly impacts the number of surviving offspring (Williams, 1994), it can have strong implications for the fitness of both parents. The ‘differential allocation hypothesis’ predicts that parents trade-off their current and future reproduction and that the attractiveness of the mate affects the optimal trade-off between these two components of reproduction (Burley, 1986; 1988; Simmons, 1987; Sheldon, 2000; Harris and Uller, 2009; Pischedda et al., 2011). Female reproductive investment increases when they mate with males of larger size in seed beetles (Fox et al., 1995), dung beetles (Kotiaho et al., 2003), house crickets (Head et al., 2006), crayfish (Galeotti et al., 2006) and Australian rainbowfish (Evans et al., 2010). Also, male attractiveness and ornamentation influences female reproductive investment in several bird species (Burley, 1986; 1988; De Lope and Møller, 1993; Petrie and Williams, 1993; Gil et al., 1999; Cunningham and Russell, 2000; Limbourg et al., 2004; Uller et al., 2005; Velando et al., 2006; Helfenstein et al., 2008). The majority of these examples provided females several opportunities to assess male quality, either by lengthy male–female interactions, nuptial gifts to the female or extensive parental care, enabling females to alter their reproductive investment accordingly. If the reproductive value of some of the young is low, parents might decrease investment to them, and so either increase the investment given to survivors or enhance their capacity for future reproduction (O'Connor, 1978; Tait, 1980; Gosling, 1986; Wright et al., 1988; Haig, 1990; Mappes et al., 1997).

Intriguingly in both invertebrates and vertebrates, DNA lesions that are carried by spermatozoa can be repaired in the embryo by maternal gene products (Generoso et al., 1979; Brandriff and Pedersen, 1981; Ashwood-Smith and Edwards, 1996; Marchetti et al., 2007). DNA repair does not occur post-meiotically in male D. melanogaster, but DNA damage carried in sperm can be repaired post-fertilization in D. melanogaster embryos via maternal repair enzymes. Agrawal and Wang (2008) exposed either high or low-condition females to sperm containing damaged DNA and then assessed the frequency of lethal mutations on paternally derived X chromosomes transmitted by these females. The rate of lethal mutations transmitted by low-condition females was 30% greater than that of high-condition females, indicating reduced repair capacity of low-condition females (Agrawal and Wang, 2008).

In many taxa of invertebrates and vertebrates, postfertilization selection includes the cannibalism/killing of eggs and/or newborns by siblings or parents (Polis, 1981; Simmons, 1988; 1997; Smith and Reay, 1991; Elgar and Crespi, 1992; Mock and Parker, 1997; 1998; Hausfater and Hrdy, 2008). Drosophila larvae often face high densities and competition for limited food in these ephemeral habitats (Atkinson, 1979; Nunney, 1990). Similar to the selection for sperm motility and speed, this type of selection is often based on hatching speed. As eggs hatch and the larvae develop, food is consumed and waste products produced. High quality individuals with elevated growth rates reduce the availability of food and contaminate the medium, making the environment a harsher place for slower-developing individuals (Bakker, 1961; Dawood and Strickberger, 1969). In aquatic and terrestrial invertebrates and vertebrates that rely on mass spawning and fertilizations, selective survival of high quality larvae and juveniles has been identified (Dobzhansky, 1947; Travis et al., 1985; Houde, 1987; 1997; Haag and Garton, 1995; Sogard, 1997; Searcy and Sponaugle, 2001; Bergenius et al., 2002, Vigliola and Meekan, 2002; Wilson and Meekan, 2002; Marshall et al., 2003; Takasuka et al., 2003; McCormick and Hoey, 2004; Hoey and McCormick, 2004; Marshall and Keough, 2007; Johnson et al., 2010). The exposure of propagules to natural selection is related to the concept of ‘selection arenas’ (Kozlowski and Stearns, 1989). The selection arena hypothesis—also called progeny choice hypothesis—has been proposed to explain the observation that many organisms produce more fertilized zygotes than they ever support in growth (Stearns, 1987b; Kozlowski and Stearns, 1989; Bruggeman et al., 2004). It states that overproduction of zygotes could be explained as part of a bet-hedging and quality control mechanism which operates in the following manner: An enlarged array of zygotes is created of which only a genetically superior subset will fully develop; zygotes with a low future fitness fail, while zygotes with a high future fitness thrive. In this way, parental energy for reproduction is invested in the most promising zygotes. Essential in the hypothesis is that the potential members of the next generation are tested at a crucial moment, namely just before substantial parental investment starts. This hypothesis further assumes that (1) zygotes are cheap to produce, (2) parental time, energy and/or risk are invested in the zygotes, (3) offspring vary in fitness, and (4) this fitness difference can be identified. Natural selection will favor such a selection arena only if the benefits of high-quality progeny outweigh the cost of overproduction. Especially if the initial cost of a zygote is low relative to the cost of an independent offspring, zygote overproduction pays off and will skyrocket (Kozlowski and Stearns, 1989). Bell and Koufopanou’s (1991) exposure of propagules to natural selection is related to the concept of ‘selection arenas’. Yet, Kozlowski and Stearns (1989) missed the point that gamete overproduction is much higher than overproduction of zygotes and, most importantly, that gamete overproduction is subjected to a stringent quality control during sexual selection cascades that act independently of natural selection. Thus, the bet-hedging and quality aspect of zygote/offspring overproduction was only discussed in relation to environmental variability and its natural selection pressures but not in terms of sexual reproduction, its evolutionary rationale, the mutagenic origin of the genetic variability and selection regimes that are unrelated to natural selection, i.e. sexual selection cascades. Due to this oversight, the theory did not allow to dismiss the flawed genetic theories of sexual reproduction (see chapter 18.1), and overall could not contribute much to the debate over the evolutionary rationale of sexual reproduction. In higher taxa with internal fertilizations the action of natural selection in ‘selection arenas’ has been largely replaced by sexual fertilization- and embryo-selection.

The complex life cycles of many organisms create the potential for interactions between conspecifics that vary greatly in size, age, and experience (e.g., insects, Cameron et al., 2007; amphibians, Eitam et al., 2005; aquatic invertebrates, McCauley and Murdoch, 1987; scleractinian corals, Edmunds and Elahi, 2007; marine algae, Schroeter et al., 1995; marine fishes, Bjørnstad et al., 1999; Webster, 2004; D’Alessandro et al., 2013). These inter-cohort competitions drive density-dependent juvenile selective bottlenecks. In fishes, juvenile growth and survival is determined by density-dependent competition with conspecific adults, with inter-cohort predation (i.e., cannibalism; Murdoch, 1994; Bjørnstad et al., 1999), and competition for limited resources, such as food or enemy-free space (Szabo, 2002; Samhouri et al., 2009). In marine fishes a high larval growth rate can ‘carry-over’ to affect fitness of subsequent developmental stages (Searcy and Sponaugle, 2001; Bergenius et al., 2002; Shima and Findlay, 2002; Wilson and Meekan, 2002; McCormick and Hoey, 2004; Holmes and McCormick, 2006; McCormick and Meekan, 2007; Samhouri et al., 2009; Smith and Shima, 2011). In some bird species, parents are rarely capable of rearing all chicks produced, some of which are actively or passively eliminated by either parents or earlier hatching siblings (Nelson, 1978; Stinson, 1979; Edwards and Collopy 1983; Drummond et al., 1986; Simmons, 1988; 1997; Anderson, 1989; Mock and Parker, 1997; 1998; Hausfater and Hrdy, 2008). Mauck et al. (2004) found a positive correlation between early hatching and early breeding success with longevity in a 40-year demographic study of Leach’s storm-petrel, Oceanodroma leucorhoa.

Selective embryo abortion is also suspected in mammals. Mitochondrial function may well represent a quality control system in the early embryo that will determine whether the embryo proceeds further into development or is quickly eliminated (Dumollard et al., 2007b; Lane, 2011b). Many DNA repair and damage response genes are expressed in early mammalian embryos. DNA repair in the newly fertilized preimplantation embryo is believed to rely entirely on the oocyte's machinery (mRNAs and proteins deposited and stored prior to ovulation) (Zheng et al., 2005; Jaroudi and SenGupta, 2007). Notably, elevated levels of NER, MMR and HR gene transcripts were detected in preimplantation rat embryos developing from DNA-damaged sperm (Harrouk et al., 2000). Embryos possessing greater cellular/molecular damage are consuming more nutrients, such as amino acids for repair processes while embryos with the highest developmental capacity are characterized by a low amino acid turnover (Houghton et al., 2002; Leese, 2002; Brison et al., 2004; Baumann et al., 2007; Sturmey et al., 2009). Low levels of some repair proteins, however, indicate that the embryo’s ability to repair DNA damage may be limited (Zheng et al., 2005). Mouse and human appear to prefer conceptuses incompatible with their own MHC (major histocompatibility complex) antigens (Bolis et al., 1985; Wedekind, 1994). This has been extensively investigated in Hutterite woman, who experience increased fetal losses when married to men with similar human leucocyte antigens (HLA) alleles (Ober, 1999; Ober et al., 1992). In coypu there seems to be selective abortion of entire litters with respect to quality and sex (Gosling, 1986). In female common shrews the large numbers of offspring per litter that often exceed the number that are reared to weaning may be related to patterns of dispersal and inbreeding in this species. Matings between genetically similar individuals occur frequently in natural populations, and females appear unable to avoid costs associated with inbreeding by selecting mates or their sperm on the basis of genetic compatibility. Production of more offspring than are weaned, coupled with multiple mating, may therefore be a strategy to promote sibling competition for maternal investment and hence selection of the most genetically fit young from mixed paternity litters (Stockley and Macdonald, 1998). Reduced litter sizes have been reported to result when male mammals are subjected to a variety of ionizing radiation or heat stress regimes (Jannes et al., 1998; Rockett et al., 2001; Zhu and Setchell, 2004; Fatehi et al., 2006; Yaeram et al., 2006; Paul et al., 2008a). These regimes did not prevent fertilization but did cause apoptosis in the two- to eight-cell stage thus preventing further development of the embryos (Zhu and Setchell, 2004; Fatehi et al., 2006). In humans, less than one–third of fertilized embryos have a chance of surviving up to a term delivery (Ruder et al., 2008). The sporadic miscarriage rate of recognized pregnancies in the general population is about 20 to 30% (Wilcox et al., 1988; Bulletti et al., 1996; Wang X et al., 2003) but based on the assumption that many pregnancies abort spontaneously with no clinical recognition (e.g. Wilcox et al., 1988; Ellish et al., 1996; Zinaman et al., 1996), only 50 to 60% of all conceptions advance beyond 20 weeks of gestation (Baird et al., 1986; Wilcox et al., 1988; Norwitz et al., 2001). Using current in vitro procedures, less than half of inseminated bovine and human oocytes reach the blastocyst stage (Hardy et al., 1989; Xu et al., 1992; Keskintepe and Brackett, 1996) and of these many do not implant or attach following embryo transfer. Genetic factors are responsible for most losses in clinically recognized pregnancies and most losses before preclinical recognition (Leigh Simpson, 1999). There is strong evidence for uterine selection against genetically disadvantaged embryos (Warburton et al., 1983; Stein et al., 1986; Neuhäuser and Krackow, 2007). In case of sperm genetic damage, a few embryos may reach the blastocyst stage, but embryo selection will ensure that most of them will abort before growing to term (Ahmadi and Ng, 1999; Sakkas et al., 1999; Morris ID et al., 2002). However, the incidence of certain birth defects — most notably Down's syndrome — rises with maternal age with strongly accelerating rate when mothers approach menopause (Penrose, 1934; Hook, 1981; Morris JK et al., 2002; 2005). Conventional explanations have focused on a rising production of defective zygotes; in contrast, an evolutionary approach suggests a relaxed maternal screen. Relaxed screening potentially explains the rising incidence of chromosomal abnormalities in live-births, the incidence of normal embryos in spontaneous abortions, and the incidence of spontaneous abortions with maternal age (Kloss and Nesse, 1992; Stein et al., 1986; Forbes, 1997; Neuhäuser and Krackow, 2007). Aneuploidy is a form of stress-inducible mutation in eukaryotes, capable of fuelling rapid phenotypic evolution as bet-hedging strategy (Chen G et al., 2012). Increased levels of cortisol are associated with a higher risk of very early pregnancy loss in humans (Nepomnaschy et al., 2006).

8.7 Sexual selection

Sexual selection is defined as selection arising from competition for access to reproductive opportunities, which can occur either through competition with members of the same sex for access to mates (intrasexual selection) or through mate choice by the opposite sex (intersexual selection) (Andersson, 1994; Borgia, 1979; Clutton-Brock, 2007). Male and female reproductive optima invariably differ because the potential reproductive rate of males almost always exceeds that of females: females are selected to maximize mate ‘quality’, while males can increase fitness through mate ‘quantity’ (Clutton-Brock and Vincent, 1991). A dynamic, sexually selected conflict therefore exists in which ‘competitive’ males are selected to override the preference tactics evolved by ‘choosy’ females (Gage et al., 2002). Darwin (1859) wrote “Amongst many animals, sexual selection will give its aid to ordinary selection, by assuring to the most vigorous and best adapted males the greatest number of offspring.” Sexual selection is that special variety to natural selection which is responsible for the evolution of traits that promote success in competition for mates. Theoretical evaluation of sexual selection suggests that the strength of selection is dependent on the amount of effort that is invested in access to mates (usually by males) versus taking part in other parental activities such as maturing eggs or rearing offspring (Sutherland, 1985; 1987). Mate choice is the process leading to the tendency of members of one sex to mate nonrandomly with respect to one or more varying traits in members of the other sex (Heisler et al., 1987). Precopulatory sexual selection in animals, in which males (generally) compete to mate with (generally) choosy females has generated traits that promote the transmission of one individual’s genes at the expense of another’s (Eberhard, 1996; Birkhead and Møller, 1998; Simmons, 2001; but see Bonduriansky, 2009). In many species, females are highly selective when it comes to mating (Darwin, 1871; Bateson, 1983; Andersson, 1994; Kokko et al., 2003). In some of these species, females are congruent in their mate preference for a particular male, while in other species females are incongruent in their preference, with each preferring a different male. Female mate preferences, like male ornaments, are condition dependent, high quality females showing the strongest mate preference (Jennions and Petrie, 1997; Hunt et al., 2005; Cotton et al., 2006).

Sexual selection theory distinguishes between two types of genetic benefits, additive and nonadditive effects, mediated by preferences for good and compatible genes, respectively. Good genes preferences should imply directional selection and mating skew among males, and thus reduced genetic diversity in the population. In contrast, compatible genes preferences should give balancing selection that retains genetic diversity. The least controversial models of female mate choice emerged from resource-based mating systems. In mating systems in which males provide direct benefits to the female or her offspring, such as food or shelter, the answer seems straightforward — females should prefer to mate with males that are able to provide more or better resources. The finding that, based on human mtDNA (transmitted only from mother to child) “The Time to the Most Recent Common Ancestor” is estimated to be twice as long as the one based on the non-recombining part of the Y chromosome (passed from father to son) indicates that in human ancient populations males faced a highly competitive context, while females were facing a much smaller female-female competition (Favre and Sornette, 2012). Females that chose a male based on direct benefits appear to have simultaneously chosen a male of higher genetic quality (Kokko et al., 2003; Hunt et al., 2004). However, there are many other mating systems (and perhaps most mating systems) in which females receive no resources from males (called nonresource-based mating systems), yet females still express a preference among males. For example, in some taxa, males display at fixed courtship, territories known as leks and these males provide only genes (i.e. sperm) to their mates. A striking characteristic of lek mating is the large variance in males assembled on any one area. On a typical vertebrate lek, as few as 5% of the males may account for as many as 70% of the matings (Bradbury et al., 1985; Höglund and Alatalo, 1995). This congruence appears paradoxical given that a female receives only genes from the male she selects (termed the ‘paradox of the lek’; reviewed in Kirkpatrick and Ryan, 1991; Tomkins et al., 2004).

Good-genes sexual selection, especially viability-based sexual selection, is important in a variety of animal species (Andersson, 1994; Kirkpatrick, 1996; Møller and Thornhill, 1998; Petrie and Kempenaers, 1998; Petrie et al., 1998; Wilkinson et al., 1998; Møller and Alatalo, 1999). The condition-dependence of male ornaments is vindicated by studies showing that the expression of traits, such as tail ornaments and combs in birds, and carotenoid pigmentation in fishes and birds, correlates with condition, particularly immunocompetence, and survival (von Schantz et al., 1989; Milinski and Bakker, 1990; Zuk et al., 1990; Hill, 1991; Houde and Torio, 1992; Nicoletto, 1993; Andersson, 1994; Møller, 1994; Thompson CW et al., 1997). The good-genes models are specifically supported by studies showing that female birds can increase offspring fitness by mating with more ornamented males without obtaining any direct benefits (Norris, 1993; Møller, 1994; Petrie, 1994; von Schantz et al., 1994; Hasselquist et al., 1996). Little genetic variance in fitness traits is expected, however, because directional selection tends to drive beneficial alleles to fixation (Fisher, 1930; Falconer, 1989; Roff, 1997). Nevertheless, evidence shows that genetic variance persists despite directional selection (Houle, 1992; Pomiankowski and Møller, 1995; Rowe and Houle, 1996) and genetic benefits and offspring fitness provided by female choice (Alatalo et al., 1998; Møller and Alatalo, 1999; von Schantz et al., 1999; Jennions et al., 2001; Byers and Waits, 2006; Simmons and Kotiaho, 2007; García-González and Simmons, 2011). A theoretical solution to the lek paradox has been proposed on the basis of two assumptions (Rowe and Houle, 1996): that traits are condition-dependent, and that condition (a summary of characteristics that reflect the general health and vigor of an individual) shows high genetic variance. Condition-dependence as evidenced by sexual ornamentation may be related to immunocompetence, stress resistance and oxidative stress susceptibility (von Schantz et al., 1999; Buchanan, 2000; Velando et al., 2008; Dowling and Simmons, 2009; Cotton, 2009; Monaghan et al., 2009; Helfenstein et al., 2010) and may be used for premating and postmating cryptic choice (Velando et al., 2008; Cotton, 2009). Various experimental approaches to manipulate or quantify the levels of oxidative stress/damage and ROS production, and then to examine the associations between these parameters and sexual trait expression and immunocompetence (Alonso-Alvarez et al., 2004b; Constantini et al., 2007; Hõrak et al., 2007; Kurtz et al., 2007; Pike et al., 2007; Torres and Velando, 2007) found evidence for ROS-mediated sexual selection (Dowling and Simmons, 2009). The condition capture model also predicts that over evolutionary time, the amount of genetic variation for sexual traits should increase as they become increasingly condition-dependent. Such an evolutionary association has been reported across species of stalk-eyed flies (Wilkinson and Taper, 1999). Ultimately, the theory relies on the assumption that condition is influenced by a large number of loci, resulting in a relatively high frequency of mutations in genes coding for condition (Rowe and Houle, 1996). Experiments showed that male courtship rate in the dung beetle Onthophagus taurus is a condition-dependent trait that is preferred by females. More importantly, male condition has high genetic variance and is genetically correlated with courtship rate (Kotiaho et al., 2001).

In addition to the female “good gene” mate choice (Møller and Alatalo, 1999; Kokko, 2001; Byers and Waits, 2006), the “compatible gene” mate choice has been proposed (Hamilton and Zuk, 1989; Zeh and Zeh, 1996; 1997; Brown, 1997; Tregenza and Wedell, 2000). Rather than providing good genes that improve offspring fitness traits, these males may be relatively more compatible with the choosing female’s genotype (Zeh and Zeh, 1996; Brown, 1997; Tregenza and Wedell, 2000; Mays and Hill, 2004; Kempenaers, 2007). While good genes effects are additive and broadly speaking are expressed independently of the maternal genome, compatibility effects are nonadditive and are determined by the interaction between the parental genotypes (Zeh and Zeh, 1996; 1997). This leads to a key difference between the good genes and compatibility explanations for female choice: while females will usually agree about which males carry good genes, the most compatible male will vary depending on the genotype of the choosing female. Compatibility does, however, not equal with dissimilarity. The relationship between genetic dissimilarity and offspring fitness is such that the highest level of fitness (compatibility) is achieved at intermediate levels of genetic dissimilarity (Bateson, 1983; Wan et al., 2013). Maximal genetic dissimilarity between parents might not yield the highest fitness of offspring (Wegner et al., 2004). The main consequence of mating between genetically dissimilar partners is increased heterozygosity of offspring (Brown, 1997; Griggio et al., 2011). Genetic heterozygosity can influence fitness via two main mechanisms: either because it avoids unmasking any deleterious recessive alleles carried by either parent (the dominance hypothesis), or because heterozygotes are inherently superior to homozygotes (overdominance: Charlesworth and Charlesworth, 1987). Particularly, disassortative mating preferences for genes of the major histocompatibility complex (MHC) have been observed in fish, reptiles, birds and mammals (Yamazaki et al., 1976; Potts et al., 1991; Brown, 1997; Penn and Potts, 1999; Landry et al., 2001; Aeschlimann et al., 2003; Freeman-Gallant et al., 2003; Olsson et al., 2003; Richardson et al., 2005; Milinski, 2006; Forsberg et al., 2007; Yamazaki and Beauchamp, 2007; Radwan et al., 2008; Agbali et al., 2010; Setchell et al., 2010; Griggio et al., 2011; Reichard et al., 2012). Whatever the underlying mechanism, homozygosity is frequently detrimental to reproductive success, having been shown to reduce hatching success or embryo survival, litter size, and survival of offspring in captive studies (Ralls et al., 1988; Pusey and Wolf, 1996; Bixler and Tang-Martinez, 2006). Accumulating data support the idea that genetic similarity among mating partners also has negative effects on offspring number and survival in the wild (Stockley et al., 1993; Coltman et al., 1998; Crnokrak and Roff, 1999; Keller and Waller, 2002; Slate and Pemberton, 2002). The detrimental effects of inbreeding may continue into adulthood, with inbred individuals suffering a reduction in survival (Jiménez et al., 1994; Keller et al., 1994; Coltman et al., 1999), ability to hold territories (Meagher et al., 2000), and reproductive success (Keller, 1998; Slate et al., 2000; Seddon et al., 2004).

Despite differences between the 2 models of mate choice for genetic quality, recent evidence suggests females might use both cues simultaneously, e.g. via both pre- and postcopulatory mechanisms, when choosing mates (Howard and Lively, 2004; Mays and Hill, 2004; Neff and Pitcher, 2005; Hoffman et al., 2007; Eizaguirre et al., 2009; Kawano et al., 2009) or alternate between the 2 cues (Roberts and Gosling, 2003; Puurtinen et al., 2005; Oh and Badyaev, 2006). Ultimately, optimal female choice and the strength of preference for genetic quality are likely influenced by the variability of heterozygosity and compatibility among potential mates and the relative costs of obtaining indirect benefits (Colegrave et al. 2002a; Roberts et al., 2006; Kempenaers, 2007). Several hypotheses propose that female birds benefit indirectly from extra-pair mating by enhancing the genetic quality of their offspring, through good genes or genetic compatibility effects (Jennions and Petrie, 2000; Kokko, 2001; Griffith et al., 2002; Neff and Pitcher, 2005). Supporting this idea, recent studies have identified a range of fitness-related traits for which extra-pair offspring are superior to their within-pair half-siblings (Hasselquist et al., 1996; Kempenaers et al., 1997; Sheldon et al., 1997; Johnsen et al., 2000; Charmantier et al., 2004; Schmoll et al., 2005; Freeman-Gallant et al., 2006; Garvin et al., 2006; Bouwman et al., 2007; O’Brien and Dawson, 2007; Dreiss et al., 2008; Fossøy et al., 2008; Losdat et al., 2011). A recent study (Gohli et al., 2013) found that more promiscuous species of passerine birds had higher nucleotide diversity at autosomal introns, but not at Z-chromosome introns. In more promiscuous species, major histocompatibility complex class IIB alleles had higher sequence diversity, and therefore should recognize a broader spectrum of pathogens. The results suggest that female promiscuity in passerine birds targets a multitude of autosomal genes for their nonadditive, compatibility benefits. Also, as immunity genes seem to be of particular importance, interspecific variation in female promiscuity among passerine birds may have arisen in response to the strength of pathogen-mediated selection (Gohli et al., 2013).

Sexual selection may not only act in pre-mating but also in post-mating choice of sperm both following internal and external fertilization (Thornhill, 1983; Thornhill and Alcock, 1983; Eberhard, 1996; Olsson et al., 1996; Newcomer et al., 1999; Reyer et al., 1999; Radwan, 2004; Fisher et al., 2006; Birkhead, 2010). In polyandry, cryptic, post-copulatory female choice might be more likely to generate ‘good gene’ or ‘compatible gene’ benefits than female choice of mates based on the expression of secondary sexual traits (Eberhard, 1996; Radwan, 2004; Fisher et al., 2006; Slatyer et al., 2012). Theoretical studies indicate that condition-dependent sexual selection can elevate population mean fitness and accelerate the rate of adaptation, making the feedback between natural and sexual selection a particularly potent force in changing environments (Proulx, 1999, 2001, 2002; Whitlock, 2000; Agrawal, 2001; Siller, 2001; Lorch et al., 2003; Fricke and Arnqvist, 2007; Clark SCA et al., 2012). For instance, sexual selection has been shown to be a key component in the process of speciation (Lande, 1981; West-Eberhard, 1983; Hoy et al., 1988; McMillan et al., 1997; Seehausen et al., 1997; Civetta and Singh, 1999; Dieckman and Doebeli, 1999; Higashi et al., 1999; Panhuis et al., 2001; Kocher, 2004; Boul et al., 2007; van Doorn et al., 2009; Maan and Seehausen, 2011).

However, in well-adapted populations, sexually successful males sired unfit daughters, indicating sexual and natural selection were in conflict. However, in populations containing an influx of maladaptive alleles, attractive males sired offspring of high fitness, suggesting that sexual selection reinforced natural selection (Long et al., 2012). Theoretical and experimental evaluations suggest that a reduction in fitness due to sexual conflict, however, may offset the fitness gain related to sexual selection (Holland, 2002; Arnqvist and Rowe, 2005; Pischedda and Chippindale, 2006; Rundle et al., 2006; Tregenza et al., 2006; Fricke and Arnqvist, 2007; Hall et al., 2008; Whitlock and Agrawal, 2009; Connallon et al., 2010; Hollis and Houle, 2011; Arbuthnott and Rundle, 2012; but see García-González and Simmons, 2011) that may also be modulated by the species’ sex chromosome system (Connallon, 2010). On the other hand, sexual selection may be opposed by antagonistic natural selection: antagonistic pleiotropy between male sexual and nonsexual fitness is likely to evolve, ultimately limiting male trait evolution (Hine et al., 2011). Thus, sexually antagonistic genes should weaken the potential indirect genetic benefits of sexual selection by reducing the fitness of opposite-sex progeny from high-fitness parents (Pischedda and Chippindale, 2006).

Using phylogenetic comparative data, indices of sexual selection were found positively related to altitudinal range and habitat range in taxa of birds and in agamid lizards. A plausible interpretation of this pattern was that sexual selection combines synergistically with natural selection, thereby increasing physiological tolerance or the propensity to adapt to novel environments (Badyaev and Ghalambor, 1998; Tobias and Seddon, 2009; Östman and Stuart-Fox, 2011). Moreover, sexual selection allowed a faster removal of deleterious genes from the population (Radwan, 2004; Sharp and Agrawal, 2008; Hollis et al., 2009; Whitlock and Agrawal, 2009). Female choice can actually increase genetic variability by supporting a higher mutation rate in sexually selected traits. Increasing the mutation rate will be selected against because of the resulting decline in mean fitness. However, it also increases the probability of rare beneficial mutations arising, and mating skew caused by female preferences for male subjects carrying those beneficials with few deleterious mutations (‘good genes’) can lead to a mutation rate above that expected under natural selection (Petrie and Roberts, 2007). From this perspective, the interaction between sexual selection and population mean fitness is positive, potentially leading to broader ecological niches (Badyaev and Ghalambor, 1998; Proulx, 2002; Tobias and Seddon, 2009), accelerated rates of adaptation (Lorch et al., 2003; Fricke and Arnqvist, 2007), and a reduced likelihood of extinction (Whitlock, 2000). Finally, sexual conflict due to coevolution between the sexes may work to elicit a rapid evolution of reproductive barriers and increased speciation rates (Parker and Partridge, 1998; Arnqvist et al., 2000; Gavrilets, 2000; Martin and Hosken, 2003; McPeek and Gavrilets, 2006) (but see Gage et al., 2002 for contrasting evidence)

9. Coevolutionary dynamics of the germ-soma conflict


Natural selection may be said to be biased in favor of youth over old age whenever a conflict of interests arises.

George C Williams (1957)

Summary

From various biological levels of conflict it has become obvious that conflict breeds antagonistic coevolution that can be described by ‘Red Queen’ dynamics. Self-reinforcing adaptation/counter-adaptation chain reactions may lead to both a higher mutational robustness and higher evolvability of the genetic network. Conflicts may arise from competition among dividing cell lineages for opportunities to propagate or for access to the germline. Conflicts can also arise among the products of meiosis and between hosts and endosymbionts. A central conflict that has to be mediated in multicellular organisms with a germline-soma division revolves around the issue of resource allocation for both reproduction and somatic tissue maintenance and resolution of the transgenerational ‘tragedy of commons’ dilemma. The fundamental conflict between germline cells and soma over the transgenerational access to and control over limited resources resulting in somatic aging and death has been discussed recently (Heininger, 2002; 2012).

Evolutionary conflict arises in interactions when the individual optima of interactants cannot be satisfied simultaneously (Parker, 2006). This disparity between optima has been labeled the “battleground” of evolutionary conflict (Godfray, 1995; Cant, 2006; 2012). Invariably, access to and control over limited resources such as food, mates, reproduction, territories, are sources of conflict (Hirshleifer, 1988; Mock and Parker, 1998; Garfinkel and Skapaderas, 2007; Batchelor and Briffa, 2010; Heininger, 2012). Conflict arises at all levels of biological organization (e.g., between genes, cell organelles, cells, individuals, and groups) (Cosmides and Tooby, 1981; Alexander, 1987; Frank, 2003; Adams and Mesterton-Gibbons, 2003; Lachmann et al., 2003; Burt and Trivers, 2006; Garfinkel and Skapaderas, 2007; Reeve and Hölldobler, 2007; Queller and Strassman, 2009). From various biological levels of conflict (e.g. predator/prey, herbivore/plant, competitors, mating partners, host/pathogen, host/(endo)symbiont, mitonuclear, and intragenomic conflicts) it has become obvious that conflict breeds coevolution (Van Valen, 1973; Jaenike, 1978; Hamilton, 1980; Bell, 1982; Rice and Holland, 1997; Zeh and Zeh, 2000; Arnqvist and Rowe, 2002; 2005; Itoh et al., 2002; Rice and Chippindale, 2002; Eberhard, 2004; Hongoh et al., 2005; Lessells, 2006; Pal et al., 2007; Poulsen et al., 2007; Kikuchi et al., 2009; Lively, 2010; Paterson et al., 2010; Wilkins, 2010; Lane, 2011a; b). Mathematical models have shown that by competing with one another, coevolutive systems evolve at increased rates (Fernández et al., 1998). Both theoretical modeling and ecological evidence indicate that evolutionary dynamics are accelerated by coevolutionary patterns (Dieckmann and Law, 1996; Marrow et al., 1996; Sorenson and Payne, 2001; Rice and Chippindale, 2002; Good-Avila et al., 2006). Thus, coevolution speeds the rate of evolution (Lack, 1947b; Holt, 1977; Brown and Vincent, 1992; Travis, 1996; Lutzoni and Pagel, 1997; Rice and Holland, 1997; Brodie and Brodie, 1999; Thompson, 1999; Abrams, 2000; Doebeli and Dieckmann, 2000; Schluter, 2000a, b; Arnqvist and Rowe, 2002; Itoh et al., 2002; Thompson and Cunningham, 2002; Summers et al., 2003; Buckling and Rainey, 2002a; Eberhard, 2004; Vamosi, 2005; Nosil and Crespi, 2006; Pal et al., 2007; Raffel et al., 2008; Kikuchi et al., 2009; Paterson et al., 2010; Gómez and Buckling, 2011; Heininger, 2012). Also intraspecies competition may lead to coevolutionary processes (Roughgarden, 1972; Rosenzweig, 1978; Gibbons, 1979; Wilson and Turelli, 1986; Doebeli, 1996; Dieckmann and Doebeli, 1999; Drossel and McKane, 2000; Bolnick, 2001; Wichman et al., 2005; Svanbäck and Bolnick, 2007). Van Valen (1973) proposed the Red Queen Hypothesis in which each species is competing in a zero-sum game against others; each game is a dynamic equilibrium between competing species where “no species can ever win and new adversaries grinningly replace the losers.” The Red Queen is characterized by a coevolutionary chain reaction: adaptation by species A counter-adaptation by enemy species B counter-adaptation by species A ..., which can lead to a protracted period of coevolution (Ehrlich and Raven 1964; Vermeij, 1983; 1987; 1994; Arnqvist and Rowe, 2002). As a result, there is selection on each interacting partner to manipulate the trait towards its own optimum and resist such manipulation by the other partner (Lessells, 2006).

Conflicts may have both beneficial and disastrous consequences. The fitness-boosting value of coevolutionary systems has been demonstrated at various levels of biological organization (Spitze, 1991; Spitze et al., 1991; Clarke et al., 1994; Lynch and Spitze, 1994; Reznick et al., 2004a; Fisk et al., 2007; Pal et al., 2007; Paterson et al., 2010). Antagonistic coevolution is a cause of rapid and divergent evolution, and is likely to be a major driver of evolutionary change within and between species. When the bacterium Pseudomonas fluorescens and its viral parasite, phage Phi2 coevolved with each other, the rate of molecular evolution in the phage was far higher than when the phage evolved against a constant host genotype (Paterson et al., 2010). Guppies collected from stream environments lacking predators were found to be inferior in every aspect of their life history profile to those evolved in other, nearby sites with predators present (Reznick et al., 2004a). Fitness gains appear to accelerate under the challenge of moderate conflicts. Differences in the evolutionary interests of males and females, a well studied conflict of interest, may provide an important route to speciation (Chapman et al., 1995; Arnqvist and Rowe, 1995; 2005; Chapman and Partridge, 1996; Rice, 1996; 1998a; b; Alexander et al., 1997; Parker and Partridge, 1998; Holland and Rice, 1999; Arnqvist et al., 2000; Gavrilets, 2000; Martin and Hosken, 2003; Chapman, 2006) and, indeed, sexual conflict seems to be a key ‘‘engine of speciation’’ (Arnqvist et al., 2000; Martin and Hosken, 2003). When selection differs between the sexual conflict partners, a mutation beneficial to the one may be harmful to the other (sexually antagonistic) and can interfere with the other's adaptive evolution (Rice, 1984; 1992; 1998b; Chapman et al., 1995). Thus, conflicts may even result in extinction following disruptive selection (Rice, 1984; 1998b; Tanaka, 1996; Parker and Partridge, 1998; Arnqvist et al., 2000; Kisdi et al., 2001; Johansson, 2008; Heininger, 2012).

A parallel evolutionary chain reaction can occur between cell populations or non-allelic genes within the genome of a single organism. Mutation, fusion between different genotypes, and infection by symbionts are the key processes that can introduce genetic heterogeneity into the population of cells that constitutes a multicellular organism. The presence of genetically distinct replicating units in a multicellular organism can lead to several different kinds of conflict among cell lineages. The conflicts may arise between dividing cell lineages competing for opportunities to propagate or for access to the germline (Buss, 1982; 1985; 1987). Conflicts can also arise among the products of meiosis (Haig and Grafen, 1991), and between hosts and (endo)symbionts (Bell and Koufopanou, 1991; Hurst and Hamilton, 1992; Frank, 1996a; Itoh et al., 2002; Hongoh et al., 2005; Kikuchi et al., 2009). Thus, like host and parasite, the gene products from the conflicting partners are part of the evolving, biotic environment of one another, and they can potentially coevolve in an antagonistic or correlational fashion (Ehrlich and Raven, 1964), via a Red Queen process (Van Valen, 1973; 1974; Stenseth and Maynard Smith, 1984; Vermeij, 1994; Rice and Holland, 1997; Rice et al., 2005; Decaestecker et al., 2007; Schulte et al., 2010). A key characteristic of antagonistic coevolution is that it can lead to a self-reinforcing adaptation/counteradaptation chain reaction that leads to both a higher mutational robustness and higher evolvability of the genetic network (Wagner, 2005a; Lenski et al., 2006; Hintze and Adami, 2008; 2010; Yukilevich et al., 2008; Kartal and Ebenhöh, 2009; Whitacre and Bender, 2010; Fierst, 2011; Whitacre, 2011). That is, it can lead to recurrent, even perpetual, gene substitutions at antagonistically interacting loci and thereby continually drive genetic and phenotypic divergence among related species, isolated populations, cell populations and genetic loci. Accordingly, coevolving populations have significantly higher levels of heterozygosity and allelic diversity (Buckling and Rainy, 2002a; Duncan and Little, 2007; Duffy MA et al., 2008; Koskella and Lively, 2009; Bérénos et al., 2011). In general terms, antagonistic coevolution speeds the rate of evolution (Lutzoni and Pagel, 1997; Thompson and Cunningham, 2002; Harvell, 2004; Landry et al., 2005; Decaestecker et al., 2007; Paaby and Schmidt, 2008; Schulte et al., 2010; Fox et al., 2011; Bérénos et al., 2012) whenever allelic substitution at one locus selects for a new allele at the interacting locus, and vice versa, so that no stable equilibrium can be achieved, or is only achieved after many iterations. As a consequence, an increased rate of allelic substitution is a footprint left behind by the antagonistic coevolution process (Rice and Holland, 1997; Itoh et al., 2002; Landry et al., 2005; Pal et al., 2007; Paterson et al., 2010; Schulte et al., 2010; Fox et al., 2011).

On all levels of biological organization, competition for limited resources is one of the most pervasive motors of evolution (Darwin, 1859; MacArthur and Levins, 1964; MacArthur, 1970; Pianka, 1974b; Lawlor and Maynard Smith, 1976a; Brown, 1981; Winther, 2005; Kopp and Hermisson, 2006; Fisher and Hoekstra, 2010; Heininger, 2012). Within plants and animals different structures may compete during development for a shared and limited pool of resources to sustain growth and differentiation (Parzer and Moczek, 2008; Sadras and Denison, 2009). Also germ cells and soma compete for resources (Renault et al., 2004), a feature that becomes particularly evident during dietary restriction when reproductive activity is delayed in favor of somatic maintenance (Heininger, 2012).

The idea that parents and their young can be in evolutionary conflict over the amount of parental resources invested in individual offspring was raised by Hamilton (1964) and discussed in more detail by Trivers (1974). Both were the first to realize that natural selection can act in different ways on genes expressed in the parent and in the young. In most animals, the specification of germ cells occurs very early during ontogeny by asymmetric distribution of germ plasm. A multi-level selection theory predicted that a reproductive division of labor is required to evolve prior to subsequent functional specialization. The frequency distribution of division of labor and the sequence of its acquisition confirm that reproductive specialization evolves prior to functional specialization (Simpson, 2012). Thus, organisms without a segregated germline (e.g. plants, benthic aquatic animals), have much less differentiated tissues than unitary animals with a segregated germline. The segregation of the germline underlies the germ-soma conflict that, on the other hand, accelerated the coevolutionary dynamics of taxa.

Many mammals, birds, fishes and insects are found living at densities at the carrying capacity of their environments (Sibly et al., 2005; Brook and Bradshaw, 2006). Populations of vertebrate and invertebrate taxa are in general regulated by the production of adult individuals being a decreasing function of population density (Klomp, 1964; Tanner, 1966; Harrison, 1995; Myers et al., 1995; Kuang et al., 2003; Sibly et al., 2005; Bassar et al., 2010) which explains the relative stability of animal populations that do not increase at rates their fertility would allow. In a variety of insects, birds and mammals, breeding density or parent nutritional stress as proxy of high parent density affect the number and quality of offspring (Anderbrant et al., 1985; Jann and Ward, 1999; Lindström, 1999; Van de Pol et al., 2006; Descamps et al., 2008). In bark beetles, more than 20 offspring per female were produced at the lowest density but only 0.6 per female at the highest density. The offspring from the lowest density were about 50% heavier than those from the highest density and also the fat content increased with decreasing density. In the next generation, offspring from the highest density produced about half as many progeny as those from the lowest densities, showing an effect of density acting over more than one generation (Anderbrant et al., 1985). When vital rates are influenced by population density, population dynamics become nonlinear. The population is characterized not by an exponential growth rate, but by the existence and stability of equilibria, by bifurcations that occur when stability is lost, and by patterns of dynamics (cycles, quasicycles, chaos) that follow the bifurcations (Neubert and Caswell, 2000 Holland, 1995; Mitteldorf, 2010). As the quality of offspring affects the fitness of parents (see also chapter 3), species should have developed mechanisms of population control which prevent overutilization and destruction of their environment and therefore their own extinction (Christian, 1961).

Exploiters (parasites and predators) are thought to play a significant role in diversification, and ultimately speciation, of their hosts or prey (Darwin, 1859; Lack, 1947b; Schluter, 2000a; b). Exploiters may drive sympatric (within-population) diversification if there are a variety of exploiter-resistance strategies or fitness costs associated with exploiter resistance (Holt, 1977; Brown and Vincent, 1992; Abrams, 2000; Doebeli and Dieckmann, 2000; Schluter, 2000a). Exploiters may also drive allopatric (between-population) diversification by creating different selection pressures and increasing the rate of random divergence (Travis, 1996; Thompson, 1999; Buckling and Rainey, 2002a). Parasites increase diversification between populations through a combination of selection for resistance and chance: the morphotype in which the resistance mutation occurred (which varied between populations) swept to high frequencies and determined subsequent population evolutionary trajectories (Buckling and Rainey, 2002a).

The soma and germline cells are irreversibly dependent on each other by their mutual interests: reproduction of the soma is tied to germline cells and the germ cells require the soma to nurse and brood them and for mating (Heininger, 2012). Current theory suggests that mutualisms are best viewed as reciprocal exploitations that nonetheless provide net benefits to each partner (Nowak et al., 1994; Leigh and Rowell, 1995; Maynard Smith and Szathmáry, 1995; Herre and West, 1997; Doebeli and Knowlton, 1998; Herre et al., 1999; Sachs et al., 2004; Foster and Wenseleers, 2006; Leigh, 2010a; Jones et al., 2012). This view stresses the disruptive potential of conflicts of interests among the erstwhile partners. The potential for conflicts of interest to shape or destabilize mutualistic associations will depend on the extent to which the survival and reproductive interests of the symbiont align with those of the host (Herre et al., 1999; Jones et al., 2012). Leigh and Rowell (1995) and Leigh (1999, 2010) pointed out that the crucial aspect of the evolution of mutualism is whether partners have a sufficient common interest. As long as partners have a sufficient common interest, they should continue to cooperate, but as soon as conditions change to boost selfish interests, one (or both) of the partners may defect and a struggle rather than a harmonious relationship ensues (van Baalen and Jansen, 2001; Jones et al., 2012). Competition for resources is a common feature of mutualisms (Holland et al., 2005; Holland and DeAngelis, 2010; Jones et al., 2012). Simple models of consumer–resource interactions revealed multiple equilibria, including one for species coexistence and others for extinction of one or both species, indicating that species’ densities alone can determine the fate of interactions (Holland and DeAngelis, 2009; 2010).

The conflict between germline cells and the soma has a parasitic/endosymbiotic phenotype being fuelled by the transgenerational conflict over exploitation of limited resources. This conflict underlies the evolutionary programming of postreproductive aging and death of the soma (Heininger, 2012). Importantly, the germ-soma conflict depends on the segregation of germline cells from the soma and should arise in both sexually and asexually reproducing organisms. The twofold ‘cost of meiosis’ reduces parent–offspring relatedness from 1, in a female that reproduces parthenogenetically, to 0.5 in a sexually reproducing female (Williams, 1975; Uyenoyama, 1984; Rice, 2002). Relatedness may reduce conflict and hence attenuate coevolutionary dynamics (Lessells, 1999; Rankin, 2011). According to a theoretical model presented by Peters and Lively (1999; 2007), antagonistic coevolution can select for recombination. Thus, the germ-soma conflict per se may favor the evolution of sexual reproduction. In a bacterial and Drosophila coevolutionary model, deleterious mutations exacted higher costs than under non-coevolutionary conditions (Cooper et al., 2005; Buckling et al., 2006; Young et al., 2009) and increased the rate at which deleterious mutations were purged from the bacterial population (Buckling et al., 2006).

In plants and benthic marine animals that have somatic totipotent stem cells and may reproduce by agametic cloning, the adult body is itself a reproductive unit that increases its fitness as a function of genet size. Accordingly, plants and immobile animals often exhibit labile sex expression but turn to sexual reproduction under environmental challenges (Harvell and Grosberg, 1988; Romme et al., 1997; Korpelainen, 1998; van Kleunen et al., 2001). The germ-soma conflict and, by consequence, antagonistic coevolution, can be expected to be boosted in sexually reproducing organisms. Theoretical models identified two general properties of antagonistically selected loci (Connallon and Clark, 2012). First, antagonistic selection inflates heterozygosity and fitness variance across a broad parameter range—a result that applies to alleles maintained by balancing selection and by recurrent mutation. Second, effective population size and genetic drift profoundly affect the statistical frequency distributions of antagonistically selected alleles. The efficacy of antagonistic selection (i.e., its tendency to dominate over genetic drift) is extremely weak relative to classical models, such as directional selection and overdominance. Alleles meeting traditional criteria for strong selection (Nes »1, where Ne is the effective population size, and s is a selection coefficient for a given sex or fitness component) may nevertheless evolve as if neutral. The effects of mutation and demography may generate population differences in overall levels of antagonistic fitness variation, as well as molecular population genetic signatures of balancing selection (Connallon and Clark, 2012).

Biotic interactions with ‘Red Queen dynamics’ often fuel chronic correlational selection, which is strong enough to maintain adaptive genetic correlations. Correlational selection builds favorable genetic correlations through the formation of linkage disequilibrium (LD) at underlying loci governing the traits (Sinervo and Svensson, 2002). However, LD built up by correlational selection are expected to decay rapidly (ie, within a few generations) due to recombination and segregation (Falconer and Mackay, 1996; Futuyma, 1998). Nevertheless, if correlational selection is strong and chronic, substantial linkage disequilibrium can be maintained owing to a balance between recombination, segregation and selection (Hartl and Clark, 1997; Lynch and Walsh, 1998). Thus, correlational selection favors optimal trait combinations (i.e. a single fitness peak). However, correlational selection is broader than this, encompassing fitness ridges or even saddles, with regions in which parallel changes in both traits (i.e. both increase or both decrease) have parallel rather than antagonistic effects on fitness (Phillips and Arnold, 1989; Roff and Fairbairn, 2007). Correlational selection can generate a suite of combinations that have equal fitnesses rather than a single fitness peak, and this may help to maintain variation in fitness trade-offs (Roff and Fairbairn, 2007). Moreover, it may facilitate evolution in fitness landscapes with its potential peaks and valleys (Gokhale et al., 2009).

10. The redox regulation of gametogenesis: mutagenesis, epigenesis, canalization and gamete quality control


Summary

Reactive oxygen and nitrogen species are signaling molecules: speed and range of signaling, almost unlimited availability and close link to the energy status of the cell are their features. Oxidative/nitrosative stress is the final common pathway of responses to a variety of biotic and abiotic stressors. Gametogenesis is subject to waves of metabolic and oxidative stress that regulate developmental epimutagenesis during primordial germ cell reprogramming, mutagenesis through both oxidative DNA damage and modulation of DNA repair, double strand breaks that are requisite for recombination, canalization by capacitators, and finally apoptosis as ultimate gamete quality assurance. Epigenetic regulatory mechanisms (RNA interference, DNA methylation and histone modifications) are thought to have evolved as defense mechanisms against genomic invaders, such as viruses and transposable elements. The redox-dependent control of transposable element mobility, genetic stability and epigenetic modulation highlights the integrated transgenerational regulation of genotype and phenotype by redox balance.

Two systems are primarily responsible for general reduction-oxidation (redox) regulation, the thioredoxin (TRX) and glutaredoxin/glutathione (GRX/GSH) systems. They maintain the redox cellular homeostasis as well as regulate several cellular processes through a thiol–redox mechanism (Holmgren, 1995; Nakamura et al., 1997; Schafer and Buettner, 2001). Thiol-based redox mechanisms rely on the special properties of Cys residues, which can adopt 10 different sulfur oxidation states from +6 to -2 (the fully reduced state) (Giles et al., 2003; Jacob et al., 2003). Reactive oxygen and nitrogen species are signaling molecules: speed and range of signaling, almost unlimited availability and close link to the energy status of the cell are their features. Environmental stress is the common denominator of the environmental modulation of a variety of developmental and life-history events (Heininger, 2001, 2012). Oxidative/nitrosative stress is the final common pathway of responses to a variety of biotic and abiotic stressors (Lindquist, 1986; Sanchez et al., 1992; Finkel and Holbrook, 2000; Heininger, 2001; Mittler, 2002; Mikkelsen and Wardman, 2003; Sørensen et al., 2003; Apel and Hirt, 2004; Ardanaz and Pagano, 2006; Rollo, 2007; Miller et al., 2008; Slos and Stoks, 2008; Jaspers and Kangasjärvi, 2010; Steinberg, 2012; Choudhury et al., 2013).

10.1 Mutagenesis

Mutagenesis has two components: generation and repair of DNA damage. Repair of DNA plays a major role in regulating mutant frequency (Klungland et al., 1999; Wang et al., 2006; Ikehata et al., 2007). The accumulation of DNA damage through misrepair or incomplete repair may determine the extent of mutagenesis (Kryston et al., 2011). Importantly, there exists a mutagenic synergism between the ability of RONS to generate DNA damage and their ability to inhibit the repair of this damage (Hu et al., 1995). Mutagens paradoxically favor loss of DNA repair, and the underlying logic comprises the metaphor ‘don’t stop for repairs in a war zone’ (Hu et al., 1995; Breivik, 2001). The key role of RONS in the generation of mutations is well established (Mello-Filho and Meneghini, 1984; Goldstein and Czapski, 1986; Imlay and Linn, 1987; 1988; Ziegler-Skylakakis and Andrae, 1987; Chevion, 1988; Imlay et al., 1988; Kazakov et al., 1988; Tachon, 1989; Blakely et al., 1990; Aruoma et al., 1991a; b; McBride et al., 1991; Juedes and Wogan, 1996; Oikawa and Kawanishi, 1996; Tamir and Tannenbaum, 1996; Rodriguez H et al., 1997; 1999; and see chapter 4.1). Genomic data mining revealed that germline caretakers that maintain DNA integrity/repair gene function are relatively dispensable for survival, and implied that milder (e.g., epimutational) male prezygotic repair defects could enhance sperm variation—and hence environmental adaptation and speciation—while sparing fertility (Zhao and Epstein, 2008). It was concluded that tumor suppressor genes that maintain DNA integrity and repair genes are general targets for epigenetically initiated adaptive evolution (Zhao and Epstein, 2008). One of the manifold relations between oxidative stress and mutagenesis converge in the base excision repair enzyme apurinic/apyrimidinic endonuclease 1 (APE1). APE1 is a key player in its role as a redox signaling factor and in the redox regulation of DNA repair (Luo M et al., 2010; Kelley et al., 2012). APE1 is the main apurinic/apyrimidinic endonuclease in eukaryotic cells, playing a central role in the DNA base excision repair (BER) pathway of all DNA lesions (uracil, alkylated and oxidized, and abasic sites), including single-strand breaks (Tell et al., 2009). APE1 is induced by oxidative stress (Grösch et al., 1998; Ramana et al., 1998) and has a pleiotropic role in controlling cellular response to oxidative stress including mitochondrial function and cell death signaling pathways. APE1 is directly responsible for the control of the intracellular ROS levels through inhibiting the ubiquitous small GTPase Rac1, the regulatory subunit of the NADPH oxidase system (Angkeow et al., 2002; Ozaki et al., 2002). APE1 protein upregulation is associated with an increase in both redox and AP endonuclease activity, followed by an increase in cell resistance toward oxidative stress and DNA damaging agents (Grösch et al., 1998; Ramana et al., 1998; Tell et al., 2005; McNeill and Wilson, 2007; Mitra et al., 2007). APE1 also coordinates recruitment of other DNA-repair proteins involved in BER through a complex network of direct protein–protein interactions and indirect interactions. In addition to its repair function in BER as an AP endonuclease, APE1, in its role as redox factor, regulates a number of transcription factors, including AP-1, Egr-1, tumor necrosis factor-alpha, NF-kappaB, p53, ATF/CREB, HIF-1alpha, STAT3, Myb, PEBP2, HLF, NF-Y, nuclear respiration factor 1 (NRF1) and others (Xanthoudakis and Curran, 1992; Xanthoudakis et al., 1992; Huang and Adamson, 1993; Mitomo et al., 1994; Huang et al., 1996; Nakshatri et al., 1996; Akamatsu et al., 1997; Jayaraman et al., 1997; Tell et al., 1998a; b; 2000; 2009; Ema et al., 1999; Gaiddon et al., 1999; Hirota et al., 1999; Ueno et al., 1999; Lando et al., 2000; Hall et al., 2001; Cao et al., 2002; Nishi et al., 2002; Ziel et al., 2004; Ando et al., 2008; Bhakat et al., 2009; Ray et al., 2010; Cardoso et al., 2012; Li et al., 2012). Thus, APE1 controls the redox status of several transcription factors that in turn regulate expression of APE1 (Grösch and Kaina, 1999; Fung et al., 2001; 2007; Haga et al., 2003; Pines et al., 2005; Zaky et al., 2008).

The biological activities of APE1 are sensitive to oxidative and nitrosative stress (Qu et al., 2007; Luo M et al., 2010; Kelley et al., 2012; Tang et al., 2012). APE1 DNA repair activity is redox dependent and can be completely abolished by higher ROS levels (Kelley and Parsons, 2001; Azam et al., 2008; Bhakat et al., 2009; Luo M et al., 2010). Paradoxically, an unbalanced increase in APE1 protein leads to genetic instability (Hofseth et al., 2003; Sossou et al., 2005). Overexpression of APE1 protein disrupts the repair of DNA mismatches and results in microsatellite instability (Chang et al., 2005). Elevated APE1 levels have been demonstrated in a variety of cancers and are typically associated with malignant transformation and progression, aggressive proliferation, genetic instability, increased resistance to therapeutic agents, and poor prognosis (Xu et al., 1997; Evans et al., 2000; Moore DH et al., 2000; Guo and Loeb, 2003; Hofseth et al., 2003; Wang D et al., 2004; Sossou et al., 2005; Yang S et al., 2007; Kim et al., 2013).

Noncytotoxic levels of H2O2 dramatically reduced the activities of the human DNA mismatch repair (MMR) system in repairing both single-base and insertion/deletion loop mismatches in a dose-dependent manner (Chang et al., 2002) and suppressed DNA MMR enzyme expression in rheumatoid arthritis (Lee SH et al., 2003). The main defense against the mutagenic effect of 8-oxoG is the BER pathway, which in eukaryotes is initiated by the OGG1 protein, a DNA glycosylase in the BER pathway that catalyzes the excision of 8-oxoG from DNA (Boiteux and Radicella, 2000). Acute oxidative stress results in nearly complete but reversible inactivation of the 8-oxoG DNA glycosylase activity of human OGG1 (Bravard et al., 2006; 2009). Chronic ROS exposure lowers the cellular activity of OGG1 in animal models (Potts et al., 2001; 2003) or in human cells (Youn et al., 2005) resulting from inhibition of gene expression at the transcriptional level. Likewise NO and RONS inhibit various DNA repair pathways (Laval and Wink, 1994; Wink and Laval, 1994; Graziewicz et al., 1996; Laval et al., 1997; Wachsman, 1997; Jaiswal et al., 2001a; b; Phoa and Epe, 2002; Sidorkina et al., 2003; Tang et al., 2012).

Hypoxia induces changes in the expression of several genes involved in DNA repair pathways. Chronic hypoxia induces the epigenetic downregulation of MMR (Jiricny 1998; Edwards et al., 2009). The expression of key genes within the MMR pathway and several critical mediators of HR, BRCA1, BRCA2, and RAD51 is downregulated by hypoxia (Mihaylova et al., 2003; Bindra et al., 2004; 2005a; b; 2007; Koshiji et al., 2005; Meng et al., 2005; Bindra and Glazer, 2007a; b; Bristow and Hill, 2008). BRCA1 is a caretaker gene that is responsible for repairing DNA, and it is able to upregulate several genes involved in the antioxidant response by controlling the activity of the transcription factors Nrf2 and NfkappaB (Benezra et al., 2003; Bae et al., 2004). HIF-1alpha is involved in the regulation of genetic instability at the nucleotide level by inhibiting the expression of a variety of DNA repair genes (Kim et al., 2001; Mihaylova et al., 2003; Bindra et al., 2005a; Koshiji et al., 2005; Meng et al., 2005; Shahrzad et al., 2005; To et al., 2005; 2006; Bindra and Glazer, 2007a; Huang LE et al., 2007; Crosby et al., 2009; Rezvani et al., 2010; Babar et al., 2011). A functional impairment of repair pathways, independent of gene expression, was also involved (Yuan et al., 2000; Bindra et al., 2004; 2005b). Defective MMR is associated with a mutator phenotype (Branch et al., 1993; 1995; Bhattacharyya et al., 1994; Eshelman and Markowitz, 1996). For example, in the absence of a MMR system targeted for oxidatively induced lesions, mutation rate is elevated 1,000 times relative to that of normal cells (Simpson, 1997). Since the same pathway is also responsible for repairing base:base mismatches, defective cells also experience large increases in the frequency of spontaneous transition and transversion mutations (Glaab et al., 1998; Umar et al., 1998; Aquilina and Bignami, 2001).

Downregulation of DNA repair under stress is adaptive at the cellular level. Embryonic stem (ES) cells deficient in DNA mismatch repair responded abnormally when exposed to oxidative DNA damage. ES cells derived from mice carrying either one or two disrupted Msh2 alleles displayed an increased survival following protracted exposures to low-level ionizing radiation as compared with wild-type ES cells. The increases in survival exhibited by ES cells deficient in DNA mismatch repair appeared to have resulted from a failure to efficiently execute cell death (apoptosis) in response to radiation exposure (DeWeese et al., 1998). Loss of MMR renders cells both resistant to the cytotoxicity of H2O2 and hypersensitive to the mutagenic effect of this oxidative stress (Lin et al., 2000). Hypoxia, along with its associated low pH, enriches for MMR-deficient cells in the surviving population and makes cells sensitive to some forms of hypoxia-induced genomic instability (Kondo et al., 2001). Hypoxic cells can acquire a mutator phenotype that consists of decreased DNA repair, an increased mutation rate and increased chromosomal instability, a phenomenon which has been well established in tumor progression (Huang LE et al., 2007; Bristow and Hill, 2008).

10.2 Recombination

As mentioned earlier, homologous recombination by itself does not generate sequence polymorphisms. However, recombination does prevent the reduction in variability caused by selective sweeps and sequential bottlenecks, thus increasing the polymorphism in a population (Suerbaum et al., 1998). Meiotic recombination is inherently hazardous because it is initiated by developmentally programmed DSBs at multiple sites in the genome (Keeney, 2007). DSB or free ends require oxidative stress for their generation (Henle and Linn, 1997; Aitken et al., 1998a; Lopes et al., 1998; Wang et al., 1998; Haber, 1999; Kemal Duru et al., 2000; Karanjawala et al., 2002; Sawyer DE et al., 2003; Barzilai and Yamamoto, 2004). ROS-induced DNA damage is recombinogenic (Brennan and Schiestl, 1998). p53 is activated by multiple cellular stress signals and oxidative stress (Lu and Lane, 1993; Wang and Ohnishi, 1997; Vousden and Lu, 2002; Murray-Zmijewski et al., 2006; 2008; Berns, 2010; Hölzel et al., 2010; Marchenko et al., 2010; Lu et al., 2011) and is involved in the control of meiotic recombination (Rotter et al., 1993; Sjöblom and Lähdetie, 1996; Stürzbecher et al., 1996; Habu et al., 2004; Di Giacomo et al., 2005; Ghafari et al., 2009; Lu et al., 2010). In many, if not most eukaryotes, crossing-over recombination can also occur during mitosis (Pontecorvo and Käfer, 1958). Mitotic recombination during spermatogenesis was observed both in Drosophila (Becker, 1974; Dewees, 1982; McKee, 2004; 2009), amphibians (Yamamoto et al., 1999) and mammals (Kelus and Steinberg, 1991; Högstrand and Böhme, 1997; Leeflang et al., 1999; Hsia et al., 2003). Likewise, mitotic recombination occurs in oogenesis of vertebrates and invertebrates (Wieschaus and Szabad, 1979; Busson et al., 1983; Perrimon et al., 1984; Hsia et al., 2003). Single-strand DNA (ssDNA) lesions, double-strand breaks (DSB) or free ends are prerequisites for mitotic and meiotic recombination (Szostak et al., 1983; McGill et al., 1993; Pâques and Haber, 1999; Cromie et al., 2001; Bleuyard et al., 2006). During the meiotic prophase, DSBs are induced in a controlled manner by the action of a specific topoisomerase II variant, Spo11. This protein is conserved from yeast to humans, and its expression is strongly induced in meiotic cells (Sun et al., 1989; Zenvirth et al., 1992; Bergerat et al., 1997; Keeney et al., 1997; 1999; Romanienko and Camerini-Otero, 1999; Celerin et al., 2000; Baarends et al., 2001; Keeney, 2007). Crossing over initiates with the appearance of DSBs in leptotene and is completed by late pachytene (Sun et al., 1989; Mahadevaiah et al., 2001; Zenvirth et al., 2003). The importance of forming DSBs to allow the process of recombination is illustrated by the observation that deletion of Spo11 results in infertility in both sexes. In males, loss of Spo11 results in failure of germ cells to progress beyond the zygotene stage of meiotic prophase and is associated with increased rates of germ cell apoptosis (Baudat et al., 2000; Romanienko and Camerini-Otero, 2000). Meiotic recombination, which promotes proper homologous chromosome segregation at the first meiotic division, normally occurs between allelic sequences on homologues (Keeney, 2007). However, recombination can also take place between non-allelic DNA segments that share high sequence identity. Such non-allelic homologous recombination can markedly alter genome architecture during gametogenesis by generating chromosomal rearrangements (Sasaki et al., 2010). In humans, DSBs are thought to resolve as crossover events in only 5–25% of the time (Jeffreys and May, 2004). Consequently, SNPs lying near the centre of a recombination hotspot (see chapter 12.2) are liable to be included within gene conversion tracts and will experience much higher effective recombination rates than predicted from crossover rates alone (The International HapMap Consortium, 2007). Meiotic gene conversion has an important role in allele diversification and in the homogenization of gene and other repeat DNA sequence families (Clarke et al., 1982; Jinks-Robertson and Petes, 1985; Powers and Smithies, 1986; Ogasawara et al., 2001; Rozen et al., 2003), sometimes with pathological consequences (Collier et al., 1993; Watnick et al., 1998).

10.3 Epimutagenesis

It is now widely recognized that heritable changes in gene expression can occur without accompanying changes in DNA sequence. Epigenetics refers to a collection of mechanisms and phenomena that define the phenotype of a cell without affecting the genotype (Wolffe and Matzke, 1999; Wolffe and Guschin, 2000; Goldberg et al., 2007). In molecular terms, it represents a range of chromatin modifications including DNA methylation, histone modifications, non-coding RNAs, remodelling of nucleosomes and higher order chromatin reorganization that determine whether, where and when genes are expressed (Jirtle and Skinner, 2007; Mattick, 2012).

Genotoxic stressors have an effect on both the genome and the epigenome (Kovalchuk and Baulch, 2008). From fungi and plants to humans, oxidative stress-dependent cellular differentiation (Sohal et al., 1986; Allen, 1991; 1998; Zs.-Nagy, 1992; Blackstone, 1999; 2009; Heininger, 2001; 2012; Orzechowski et al., 2002; Foyer and Noctor, 2005; de Magalhães and Church, 2006; Gapper and Dolan, 2006; Pitzschke et al., 2006; Hitchler and Domann, 2007; Covarrubias et al., 2008; Nasution et al., 2008; Scott and Eaton, 2008; Owusu-Ansah and Banerjee, 2009; Hernández-García et al., 2010; Vincent and Crozatier, 2010; Sardina et al., 2012) is associated with epigenetic changes. Epigenetic profiles, including DNA methylation, histone modifications, and non-coding RNA-mediated regulatory events, are modifiable during cellular differentiation but are phylogenetically conserved to effect the heritable and long-lasting changes in gene expression (Wu and Sun, 2006; Hitchler and Domann, 2007; Sasaki and Matsui, 2008; Singh et al., 2009; Feng et al., 2010; Hu B et al., 2010; Iorio et al., 2010; Mahpatra et al., 2010; Ficz et al., 2011; Wu and Zhang, 2011; Briones and Muegge, 2012; Hu et al., 2012; Sabin et al., 2013). A major part of the epigenetic variation is triggered by (oxidative) stress or changes in the environment (Lertratanangkoon et al., 1997; Finnegan, 2002; Labra et al., 2002; Wada et al., 2004; Rapp and Wendel, 2005; Grant-Downton and Dickinson, 2006; Richards, 2006; Choi and Sano, 2007; Bossdorf et al., 2008; Boyko and Kovalchuk, 2008; Mason et al., 2008; Jablonka and Raz, 2009; Turner, 2009; Angers et al., 2010; Halfmann and Lindquist, 2010; Verhoeven et al., 2010a; Flatscher et al., 2012; Grativol et al., 2012). Moreover, if conditions return to their original state, spontaneous back-mutation of epialleles can restore original phenotypes [e.g., in position-effect variegation (Richards, 2006; Flatscher et al., 2012)]. The discovery of position–effect variegation (PEV) by H. J. Muller in 1930 provided the first description of a phenomenon with an underlying epigenetic basis (Wakimoto, 1998).

Epigenetics is closely linked to cellular bioenergetics (Xie et al., 2007; Naviaux, 2008; Smiraglia et al., 2008; Minocherhomji et al., 2012) and is, at least in part, regulated by mtDNA copy number as a component of retrograde signaling from mitochondria to the nucleus. Growth and replication of the nucleus are limited by mitochondrial energy production and thus calorie availability. The regulation of nuclear replication and gene expression by calorie availability is mediated by mitochondrial energetics. This is achieved by coupling of nDNA chromatin structure and function by modification via high energy intermediates: phosphorylation by ATP, acetylation by acetyl-Coenzyme A (ac-CoA), deacetylation by nicotinamide adenine dinucleotide (NAD+), and methylation by S-adenosyl-methionine (SAM). Numerous characterized epigenetic marks, including histone methylation, acetylation, and ADP-ribosylation, as well as DNA methylation, have direct linkages to central metabolism through critical redox intermediates such as NAD+, SAM, and 2-oxoglutarate (Wallace and Fan, 2010; Cyr and Domann, 2011). SAM is produced in the cytosol by the reaction L-methionine + ATP to give SAM + Pi + PPi. Thus, SAM links energy production to the methylation of lysines and arginines in proteins and cytosines in DNA (Wallace and Fan, 2010). SAM has both antioxidant (Evans et al., 1997; Caro and Cederbaum, 2004) and pro-oxidant (Wang and Frey, 2007) properties. A significant percentage of proteins across all organisms are enzymes that catalyze the redox-dependent transfer of a methyl group from the cofactor SAM to a substrate (Cheng and Blumenthal, 1999; Martin and McMillan, 2002; Katz et al., 2003; Schubert et al., 2003; Petrossian and Clarke, 2009). Members of this superfamily, the radical SAM enzymes, contain an iron-sulphur (Fe-S) cluster and use SAM to catalyse a variety of radical reactions (Wang and Frey, 2007). Fe–S clusters are used as cofactors by many redox proteins and radical SAM enzymes (Boyd et al, 2009), and are important for electron transfer reactions and gene regulation as they sense changes in redox status and oxidative stress (Beinert and Kiley, 1999; Fontecave, 2006; Outten, 2007; Kaur et al., 2010). Prevalent calories increase ATP, acetyl-CoA, SAM and NADH, causing the modification of histones, the opening of the chromatin, increased gene expression, and the stimulation of growth and reproduction. Reduced calorie levels deplete ATP, ac-CoA, SAM, and increase NAD+, reducing histone modification, compacting the chromatin, and reducing gene expression, growth and reproduction (Wallace and Fan, 2010).

10.3.1 DNA methylation

Almost 40 years ago, it was proposed that cytosine DNA methylation in eukaryotes could act as a stably inherited modification affecting gene regulation and cellular differentiation (Holliday and Pugh, 1975; Riggs, 1975). The percentage of methylated cytosines ranges from 0–3% in insects, 2% to 8% in mammals and birds, 10% in fish and amphibians to up to 50% in higher plants (Doerfler, 1983; Adams, 1996). Whereas DNA methylation in plants can occur at C bases in diverse sequence contexts, in mammals DNA methylation occurs almost exclusively in symmetric CpG (C followed by G) dinucleotides and is estimated to occur at ~70–80% of CpG dinucleotides throughout the genome (Ehrlich et al., 1982; Law and Jacobsen, 2010). DNA methyl transferases (DNMTs) fall under two categories: de novo and maintenance (Goll and Bestor, 2005). Patterns of DNA methylation are initially established by the de novo DNA methyltransferases DNMT3A and DNMT3B during the blastocyst stage of embryonic development (Okano et al., 1998b; 1999). These methyl marks are then faithfully maintained during cell divisions through the action of the maintenance methyltransferase, DNMT1, which has a preference for hemi-methylated DNA (duplex DNA in which only one of the two strands is methylated), copying pre-existing methylation patterns onto the newly synthesized DNA strands during DNA replication and repair (Bestor and Ingram, 1983; Bestor et al., 1988; Hermann et al., 2004; Mortusewicz  et al., 2005; Chen and Li, 2006; Wu and Zhang, 2010). Both the establishment and maintenance of DNA methylation patterns are crucial for development as mice deficient in DNMT3B or DNMT1 are embryonic lethal (Li et al., 1992; Okano et al., 1999) and DNMT3A-null mice die by 4 weeks of age (Okano et al., 1999). In general, the presence of DNA methylation, in and around the promoter regions of genes, is associated with gene silencing, and loss of methylation accompanies transcription (Bird, 2002). Exposure to ROS can alter DNA methylation profiles (Panayiotidis et al., 2004; Lim et al., 2008; Tunc and Tremellen, 2009; Donkena et al., 2010; Min et al., 2010; Quan et al., 2011; Ziech et al., 2011). DNA oxidative injury, like other types of DNA damage (alkylation of bases, abasic sites, photodimers, etc.), plays a causal role in the formation of DNA methylation patterns (Weitzman et al., 1994; Cerda and Weitzman, 1997; Wachsman, 1997; Dalton et al., 1999; Franco et al., 2008; Tunc and Tremellen, 2009). Oxidative stress can cause both DNA methylation and demethylation (Lertratanangkoon et al., 1997; Panayiotidis et al., 2004; Campos et al., 2007; Franco et al., 2008; Lim et al., 2008; Bhusari et al., 2010; Donkena et al., 2010; Min et al., 2010; Quan et al., 2011; Ziech et al., 2011; Bernal et al., 2013). High doses of ionizing radiation, a known inducer of oxidative stress, result in epigenetic modifications in adult mice (Tawa et al., 1998). They also induce genomic instability and the bystander effect, a phenomenon in which nonirradiated cells exhibit radiation damage even though they are not directly exposed (Koturbash et al., 2008; Kovalchuk and Baulch, 2008). The bystander effect is dependent on epigenetic signaling, since targeted disruption of Dnmt1 and Dnmt3a in cultured cells eliminates the transmission of genomic instability (Rugo et al., 2011). Hydroxyl radical-induced DNA lesions, such as 8-oxoguanine (8-oxoG) and 8-OHdG (Kuchino et al., 1987; Weitzman et al., 1994; Turk et al., 1995; Turk and Weitzman, 1995; Cerda and Weitzman, 1997), O6-methylguanine (Tan and Li, 1990; Hepburn et al., 1991) and single stranded DNA (Christman et al., 1995) have all been shown to contribute to decreased DNA methylation by means of interfering with the ability of DNA to function as a substrate for the DNMTs and thus resulting in global hypomethylation (Panayiotidis et al., 2004; Franco et al., 2008). Oxidative damage of parental strand guanines would permit normal copying of methylation patterns through maintenance methylation, while oxidative damage of guanines in the nascent (and unmethylated) strand DNA, one or two bases 3’ to the target cytosine, would inhibit such methylation resulting in differences in methylation rates as great as 13-fold (Turk et al., 1995). DNA methylation also appears to affect factors such as chromatin organization (Jones and Wolffe, 1999; Segal and Widom, 2009; Chodavarapu et al., 2010). Aberrant DNA methylation patterns induced by oxidative insult could therefore further alter oxidative susceptibility of regions of DNA as well as affect DNA repair activity (Evans and Cooke, 2004). DNA hypomethylation affects point mutation rates in mammalian cells (Chan et al., 2001) and elicits an increased rate of rearrangements and gene loss by mitotic recombination (Chen et al., 1998).

An additional modified base, 5-hydroxymethylcytosine (5hmC), has been found to be a normal component of mammalian DNA (Kriaucionis and Heintz, 2009; Tahiliani et al., 2009) and appears to be generated by oxidation of m5C in a reaction catalyzed by the ten eleven translocation (Tet)-family of enzymes (Tahiliani et al., 2009). Tet proteins can convert 5mC into 5hmC, 5-formylcytosine, and 5-carboxylcytosine through three consecutive oxidation reactions. Recent studies have raised the possibility that 5hmC is an intermediate during DNA demethylation and that the Tet family enzymes are critical for this process (Guo et al., 2011; He YF et al., 2011; Ito et al., 2011; Williams et al., 2011; Wossidlo et al., 2011; Wu and Zhang, 2011).

DNA methylation via SAM can also be modulated by mitochondrial function (Wallace and Fan, 2010). ROS can lead to de novo DNA methylation (Campos et al., 2007; Lim et al., 2008; Donkena et al., 2010; Min et al., 2010; Quan et al., 2011; Ziech et al., 2011), e.g. by recruiting histone deacetylase 1 and DNA methyltransferase 1 as well (Lim et al., 2008). The level and pattern of 5-mC are determined by both DNA methylation and demethylation processes (He XJ et al., 2011). Demethylation of DNA can be passive and/or active. Passive DNA demethylation occurs when maintenance methyltransferases are inactive during the cell cycle following DNA replication, which results in a retention of the unmethylated state of the newly synthesized strand. Active DNA demethylation involves one or more enzymes and can occur independently of DNA replication (Zhu, 2009).

These epigenetic modifications constitute a unique profile in each cell and define cellular identity by regulating gene expression. Importantly, DNA methylations are not gene-locus specific and have a substantial stochastic component (Silva et al., 1993, Ushijima et al., 2003; Reiss and Mager, 2007; Raj and van Oudenaarden, 2008; Huang, 2009; Mohn and Schübeler, 2009; Feinberg and Irizarry, 2010; Petronis, 2010). The degree of fidelity in epigenetic transmission is about three orders of magnitude lower than that of DNA sequence (an error rate of 1 in 106 and 1 in 103 for DNA sequences and DNA modification, respectively) (Ushijima et al., 2003; Laird et al., 2004; Riggs and Xiong, 2004; Genereux et al., 2005; Fu et al., 2010; Petronis, 2010). The cardinal signs of epigenetic effects on gene transcription are variable expression of a gene in a population of isogenic individuals (variable expressivity) and/or a mosaic pattern among cells of the same type within an individual (variegation) (Whitelaw and Martin, 2001). The term ‘‘metastable epialleles’’ was coined to describe alleles with more than one stable epigenetic state driving variable expressivity and variegation (Rakyan et al., 2002). At each new generation, the establishment of a metastable epiallele’s epigenetic state is probabilistic (Peaston and Whitelaw, 2006). Metastable epialleles are most often associated with retroelements and transgenesis, resulting in ectopic or aberrant transcription of nearby genes (Ekram et al., 2012; Faulk et al., 2013). Thus, in addition to genetic variation, epigenetic variation that is higher than genetic variation, represents another level of diversity (Petronis et al., 2003; Wong et al., 2005; Peaston and Whitelaw, 2006; Rakyan and Beck, 2006; Richards, 2006; 2011; Zhai et al., 2008; Lister et al., 2009; Feinberg and Irizarry, 2010; Tal et al., 2010; Verhoeven et al., 2010a; Bell and Spector, 2011; Massicotte et al., 2011; Schmitz et al., 2011; Ekram et al., 2012; Havecker et al., 2012; Faulk et al., 2013). Variation in epigenetic modifications is at least partly independent from variation in the DNA sequence (e.g. Cubas et al., 1999; Cervera et al., 2002; Riddle and Richards, 2002; Keyte et al., 2006; Shindo et al., 2006; Vaughn et al., 2007; Bossdorf et al., 2008). For instance, Cervera et al. (2002) and Vaughn et al. (2007) found large and consistent ecotypic variation of DNA methylation in A. thaliana that was not correlated with genetic variation. Keyte et al. (2006) explored DNA methylation polymorphism in 20 accessions of cotton and found that the levels of epigenetic variation greatly exceeded genetically based estimates of variation (Bossdorf et al., 2008).

10.3.1.1 CpG mutation bias

ROS react with 5-methylcytosine to oxidize the 5,6-double bond; the intermediate product, 5-methylcytosine glycol, then deaminates to form thymine glycol. Thymine glycol base pairs with A and results in a CT transition (Marnett and Plastaras, 2001). There is a large body of evidence implicating cytosine-5 DNA methylation in transition mutations at CpG dinucleotides in vertebrates (Gojobori et al., 1982; Wang et al., 1982; Li et al., 1984; Bulmer, 1986; Cooper and Krawczak, 1990; Sved and Bird, 1990; Laird and Jaenisch, 1996; Yang et al., 1996a; O’Neill and Finette, 1998; Chan et al., 2001). Methylated CpGs are mutational hotspots and undergo mutation at a higher rate than the 4 unmodified bases. Cytosine methylation followed by a spontaneous deamination event creating TpG or CpA dinucleotides induces 40–70% of all spontaneous somatic mutations of the multiple classes at CpG and CpNpG sites and flanking nucleotides (Razin and Riggs, 1980; Holliday and Grigg, 1993; Mazin, 2009; Cooper et al., 2010). The net result is that methyl-CpGs mutate at 10–50 times the rate of C in any other context (Coulondre et al., 1978; Bird 1980; Duncan and Miller, 1980; Razin and Riggs 1980; Bulmer, 1986; Sved and Bird, 1990; Walser and Furano, 2010), or of any other base (Hwang and Green, 2004). In both eukaryotes (Fryxell and Zuckerkandl, 2000; Galtier et al., 2001; Birdsell, 2002) and prokaryotes (Birdsell, 2002), mutation processes produce more AT mutations than GC mutations, which may, at least in part, be compensated by the GC-biased gene conversion system (Marais, 2003). With much higher efficiency than other types of mutational events, CpG deamination events can create new transcription factor-binding sites in promoters of human genes, transposons, and in genomic regions bound by key transcription factors, contributing to variability in gene regulation (Zemojtel et al., 2009; 2011). That in vertebrates the mutation of cytosine to thymine (CT) in the context of CpG dinucleotides has the highest rate among all base substitutions, is reflected in the ongoing genomic depletion of CpG dinucleotides and has led to a decrease in frequency of amino acids coded by CpG dinucleotides in organisms with CpG methylation (Sved and Bird, 1990; Jones PA et al., 1992; Schorderet and Gartler, 1992; Yang et al., 1996b; Pfeifer, 2006). Notably, in humans CpG dinucleotides occur at a frequency ~21% of that predicted by random chance (International Human Genome Consortium, 2001).

Fully 30–40% of all human germline point mutations are thought to be methylation-induced even though the CpG dinucleotide is under-represented (Jones PA et al., 1992). Thus, CpG transition mutations are responsible for approximately one-third of all human hereditary disease mutations (Cooper and Krawczak, 1990). CpG transition mutations represent the single most common type of somatic point mutation of the p53 gene in human cancer (Greenblatt et al., 1994; Hollstein et al., 1994; Denissenko et al., 1997; Hussain and Harris, 1999; Pfeifer, 2000). CpG transition mutations are the most common type of point mutation found in mutation assays in vivo and in vitro (Jackson-Grusby et al., 1997; O’Neill and Finette, 1998; Ikehata et al., 2000; Ono et al., 2000). Work with cytosine-5 methyltransferases has shown that the enzymes themselves can contribute to deamination of cytosines in the target recognition sequence e.g. under conditions involving a limiting supply of the methyl donor S-adenosylmethionine (Shen et al., 1992; Laird and Jaenisch, 1994; Bandaru et al., 1996; Zingg et al., 1996; Okano et al., 1998a; b; 1999). There is a repair mechanism which specifically recognises G•T mispairs, and replaces thymine with cytosine. However, this repair is not fully efficient, because the 5mCT transition mutation occurs about 10 times as frequently as other transitions (Holliday and Grigg, 1993). CpG deamination has been ascribed a role in silencing of transposons and induction of variation in regional methylation. CpG deamination events can create transcription factor–binding sites with much higher efficiency than other types of mutational events (Zemojtel et al., 2009; 2011).

Comparison of the human and chimpanzee genomes has shown that 14% of the single amino acid changes are due to the biased instability of CpG sequences, which can be subject to methylation and thence to mutations (Misawa et al., 2008). The methylation of CpGs is a major contributing factor to mutation in RB1, a gene in which allelic inactivation leads to the developmental tumor retinoblastoma (Mancini et al., 1997). Intriguingly, the CpG content is strongly correlated with a higher rate of neutral mutation at non-CpG sites (Walser et al., 2008; Walser and Furano, 2010), which suggests that CpGs play a role in influencing the mutation rate of DNA not containing CpG, perhaps by influencing the chromatin conformation surrounding the CpG and making it more accessible to other modifying processes. Furthermore, CpG content also appears to influence the type of mutation that occurs, with a higher ratio of transition-to transversion mutations observed in parallel with the non-CpG mutation rate (Walser and Furano, 2010).

However, there are local exceptions to the hypermutability of CpG sequences. Short sequences (~1kb) in which CpGs appear in high density in a mostly unmethylated form are termed CpG islands (CGIs) (Antequera and Bird, 1993). CGIs, though only representative of a fraction of all CpG dinucleotides, occur in the promoter sequences of a vast number of human genes (Illingworth and Bird, 2009; Illingworth et al., 2010). These islands are thought to persist based upon methylation status; since they remain largely unmethylated, minimal deamination takes place and the islands maintain their integrity. On the other hand, a subset of CGIs appear to have tissue-specific roles due to cell type–specific DNA methylation at CGIs during differentiation (Deaton et al., 2011; Deaton and Bird, 2011).

10.3.2 Histone modifications

Histone modifications including acetylation, methylation, ADP-ribosylation, SUMOylation, phosphorylation, ubiquitylation, and deamination are other redox-dependent key mechanism in transcriptional regulation, which are well conserved through a host of plant and animal species (Bird and Wolffe, 1999; Kornberg and Lorch, 1999; Cheung et al., 2000; Strahl and Allis, 2000; Jenuwein and Allis, 2001; Rice and Allis, 2001; Turner, 2002; Verdin et al., 2003; Bode and Dong, 2005; Causevic et al., 2006; Kondo, 2009; Sundar et al., 2010; Cyr and Domann, 2011). The ‘‘histone code’’ hypothesis has been formulated to account for the vast capacity of histones for transient information storage on top of DNA sequence (Strahl and Allis, 2000; Jenuwein and Allis, 2001; Musselman et al., 2012). Histone modifications have both active and repressive effects on chromatin function (Chen et al., 1999; Strahl et al., 2001; Wang H et al., 2001; Arney et al., 2002; Cowell et al., 2002; Erhardt et al., 2003; Reik et al., 2003; Kourmouli et al., 2004; Lepikhov and Walter, 2004; Liu et al., 2004; Yu MC et al., 2006; Berger, 2007; Klose and Zhang, 2007; Lin H et al., 2010; Musselman et al., 2012). For instance, a common mark associated with active chromatin is the trimethylation of histone 3 (H3) at lysine 4 (K4), or H3K4me3, which is often found at promoters of actively transcribed genes (Black et al., 2012). Conversely, marks associated with silenced heterochromatin include di- and trimethylated H3K9, as well as trimethylation of H3K27 (Black et al., 2012). Histone methylation patterns define the vast majority of mammalian recombination hotspots (Borde et al., 2009; Buard et al., 2009; Grey et al., 2011; Ségurel et al., 2011; Smagulova et al., 2011). It is generally assumed that chromatin condensation and gene repression are linked to the activity of histone deacetylases (HDAC), whereas chromatin relaxation and the promotion of gene expression involve the activity of histone acetyltransferases (HAT) (Davie and Spencer, 1999). Acetylation by HATs of specific lysine residues on the N-terminal tail of core histones results in uncoiling of the DNA and increased accessibility to transcription factor binding. In contrast, histone deacetylation by HDAC represses gene transcription by promoting DNA winding thereby limiting access to transcription factors. ROS are involved in histone acetylation and deacetylation and their balance (Rahman et al., 2002; 2004; Tomita et al., 2003; Ito et al., 2004; Adcock et al., 2005; Sundar et al., 2010). Accumulating evidence suggests that an epigenetic cross-talk, i.e. interplay between DNA methylation and histone modification, is involved in the process of gene transcription and gene silencing (Fuks, 2005; Viré et al., 2006; Rush et al., 2009; Margueron and Reinberg, 2011). The hierarchical order of events and dependencies leading to gene silencing, however, remains largely unknown. While some studies suggest that DNA methylation patterns guide histone modifications during gene silencing, other studies argue that DNA methylation takes its cues primarily from histone modification states (Causevic et al., 2006; Vaissière et al., 2008; Kondo, 2009). H3K9 methylation, for example, is a prerequisite for DNA methylation in Neurospora crassa and guides maintenance DNA methylation in Arabidopsis (Malone and Hannon, 2009; Law and Jacobsen, 2010).

Recent evidence links hypoxia-elicited oxidative stress and epigenetic regulation via histone methylation/demethylation and acetylation/deacetylation (Kato et al., 2004; Maltepe et al., 2005; Chen et al., 2006; Johnson and Barton, 2007; Beyer et al., 2008; Johnson et al., 2008; Wellmann et al., 2008; Pollard et al., 2008; Baccarelli and Bollati, 2009; Xia et al., 2009; Yang J et al., 2009; Krieg et al., 2010; Watson et al., 2010; Zhong et al., 2010). ROS have been shown to inhibit binding of methyl-CpG binding protein 2, a critical epigenetic regulator that recruits cytosine methyl transferases and histone deacetylases to DNA (Valinluck et al., 2004). It is becoming increasingly apparent that epigenetics plays a crucial role in the cellular response to hypoxia relayed by the hypoxia-induced transcription factor (HIF) family (Wellmann et al., 2008; Watson et al., 2010). This includes the role of epigenetics in both the stabilization and binding of HIF to its transcriptional targets, the role of histone demethylase enzymes following direct HIF transactivation, and the impact of hypoxic environments on global patterns of histone modifications and DNA methylation (Pollard et al., 2008; Watson et al., 2010). In addition to the traditional transcriptional regulation by HIF, recent studies have shown that epigenetic modulation such as histone methylation, acetylation, and DNA methylation can change the regulation of the response to hypoxia (Mimura et al., 2011).

The epigenetic, selectable variation might enable a lineage to adapt and “hold” the adaptation until genetic changes take over; thus, the heritable epigenetic variations in protein architecture pave the way for genetic adaptation (True et al., 2004; Sangster et al., 2004; Jablonka and Raz, 2009).

10.3.3 Noncoding RNA

Around 70 to 90% of the genome is transcribed in a variety of eukaryotes from fission yeast to humans (Carninci et al. 2005; Cheng et al. 2005; Willingham and Gingeras 2006; Birney et al., 2007; Wilhelm et al. 2008; Bühler, 2009) and plants (Chekanova et al., 2007; Matzke et al., 2009). Most of the transcripts correspond to nonprotein-coding (nc) RNAs of unknown function. Increasing evidence suggests, however, that the ncRNAs themselves and/or the act of transcription play key roles in establishing and maintaining the epigenetic architecture of eukaryotic genomes. In some cases, long ncRNAs are involved directly in recruiting chromatin factors (Nagano et al., 2008; Pandey et al., 2008; Lee, 2012), whereas in other instances they are processed by the RNAi machinery to generate short interfering RNAs that guide chromatin modifications to homologous regions of the genome (Kloc and Martienssen, 2008; Bühler, 2009). Small non-coding RNAs (sncRNAs), ranging from 19 to 30 nucleotides (nt) in length, constitute a large family of regulatory molecules with diverse functions in invertebrates, vertebrates, plants, and fungi (Bartel, 2004; Nakayashiki, 2005; Großhans and Filipowicz, 2008; Kawaji and Hayashizaki, 2008). The growing family of small RNAs, including microRNA (miRNA), small interfering RNA (siRNA), piwi-interacting RNA (piRNA), repeat-associated-siRNA (rasiRNA) and heterochromatic small RNA (hcRNA), forms the most abundant class of endogenous RNA in metazoans (Joly-Tonetti and Lamartine, 2012), but has also been characterized in plants, unicellular algae (Zhao et al., 2007), DNA viruses (Pfeffer et al., 2004) and, controversially, in retroviruses (Klase et al., 2007). All of the known sncRNA species have been found to be abundantly expressed in the testis (Aravin et al., 2006; Girard et al., 2006; Grivna et al., 2006a; b; Kim, 2006; Lau et al., 2006; Ro et al., 2007a; b; Kuramochi-Miyagawa et al., 2010; Linsen et al., 2010; Song R et al., 2011) and play critical roles in testicular development and spermatogenesis in mice (Deng and Lin, 2002; Hayashi et al., 2008; Kuramochi-Miyagawa et al., 2008; Ma et al., 2009; Frost et al., 2010; Korhonen et al., 2011; Romero et al., 2011; Pillai and Chuma, 2012). Although a variety of small RNAs have a host of functions in germline cells (e.g. Zhou X et al., 2010), here my focus is on miRNA and piRNA. MicroRNA (miRNA) was discovered nearly 20 years ago in C. elegans (Lee et al., 1993; Wightman et al., 1993). miRNAs are ~22-nt-long non-coding RNAs (Lagos-Quintana et al., 2001; Lau et al., 2001; Lee and Ambros, 2001) that regulate gene expression in eukaryotes and have been identified in various organisms including primates, rodents, birds, fish, worms, flies and viruses (Cullen, 2006; Kim and Nam, 2006; Ibanez-Ventoso et al., 2008). In metazoa, miRNAs target the RNA-induced silencing complex (RISC) usually to the 3´untranslated region of mRNA genes in a sequence-specific manner leading to mRNA cleavage (RNA interference). Or they target transcripts that are being translated, leading to inhibition of translation or changes in mRNA stability (Ambros, 2004; Bartel, 2004; Rana, 2007). miRNA-mediated translational repression is a reversible process in mammalian cells. The miRNA mode of action seems to be more of a fine-tuning of expression rather than degradation of mRNA, a mechanism by which RNA interference works (Engels and Hutvagner, 2006). Recently it was shown that miRNA have the potential to activate translation under certain conditions (Henke et al., 2008; Jopling et al., 2008; Orom et al., 2008) and have the ability to switch from translational repression to translational activation in cell-cycle-arrested cells (Vasudevan et al., 2007; 2008; Vasudevan and Steitz, 2007). miRNAs are currently thought to regulate the expression of most genes and consequently play critical roles in the coordination of a wide variety of processes, including differentiation, proliferation, metabolism, inflammation and cancerogenesis (Pasquinelli and Ruvkun, 2002; Bartel and Chen, 2004; Karp and Ambros, 2005; Miska, 2005; Kloosterman and Plasterk, 2006; Bushati and Cohen, 2007; Schetter et al., 2010).

Many miRNA genes were found to be evolutionarily conserved and this was thought to be a general characteristic of miRNAs. However, a number of nonconserved miRNAs have been recently discovered (Bentwich et al., 2005). miRNA controlled genes evolve under extremely high constraints and are more likely to undergo intense purifying selection than other genes (Chen and Rajewsky, 2006; Hertel et al., 2006; Sempere et al., 2006; Prochnik et al., 2007; Saunders et al., 2007; Heimberg et al., 2008; Nielsen et al., 2009). Importantly, after a positive selection-driven allele replacement is over, positive selection transforms into negative selection (Bazykin and Kondrashov, 2011). miRNAs are estimated to comprise 1%–5% of animal genes (Bartel, 2004; Bentwich et al., 2005; Berezikov et al., 2005), making them one of the most abundant classes of regulators: e.g., there are more than 1,000 miRNAs in humans (Bentwich et al., 2005; Berezikov et al., 2005; Xie et al., 2005; Rigoutsos et al, 2006) that are conserved throughout evolution with constitutive or spatially and temporally regulated expression. Target site predictions (Enright et al., 2003; Rajewsky, 2006) reveal that these human miRNAs have the potential to regulate thousands of human genes. Computational analyses suggest that a single transcript may be regulated by multiple miRNAs (Lindow and Gorodkin, 2007) and that each miRNA can target tens to hundreds of transcripts (Baek et al., 2008; Selbach et al., 2008), leading to the conclusion that miRNAs as a whole regulate the expression of at least 30% of human gene transcripts (Lewis et al., 2005; Rajewsky, 2006).

Recent evidence indicates that ncRNAs play an important role in both the generation and repair of DSBs (Shaham et al., 1999; Adamo et al., 2012; Adamo and Volpe, 2012). The DNA-damage response is a signaling pathway that originates from a DNA lesion and arrests cell proliferation. Evolutionarily conserved, ncRNA have been shown to function in the DNA-damage response (Francia et al., 2012; Liu and Lu, 2012; Tang and Ren, 2012; Kang HC et al., 2013). DSBs trigger production of ~21-nucleotide small RNAs from sequences flanking DSB sites in Arabidopsis, Drosophila and human cells (Michalik et al., 2012; Wei et al., 2012). Mutations in proteins involved in the biogenesis of DSB-induced small RNAs caused significant reduction in DSB repair efficiency (Wei et al., 2012). The small RNAs can repress homologous sequences in trans and may therefore–in addition to putative roles in repair–exert a quality control function by clearing potentially truncated messages from genes in the vicinity of the break (Michalik et al., 2012).

10.3.4 Epigenetic reprogramming

Epigenetic reprogramming is the process by which most genomic methylation patterns are erased and re-established in a sex-specific fashion. The methylation status of plant genomes was thought to be not reset each generation to the same degree as mammalian genomes are, and a considerable proportion of the methylation marks are stably transmitted across generations (Cervera et al., 2002; Riddle and Richards, 2005; Vaughn et al., 2007). However, more recent evidence suggests that epigenetic reprogramming also occurs in plant gametogenesis (Gehring et al., 2009; Hsieh et al., 2009; Slotkin et al., 2009; Jullien and Berger, 2010; Gutierrez-Marcos and Dickinson, 2012; Jullien et al., 2012). Importantly, epigenetic modifications, such as DNA methylation, are not entirely erased between generations and could underlie transgenerational epigenetic inheritance (Hadchouel et al., 1987; Reik, et al., 1987; Sapienza et al., 1987; Swain et al., 1987; Kearns et al., 2000; Sutherland et al., 2000; Lane et al., 2003). Two periods of mammalian epigenetic reprogramming have been detected: during primordial germ cell (PGC) migration (Hajkova et al., 2002; Yamazaki et al., 2003; Seki et al., 2007) and immediately after fertilization when the paternal genome is preferentially and actively demethylated (Monk et al., 1987; Mayer et al., 2000; Oswald et al., 2000; Abdalla et al., 2009; Okada et al., 2010; Smallwood and Kelsey, 2012). In the mouse male germline, the establishment of novel methylation marks for imprinted genes begins around day 15.5 post conceptionem, but is finished only after birth (Davis et al., 1999; 2000; Li JY et al., 2004). The SAM radical domain (see chapter 10.3) is involved in the active process of paternal genome demethylation (Okada et al., 2010). After fertilization, the one-cell zygote undergoes several cell divisions that ultimately lead to formation of the blastocyst. During this developmental period, maternally contributed DNMT1 is excluded from the nucleus (Carlson et al., 1992) and the maternal genome undergoes passive DNA demethylation (Li, 2002). A gradual loss of DNA methylation occurs with each cell division (Monk et al., 1987) in a replication-dependent manner (Howlett and Reik, 1991). During germ cell development, epigenetic reprogramming resets parent-of-origin based genomic imprints and restores totipotency to gametes. In PGCs, epigenetic reprogramming occurs on a genome-wide scale, which includes demethylation of DNA and remodeling of histones and their modifications (Reik et al., 2001; Surani, 2001; Hajkova et al., 2002; 2008; 2010; Li, 2002; Allegrucci et al., 2005; Morgan et al., 2005; Ohinata et al., 2006; Sasaki and Matsui, 2008; Feng et al., 2010; Popp et al., 2010; Guibert et al., 2012; Smallwood and Kelsey, 2012). Demethylation in PGCs is global and encompasses genic, intergenic and transposon sequences, with a median methylation level of 16.3% and 7.8% in male and female mouse PGCs, respectively (Popp et al., 2010). Methylation levels in fetus, embryonic stem cells and sperm are high (73.2–85%), whereas those in placenta are intermediate (42.3%) (Popp et al., 2010). PGCs at the endpoint of reprogramming have therefore attained a unique epigenetic state, with genome-wide demethylation of DNA, and loss of the repressive histone marks H3K9me2 and H2A/H4R3me2 together with H2AZ, as well as loss of the active histone mark H3K9ac (Seki et al., 2007; Hajkova et al., 2008; Popp et al., 2010; Guibert et al., 2012). DNA hypomethylation can impart genomic instability with elevated mutation rates, an increased rate of rearrangements and microsatellite slippage, activation of endogenous retroviral elements and gene loss by mitotic recombination (Jähner et al., 1982; Chen et al., 1998; Gaudet et al., 2003; Eden et al., 2003; Kim M et al., 2004; Wang and Shen, 2004; Howard et al., 2008). A link between DNA hypomethylation, chromosomal and genomic instability and carcinogenesis has been established (Chen et al., 1998; Saito et al., 2002; Fan et al., 2003; Kisseljova and Kisseljov, 2005; Shvachko, 2009; Kanai, 2010).

A recent study in C. elegans highlights the importance of reprogramming between generations (Daxinger and Whitelaw, 2012). Mutants that are null for the H3K4me2 demethylase spr-5 (the mammalian orthologue is KDM1A) exhibit progressive sterility over many generations (Katz et al., 2009). This sterility correlates with the misregulation of genes in spermatogenesis and the transgenerational accumulation of H3K4me2, suggesting that H3K4me2 needs to be cleared between generations.

Ectopic expression of the reprogramming factors Oct4, Sox2, Klf4, and c-Myc or Oct4, Sox2, and Klf4 allows the reprogramming of somatic cells to a pluripotent state (Takahashi and Yamanaka, 2006; Park et al., 2008; Hochedlinger and Plath, 2009; Yamanaka and Blau, 2010; Okita and Yamanaka, 2011). This reprogramming-induced pluripotency allows important inferences on the processes and states of PGC and embryonal reprogramming. There is a direct link between epigenetic reprogramming and increased DNA DSBs (González et al., 2013). p53 integrates several stress response pathways and coordinates the cellular response to a wide range of insults (Vazquez et al., 2008). A key role of p53 in PGC early mammalian embryo reprogramming has been identified (Kanatsu-Shinohara et al., 2004; Menendez et al., 2010). In contrast, wild type p53 operates in preventing reprogramming in somatic cells (Kawamura et al., 2009; Tapia and Schöler, 2010; Yi et al., 2012). Reprogramming to a pluripotent state has been shown to increase the number of cells with phosphorylated histone H2AX nuclear foci, one of the earliest cellular responses to DSBs (Huang and Darzynkiewicz, 2006; Kawamura et al., 2009; Marion et al, 2009; Müller et al., 2012; González et al., 2013). Demethylation of DNA involves potentially mutagenic DNA modifications that need to be processed through DNA repair mechanisms (Teperek-Tkacz et al., 2011; González et al., 2013). Thus, efficient reprogramming requires key homologous recombination genes, including Brca1, Brca2, and Rad51 (González et al., 2013) and possibly additional DNA repair pathways (Fong et al., 2011). Overall, reprogramming increases genomic instability and is mutagenic (Mayshar et al., 2010; Ramos-Mejia et al., 2010; Blasco et al., 2011; Gore et al., 2011; Hussein et al., 2011; Laurent et al., 2011; Pasi et al., 2011; Chen Z et al., 2012; Liang et al., 2013b; González et al., 2013). Importantly, single cell transcriptional profiling revealed an increased heterogeneity of reprogrammed pluripotent stem cells (Narsinh et al., 2011).

BORIS is a sperm cell-specific protein and cancer/testis gene (see chapter 7.3.2) with an 11-zinc-finger domain which interacts with demethylases to remove DNA methylation patterns by epigenetic reprogramming during the final round of mitosis in spermatogenesis (Klenova et al., 2002; Loukinov et al., 2002; Vatolin et al., 2005). Tet protein-mediated oxidative loss of 5mC and prevention of  unwanted DNA methyltransferase activity is central to PGC and early mammalian embryo epigenetic reprogramming (Wossidlo et al., 2011; Wu and Zhang, 2011; Xu et al., 2011). Studies in plants (Choi et al., 2002; Gong et al., 2002; Kapoor et al., 2005; Zhu, 2009), zebrafish (Rai et al., 2008) and mammalian cells (Barreto et al., 2007; Zhu, 2009; Cortellino et al., 2011) have suggested that active DNA demethylation can occur through various DNA repair mechanisms. DNA demethylation in the mouse PGCs is mechanistically linked to the appearance of single-stranded DNA breaks and the activation of the BER pathway (Hajkova et al., 2010). The most commonly recognized effect of global hypomethylation is to facilitate genomic instability (Gaudet et al., 2003; Kim M et al., 2004; Hoffmann and Schulz, 2005; Weber and Schübeler, 2007; Daskalos et al., 2009), probably mediated by hypomethylated genome-associated replication-independent DNA DSB error-prone repair (Pornthanakasem et al., 2008; Kongruttanachok et al., 2010). The importance of maintaining methylation on various types of repeats has been demonstrated in a number of methyltransferase-deficient systems. DNA methylation in mammals is associated with repeat stability: demethylation of minor satellites, subtelomeric satellites, microsatellites and selfish repeats appears to lead to increased recombination and mutagenesis and may result in destabilisation of the chromosome on which they reside (Hansen et al., 1999; Okano et al., 1999; Xu et al., 1999; Bourc’his and Bestor, 2004; Guo G et al., 2004; Kazazian, 2004; Kim M et al., 2004; Wang D et al., 2004; Gonzalo et al., 2006; Lees-Murdock and Walsh, 2008). During PGC reprogramming single-copy and imprinted sequences are demethylated, presumably by active demethylation (Kafri et al., 1992; Brandeis et al., 1993; Lee J et al., 2002; Hajkova et al., 2002; Popp et al., 2010). Some demethylation of the transposable Intracisternal A Particle (IAP) elements and Line1 elements was also reported in PGC (Walsh et al., 1998; Hajkova et al., 2002; Lane et al., 2003). In PGC both IAP and Line1 methylation was fairly high (74%, 65%), followed by considerable demethylation of Line1 (to 32% and 17% for combined male and female cells, respectively) and IAPs (40% male, 34% female) (Lane et al., 2003).

A high level of transposable element (TE) expression is usually deleterious for the organism, leading to double-strand breaks, disruption of protein-coding genes, mutations, chromosomal rearrangements, and an alteration in the transcription network (McClintock, 1951; Kidwell and Lisch, 2001;Gilbert et al., 2002; Symer et al., 2002; Deininger et al., 2003; Kazazian, 2004; Brookfield, 2005; Gasior et al., 2006; Hedges and Deininger, 2007; Goodier and Kazazian, 2008; Konkel and Batzer, 2010). Notably, the number of L1-induced DSBs is greater than the predicted numbers of successful insertions, suggesting a significant degree of inefficiency during the integration process (Gasior et al., 2006; Hedges and Deininger, 2007). Therefore, activity of mobile elements is thought to be under keen cellular control. Transcription of TEs is restricted in most differentiated tissues of animals and plants due to silencing directed by DNA methylation, histone modifications and RNA interference (Matzke et al., 2000; Miura et al., 2001; Slotkin and Martienssen, 2007; Law and Jacobsen, 2010). In germline cells, silencing of selfish elements is realized through short RNA species, called both repeat associated short interfering RNAs (rasiRNAs) (Aravin et al., 2003; 2004; Saito et al., 2006; Vagin et al., 2006; Gunawardane et al., 2007) and Piwi-interacting RNAs (piRNAs) (Brennecke et al., 2007). piRNAs of 24–32 nt in length are longer than 21–22 nt siRNAs derived from dsRNA or 21–23 nt endogenous miRNAs (Aravin et al., 2003; 2004). piRNAs play evolutionarily conserved roles in the regulation of TE in insects, mammals and zebrafish (Aravin et al., 2007a; Carmell et al., 2007; Houwing et al., 2007) and are accumulated specifically in the germline (Aravin et al., 2006; Girard et al., 2006; Lau et al., 2006; Klenov et al., 2007). Approximately 80% of the piRNAs in D. melanogaster are rasiRNAs with a sequence that corresponds to or is complementary to TEs (Aravin et al., 2003; Saito et al.. 2006; Brennecke et al., 2007; Gunawardane et al., 2007). piRNAs are specialized in the repression of mobile elements in the germline (Aravin et al., 2006; Girard et al., 2006; Grivna et al., 2006b; Saito et al., 2006; Vagin et al., 2006; Brennecke et al., 2007). A lack of piRNAs has been shown to be incompatible with normal spermatogenesis and male fertility (Carmell et al., 2007; Frost et al., 2010; Zheng et al., 2010; Watanabe et al., 2011). In plant meristems and gametogenesis, similar TE mobilization processes may be operative (Van Ex et al., 2011; Bucher et al., 2012; Martínez and Slotkin, 2012; Migicovsky and Kovalchuk, 2012).

Classical theory holds that parental-origin specific DNA methylated regions necessitate a process of epigenetic reprogramming during gamete development to ensure the successful development of future offspring (Radford et al., 2011). However, if sexual reproduction requires epigenetic reprogramming that not only serves to reset parental-origin-specific imprinting but also carries the increased risk of TE mobilization, this increased risk should be considered as an additional cost of sexual reproduction. Doesn’t evolution play here Russian roulette with a loaded and unlocked gun, particularly in cells that are thought to be vital for the transmission of stable genetic information?

10.4 Canalization

Living things must be able to dampen variable inputs (in nutrition, temperature, humidity, genetic background, etc.) to achieve the remarkable stability in the output (development, physiological responses, gene expression, etc.) (Wu et al., 2009). If organisms always produce the same phenotype, regardless of variation in genotype or environment, the relationship is described as canalization. Genetic canalization describes the insensitivity of a character to mutations, insensitivity to environmental factors is called environmental canalization (Wagner et al., 1997). The major fitness benefit driving the fixation of canalizing alleles derives from a reduction in environmental influences on phenotypic variation (Meiklejohn and Hartl, 2002). Canalization is therefore recognized as a property of organisms that influences their variability, or their propensity to vary (Wagner and Altenberg, 1996), a property that is selected for under fluctuating selection regimes (Kawecki, 2000). The terms ‘phenotypic plasticity’ (see chapter 12.5) and ‘canalization’ indicate whether environmental variation has a large or small effect on the phenotype. The heat shock protein 90 is considered the key mediator of canalization (Rutherford and Lindquist, 1998; Wagner et al., 1999).

A main feature of the heat shock response from bacteria to man is the vigorous but transient activation of a small number of specific genes previously either silent or active at low levels. New mRNAs are actively transcribed from these genes and translated into proteins which are collectively referred to as the heat shock proteins, or hsps (Schlesinger et al., 1982; Burdon, 1986). An intriguing feature is that the hsps have been highly conserved from yeast to plants and animals (Ingolia et al., 1982; Hackett and Lis, 1983; Farrelly and Finkelstein, 1984; Hunt and Morimoto, 1985; Rochester et al., 1986). A number of isoforms as members of several hsp families has been described and in eukaryotic cells the endomembrane systems [endoplasmic reticulum (ER), mitochondria, chloroplasts] harbor their own sets of proteins related to the hsp families (Nover and Scharf, 1997). The heat shock response is controlled by heat shock transcription factors (Hsfs) that act by binding to the highly conserved heat shock element (HSE) in the promoters of target genes. It is known that, besides heat, hsps and Hsfs are involved in cellular response to various forms of stress. Hsfs are also involved in different pathological conditions, cellular responses to oxidative stress, heavy metals, amino acid analogues and metabolic inhibitors, wounding, pathogen infection, and certain developmental and differentiation processes (Sorger and Pelham, 1988; Park and Craig, 1989; Jedlicka et al., 1997; Morimoto, 1998; Hahn et al., 2004; Swindell et al., 2007). An intimate relationship appears to exist between oxidative stress and the heat shock response (Jacquier-Sarlin and Polla, 1996; Liu and Thiele, 1996; McDuffee et al., 1997; Ahn and Thiele, 2003; Miller and Mittler, 2006). Hsfs may function as one of the hydrogen peroxide sensors in plants and animals (Manalo et al., 2002; Ahn and Thiele, 2003; Miller and Mittler, 2006; Volkov et al., 2006; Vandenbroucke et al., 2008).

Hsps act as molecular chaperones. Chaperones are proteins that assist other proteins in folding, and that can help refold misfolded proteins (Hartl, 1996; Young et al., 2001; Walter and Buchner, 2002). Protein misfolding can result from mutations in protein coding regions (Walter and Buchner, 2002; Peterson et al., 2010) or from environmental changes, such as heat stress, which can lead to protein denaturation (McClellan et al., 2007). Because proteins are involved in forming and maintaining every phenotypic trait, misfolded proteins often have detrimental effects on phenotypes (Hartl, 1996; Walter and Buchner, 2002). Proteins that can mitigate these effects can render organisms more robust against genetic or environmental perturbations. Thus, chaperones are one of several ways in which phenotypes can become robust to genetic, epigenetic and environmental change (Wagner, 2005a; Salathia and Queitsch, 2007; Chen and Wagner, 2012). Robustness to genetic and environmental change is often associated with one another (Queitsch et al., 2002; Milton et al., 2003; Hermisson and Wagner, 2004; Masel and Siegal, 2009; Jarosz and Lindquist, 2010). On evolutionary time scales, robustness to genetic change has an important consequence on the genetic constitution of a population: It allows mutations to accumulate that are not phenotypically visible, precisely because phenotypes are robust to such mutations. The resulting genetic variation is often also called cryptic variation (Gibson and Dworkin, 2004; Schlichting, 2008; McGuigan and Sgrò, 2009; Chen and Wagner, 2012). Such variation need not stay cryptic forever, however. It can become phenotypically visible in the presence of yet other mutations or after environmental change (Rutherford and Lindquist, 1998; Queitsch et al., 2002; Gibson and Dworkin, 2004; Sangster et al., 2008; Schlichting, 2008; Chen and Wagner, 2012). The resulting phenotypic change can be detrimental, but also beneficial, leading to new evolutionary adaptations (Maisnier-Patin et al., 2005; Sangster et al., 2008; Masel and Siegal, 2009; Jarosz and Lindquist, 2010). Cryptic genetic variation is likely to be an effective source of useful adaptations at a time of environmental change, relative to an equivalent source of variation that has not spent time in a hidden state (Masel, 2006).

The Hsp90 stress response protein is an ancient, abundant and nearly ubiquitous protein chaperone that interacts in an ATP-dependent system with more than 100 'client proteins' in the cell, most of which are involved in signaling pathways, including protein kinases, transcription factors and others, and either facilitates their stabilization and activation or directs them for proteasomal degradation. Both genetic (Cossins, 1998; Wagner et al., 1999; Marshall, 2002; Mitchell-Olds and Knight, 2002; Stearns, 2002; Velkov, 2002) and epigenetic (Pigliucci, 2003; Ruden et al., 2003; 2005; Rutherford and Henikoff, 2003; Sangster et al., 2003; Sollars et al., 2003) mechanisms likely explain the evolutionary capacitor function of Hsp90. By this means, Hsp90 displays a multifaceted ability to influence signal transduction, chromatin remodelling and epigenetic regulation, development and morphological evolution in nearly every organism and cell type examined (Rutherford and Zuker, 1994; Richter and Buchner, 2001; Nollen and Morimoto, 2002; Carey et al., 2006; Pearl and Prodromou, 2006; Salathia and Queitsch, 2007; Yeyati et al., 2007; Pearl et al., 2008). Hsp90 is extremely abundant – constituting ~1% of total protein under normal growth conditions (Welch and Feramisco, 1982) – and these levels are increased approximately twofold by environmental stress in yeast (Borkovich et al., 1989) and may even increase up to tenfold both in prokaryotes and in eukaryotes (Buchner, 1999). Complete loss of Hsp90 function is lethal, as multiple essential pathways are inactivated (Rutherford et al., 2007a). Hsp90 is constitutively expressed (Lindquist and Craig, 1988; Welch, 1990) and Hsp90 protein function is required at all times in eukaryotic cells (Borkovich et al., 1989; Cutforth and Rubin, 1994), but under stress conditions higher levels are achieved through induction of a heat shock response (Borkovich et al., 1989). By linking genetic variation to phenotypic variation via environmental stress, the Hsp90 protein folding reservoir might promote both stasis and change (Jarosz and Lindquist, 2010). The Hsp90 chaperone system alters relationships between genotypes and phenotypes under conditions of environmental stress (Rutherford and Lindquist, 1998; Queitsch et al., 2002; Cowen and Lindquist, 2005; Carey et al., 2006; Sangster et al., 2007; Sangster et al., 2008; Jarosz et al., 2010; Jarosz and Lindquist, 2010; Chen G et al., 2012) and, in so doing, provide at least two routes to the rapid evolution of new traits: (i) Acting as a potentiator, Hsp90’s folding reservoir allows individual genetic variants to immediately create new phenotypes; when the reservoir is compromised, the traits previously created by potentiated variants disappear. (ii) Acting as a capacitor, Hsp90’s excess chaperone capacity buffers the effects of other variants, storing them in a phenotypically silent form; when the Hsp90 reservoir is compromised, the effects of these variants are released, allowing them to create new traits. A variety of morphological abnormalities are expressed when Hsp90 is partially disabled in heterozygous Drosophila mutants or when developing flies or Arabidopsis seedlings are treated with sublethal doses of Hsp90-inhibitory drugs—conditions expected to mimic natural reductions of Hsp90 function by more extreme environmental stress (Rutherford and Lindquist, 1998; Nollen and Morimoto, 2002; Queitsch et al., 2002; Milton et al., 2003; 2006; Rutherford et al., 2007b; Sangster et al., 2007; Sangster et al., 2008). Modulation of Hsp90 activity not only is able to unmask cryptic, pre-existing variation but also to create the expression and assimilation of novel morphological phenotypes (Ruden et al., 2003; Sollars et al., 2003; Tariq et al., 2009; Mittelman and Wilson, 2010). Proteotoxic stress, caused by transient Hsp90 inhibition or heat shock, markedly increased chromosome instability to produce a yeast cell population with high karyotype diversity. Continued growth in the presence of an Hsp90 inhibitor resulted in the emergence of drug-resistant colonies with chromosome XV gain. Short-term exposure to Hsp90 stress potentiated fast adaptation to unrelated cytotoxic compounds by means of different aneuploid chromosome stoichiometries (Chen G et al., 2012). These findings demonstrate that aneuploidy is a form of stress-inducible mutation in eukaryotes, capable of fuelling rapid phenotypic evolution and drug resistance, and reveal a new role for Hsp90 in regulating the emergence of adaptive traits under stress. The loss of Hsp90 function under high stress may be due to its ATP-dependent functioning when ATP becomes limiting energetic stress-dependently (Csermely and Kahn, 1991; Csermely et al., 1993; Obermann et al., 1998; Panaretou et al., 1998; 2002; Buchner, 1999; Rutherford et al., 2007b), and possibly cochaperone-dependently (Stankiewicz and Mayer, 2012). Moreover, ROS-dependent degradation of Hsp90 protein may result in the loss of Hsp90 chaperone function, leading to client protein degradation (Pantano et al., 2003; Panopoulos et al., 2005; Shen et al., 2008; Beck et al., 2009; 2011; 2012), possibly by an ADP- and iron-dependent local generation of hydroxyl radicals through a Fenton-type reaction (Beck et al., 2012). Thus, canalization and phenotypic plasticity are two sides of the same coin. Under environmental stress the function of Hsp90 breaks down affecting the odds for a change of the redox-dependent canalization-phenotypic plasticity balance. HSP90 has also been shown to impact genomic stability and TE activity through regulation of the Piwi-interacting RNA pathway (Specchia et al., 2010; Gangaraju et al., 2011).

Numerical simulations of complex gene networks, as well as genome-scale expression data from yeast single-gene deletion strains illustrated that most, and perhaps all, genes reveal phenotypic variation when functionally compromised, and that the availability of loss-of-function mutations accelerates adaptation to a new optimum phenotype. However, this effect does not require the mutations to be conditional on the environment. Thus, there might exist a large class of evolutionary capacitors whose effects on phenotypic variation complement the systemic, environment-induced effects of Hsp90 (Hartman et al., 2001; Bergman and Siegal, 2003; Wagner, 2003; Kitani, 2004; Cooper et al., 2006; Suzuki and Nijhout, 2006; Wang GZ et al., 2011). For instance, microRNAs mediate another mechanism of canalization (Stark A et al., 2005; Hornstein and Shomron, 2006; Wu et al., 2009).

10.5 Apoptosis

Earlier, I presented compelling evidence that both compulsory measures and deceit succeeded to coerce cells into, and trapp them in, stress pathways with “dead-ends” (Heininger, 2001). Thus, apoptosis is no “altruistic suicide” but programmed cytocide in which social control and cellular competition turn signaling networks that evolved as survival pathways into “programmed cell death” (Heininger, 2001). Gamete quality control is exerted by germ cell life and death decision pathways. There is no doubt that ROS can be a double-edged sword: depending on the cellular context they exert signaling and executive functions in cellular survival and apoptosis decisions (Heininger, 2001; Martin and Barrett, 2002; Martindale and Holbrook, 2002; Fruehauf and Meyskens, 2007; Moreira da Silva et al., 2010; Maryanovich and Gross, 2013). Thus, regulation and execution of cell death is controlled by oxidative stress in both fungi, plants and animals (Buttke and Sandstrom, 1994; Jacobson, 1996; Tan et al., 1998; Jabs, 1999; Kannan and Jain, 2000; Simon et al., 2000; Jones, 2001; Ueda et al., 2002; Kern and Kehrer, 2005; Le Bras et al., 2005; Gechev et al., 2006; Orrenius, 2007; Matés et al., 2008; Trachootham et al., 2008; Circu and Aw, 2010; Doyle et al., 2010).

10.6 Transposable elements and epigenetics

In contrast to the relatively modest changes in the proteome through evolution, the amount of non-protein-coding DNA has increased dramatically and accounts for >98% of the human genome sequence (Taft et al., 2007). Transposable elements (TEs) are DNA sequences that can move (transpose) from one chromosomal location to another within the genome. While the genome of the nematode C. elegans has less than 5% (Kidwell and Lisch, 2000), and the Drosophila genome 7–8% (Smith et al. 2007), in mammals TEs account for nearly half of the genome (International Human Genome Consortium, 2001; Mouse Genome Sequencing Consortium, 2002), and in some plants they may constitute more than 90% of the genome (Leeton and Smyth, 1993; SanMiguel et al., 1996; Wessler, 2006; Tenaillon et al., 2010). TEs can be broadly categorized as retrotransposons (class I), which move via an RNA intermediate by a ‘‘copy and paste’’ mechanism, or DNA transposons (class II), which mobilize through a ‘‘cut-and-paste’’ mechanism (Slotkin and Martienssen, 2007; Goodier and Kazazian, 2008). As a consequence, class I retroelements generate an additional copy with every transposition, whereas the transposition of class II DNA transposons is conservative, where one site loses the transposon while another gains it. Long interspersed element-1, LINE-1 or L1 retrotransposons have successfully populated and modified eukaryotic genomes for hundreds of millions of years (Singer and Skowronski, 1985; Dombroski et al., 1991; Smit et al., 1995; Han and Boeke, 2005). L1, the most abundant class of retrotransposons in mammals, has approximately 500,000 copies in the haploid genome and about 100 and 3000 of them are estimated to be functional in humans and mice, respectively (Kazazian, 2004). The other major component of mammalian repetitive DNA is short interspersed nuclear elements (SINEs), comprising predominantly Alu and Alu-like elements (Singer, 1982). Such active TEs are typically expressed in the metazoan germline, wherein they generate newly transposed copies and pass them onto the next generation. L1 expression in the germline and cells that are closely associated with the germline has been previously reported (Branciforte and Martin, 1994; Trelogan and Martin, 1995; Ergun et al., 2004). DNA methylation is one important mechanism involved in the silencing of transposons in plant, mammalian, and fungal germlines (Yoder et al., 1997; Martienssen and Colot, 2001; Selker, 2004). Once thought to be purely selfish genomic entities (Doolittle and Sapienza, 1980; Orgel and Crick, 1980), TEs are now recognized to occupy a continuum of relationships, ranging from parasitic to mutualistic, with their host genomes (Zeh et al., 2009). For instance, a number of formerly selfish or parasitic element sequences have been exapted to provide regulatory and/or coding sequences that serve to increase the fitness of the host (Kidwell and Lisch, 2000; Kazazian, 2004; Feschotte, 2008; Huda et al., 2010). TEs can regulate host genes by serving as the targets of epigenetic histone modifications that spread into adjacent gene loci (Lippman et al., 2004; Mikkelsen et al., 2007).

In a variety of taxa, TE activity increases in response to extrinsic stress and oxidative stress (McClintock, 1984; Ratner et al., 1992; Arnault and Dufournel, 1994; Hirochika et al., 1996; Wessler, 1996; Grandbastien, 1998; Capy et al., 2000; Jiang N et al., 2003; Kikuchi et al., 2003; Lu and Ramos, 2003; Nakazaki et al., 2003; Jorgensen, 2004; Farkash et al., 2006; de la Vega et al., 2007; Slotkin and Martienssen, 2007; Bouvet et al., 2008; Cam et al., 2008; Desalvo et al., 2008; Oliver and Greene, 2009a; Zeh et al., 2009; Rebollo et al., 2010; 2012; Stoycheva et al., 2010; Belancio and Roy-Engel, 2011; Casacuberta and González, 2013). TEs serve as broad-spectrum mutator elements and are responsible for genetic variation in the host genome (Kidwell and Lisch, 1997; Miller and Capy, 2004; Kidwell, 2005). Thus, TEs have a major role in generating intraspecies genetic variability at the level of cytosine methylation and through insertions and non-homologous recombination events between elements, leading to various chromosomal rearrangements, duplications and deletions (Deininger et al., 2003; Sandovici et al., 2005; Boissinot et al., 2006; Rangwala et al., 2006; del Carmen Seleme et al., 2006; Wang and Dooner, 2006; Reiss and Mager, 2007; Slotkin and Martienssen, 2007). Like new mutations produced by any mutator mechanism, the majority of new TE-induced mutations are expected to be deleterious to their hosts (Kidwell and Lisch, 1997). TEs produce their mutagenic effects not simply on initial insertion into host DNA but may also produce mutations when they excise, leaving either no identifying sequence or only small ‘‘footprints’’ of their previous presence (Kidwell and Lisch, 1997). In E. coli, TEs increase mutational supply and occasionally generate variants with especially large phenotypic effects, e.g. by activating cryptic genes (Reynolds et al., 1981; Hall, 1999), but may be outcompeted by mismatch repair mutator alleles (Fehér et al., 2012). It has been estimated that 80% of the spontaneous mutations seen in Drosophila genetics result from TEs (Ashburner, 1992) that constitute 7–8% of the genome (Smith et al., 2007). TEs have repeatedly contributed regulatory and coding sequences to their hosts, leading to the emergence of new lineage-specific gene regulations and functions and were domesticated as drivers of genomic and biological diversity and evolution (Bowen and Jordan, 2002; 2007; Miller and Capy, 2004; Volff, 2006; Muotri et al., 2007; Piriyapongsa et al., 2007a; Böhne et al., 2008; Bourque, 2009; Sinzelle et al., 2009; Lankenau and Volff, 2009; Nakayashiki, 2011), including speciation (Rebollo et al., 2010). TEs have been implicated in the evolution of several key innovations, including acquired immunity in vertebrates (Kapitonov and Jurka, 2005) and placentation in mammals (Harris, 1998; Sekita et al., 2008). Thus, it is thought that the transposable element, MER20, contributed to the rewiring of gene regulatory networks and to the evolution of pregnancy in mammals (Lynch et al., 2011). Regulatory regions of heat shock genes are associated with constitutively uncondensed chromatin that not only facilitates rapid up-regulation in response to thermal stress, but also enables any TEs present in the region to escape transcriptional silencing, with few options for host countermeasures. In Drosophila, promoters of heat shock genes are indeed enriched for P elements whose effects contribute substantially to phenotypic variation (Chen B et al., 2007). In mammals, global epigenetic reprogramming occurs during preimplantation embryonic development and PGC development and coincides with dramatic increases in TE expression (Dupressoir and Heidmann, 1996; Loebel et al., 2004; Peaston et al., 2004; Svoboda et al., 2004; Taruscio and Mantovani, 2004; Maksakova et al., 2008; Leung and Lorincz, 2012). Developmentally regulated demethylation thus enables TEs to escape from transcriptional repression during critical stages and cause heritable transposition mutations. TE expression at these pre-meiotic stages can also yield clusters of progeny carrying identical transposition-induced mutations (Woodruff and Zhang, 2009). Such pre-meiotic clusters have major implications for transposon-mediated host evolution because they increase the fixation probability of a new mutant allele and the likelihood that the new mutation will precipitate reproductive isolation and speciation (Woodruff and Thompson, 2002; Rebollo et al., 2010). TE transposition has been revealed to take place during cellular differentiation in neuronal cells creating cellular diversity and neuronal plasticity (Muotri et al., 2005; 2007; 2009; Muotri and Gage, 2006; Vogel, 2011; Thomas et al., 2012) and involving both the DNA damage response (Coufal et al., 2011) and DNA methylation/demethylation machinery (Zhao et al., 2003; Muotri et al., 2010).

Because all known epigenetic pathways act on all TEs, it is likely that the specialized epigenetic regulation of regular host genes was co-opted from this ubiquitous need for the silencing of TEs and viruses. Growing evidence indicates that epigenetic regulation evolved to suppress TE (Bestor, 2003; Slotkin and Martienssen, 2007; Huda et al., 2010). It was argued that epigenetic regulatory mechanisms (RNA interference, DNA methylation and histone modifications) originally evolved as defense mechanisms against genomic invaders, such as viruses and TEs, and genome defense by suppressing TE mobilization (an immune system for the genome) appears to remain the primary function of this class of regulatory mechanisms (Barlow, 1993; Yoder et al., 1997; Plasterk, 2002; Vagin et al., 2006; Brennecke et al., 2007; Aravin et al., 2008; Obbard et al., 2009; Lisch and Bennetzen, 2011). Only later, gene silencing mechanisms were co-opted to serve the regulatory needs of the host organism (Yoder et al., 1997; Matzke et al., 2000; Lippman et al., 2004; McDonald et al., 2005; Piriyapongsa et al., 2007b; Piriyapongsa and Jordan, 2008; Obbard et al., 2009; Lisch and Bennetzen, 2011). However, physiological stress, induced by climate change or invasion of new habitats as well as oxidative stress disrupt epigenetic regulation and unleash TE (McClintock, 1984; Ratner et al., 1992; Arnault and Dufournel, 1994; Wessler, 1996; Grandbastien, 1998; Capy et al., 2000; Ikeda et al., 2001; Chen et al., 2003a; Daboussi and Capy, 2003; Lu and Ramos, 2003; Jorgensen, 2004; Farkash et al., 2006; McGraw and Brookfield, 2006; de la Vega et al., 2007; Bouvet et al., 2008; Cam et al., 2008; Desalvo et al., 2008; Perez-Hormaeche et al., 2008; Oliver and Greene, 2009a; Zeh et al., 2009; Rebollo et al., 2010; Stoycheva et al., 2010; Casacuberta and González, 2013).

Increasingly, non-coding RNAs have been identified to originate from within already characterized sequences such as genes and TEs (Matzke et al., 2000; Aravin et al., 2001; Sijen and Plasterk, 2003; Vastenhouw and Plasterk, 2004; Slotkin et al., 2005; Smalheiser and Torvik, 2005; 2006; Vagin et al., 2006; Brennecke et al., 2007; Piriyapongsa and Jordan, 2007; 2008; Piriyapongsa et al., 2007b; Conley et al., 2008; Zhang R et al., 2008; Kuang et al., 2009). TE-derived small RNAs have been traditionally treated as functionally distinct from gene-regulating small RNAs, such as miRNAs (Slotkin and Martienssen, 2007; Czech and Hannon, 2010; Saito and Siomi, 2010; McCue and Slotkin, 2012). TEs are major producers of endogenous small interfering RNAs (endo-siRNAs) in plants and animals and piRNAs in animals (Cox et al., 2000; Saito et al., 2006; Vagin et al., 2006; Brennecke et al., 2007; Yin and Lin, 2007). These two classes of small RNAs act to repress TE mRNA accumulation post-transcriptionally (Saito and Siomi, 2010) and, in some cases, to induce DNA methylation and repressive histone tail modifications at TE loci to maintain the element in a transcriptionally repressed heterochromatic state (Zilberman et al., 2003; Xie et al., 2004; Aravin and Bourc’his, 2008; Wang and Elgin, 2011). In contrast to TE-regulating small RNAs, gene-regulating small RNAs, such as miRNAs and the plant-specific tasiRNAs, are derived from single-copy or low-copy regions of the genome and are processed by different pathways compared with TEs (Montgomery et al., 2008; Felippes and Weigel, 2009; Czech and Hannon, 2010; McCue and Slotkin, 2012). However, two recent reports in Drosophila and Arabidopsis (Rouget et al., 2010; McCue et al., 2012; McCue and Slotkin, 2012) that experimentally revealed the direct regulation of a host gene mRNA by TE small RNAs have blurred the lines of this distinction. In both examples, epigenetically and developmentally regulated bursts in TE expression produced gene-regulating small RNAs (McCue et al., 2012). Recent data have also uncovered the less expected role of piRNAs in the regulation of gene expression through the control of mRNA stability (Rouget et al., 2010; Simonelig, 2011). Several lines of evidence indicate that piRNA-based gene regulation is likely to be widespread. For instance, a proportion of piRNAs is not produced from transposable element sequences, but from either intergenic sequences or 3´ UTR of cellular mRNAs (Senti and Brennecke, 2010; Siomi et al., 2011).

10.6.1 Role of microRNAs in gametogenetic DNA instability

Dicer, an RNaseIII endonuclease that is required for miRNA and small interfering RNA biogenesis, is required for primordial germ cell development and spermatogenesis (Hayashi et al., 2008). Selective ablation of Dicer in mouse Sertoli cells leads to infertility due to complete absence of spermatozoa and progressive testicular degeneration and Dicer and Dicer-dependent small RNAs have a continuous and cumulative effect on the process of spermatogenesis (Maatouk et al., 2008; Papaioannou et al., 2009; Korhonen et al., 2011; Romero et al., 2011; Liu D et al., 2012). Induction of miRNAs by hypoxia/HIF and RONS has been documented in a wide range of taxa (Hua et al., 2006; Donker et al., 2007; Hebert et al., 2007; Kulshreshtha et al., 2007; 2008; Fasanaro et al., 2008; Gaedicke et al., 2008; Kenneth and Rocha, 2008; Dekanty et al., 2010). On the other hand, miRNAs can also activate the generation of ROS (Chen et al., 2010; Favaro et al., 2010). Some studies have identified a miRNA signature associated with hypoxia in certain cell types (Kulshreshtha et al., 2007; 2008; Chan and Loscalzo, 2010; Loscalzo, 2010).

Hypoxia induces specific microRNAs collectively referred to as hypoxamirs (Chan and Loscalzo, 2010). A study investigating mechanistic manipulations of miRNAs in hypoxia has been conducted recently (Fasanaro et al., 2008). The study focused on one particular miRNA, miR-210, the master hypoxamir (Chan et al., 2012) that is consistently upregulated in all published studies, in both normal and transformed hypoxic cells (Kulshreshtha et al., 2007; Camps et al., 2008; Corn, 2008; Giannakakis et al., 2008; Ivan et al., 2008; Pulkkinen et al., 2008, Crosby et al., 2009; Pocock, 2011). miR-210 is involved in repressing mitochondrial respiration (Chan et al., 2009; Favaro et al., 2010) exaggerates production of undesired mitochondrial ROS (Favaro et al., 2010) and antagonizes DNA repair (Crosby et al., 2009). Intriguingly, VEGF that is central to oogenesis and follicular quality control (see chapters 7.2.4 and 8.2) is one of the most prominent miRNA targets, particularly of miR-210 (Hua et al., 2006; Foekens et al., 2008; Liu F et al., 2012; Yamasaki et al., 2012). miR-210 expression has been identified in gametogenesis of Drosophila, Xenopus, fishes and mammals (Grün et al., 2005; Madison-Villar and Michalak, 2011; Torley et al., 2011; Ambady et al., 2012; Bizuayehu et al., 2012; El Naby, 2012; Miles et al., 2012). In addition, several of the transcription factors activated by hypoxia including p53, NF-kappaB, as well as c-Myc can induce miRNAs (O’Donnell et al., 2005; Taganov et al., 2006; Chang et al., 2007; Raver-Shapira et al., 2007). miRNAs are up- or down-regulated in biotic and abiotic stress responses in plants (Phillips et al., 2007; Shukla et al., 2008; Sunkar, 2010) and animals (Babar et al., 2008; Leung and Sharp, 2010; Suzuki and Miyazono, 2010). Tumor suppressors p53, p63, and p73 that bind to conserved p53 response elements in promoter DNA, function as both positive and negative regulators of miRNAs maturation in response to DNA damage (Suzuki HI et al., 2009; Boominathan, 2010; Knouf et al., 2012). Various hypoxia- and oxidative stress-dependent miRNAs downregulate homologous recombination-mediated DNA repair, nucleotide excision repair and mismatch repair, enhance genomic instability and induce a mutator phenotype (Crosby et al., 2009; Lal et al., 2009; Valeri et al., 2010a; b; Moskwa et al., 2011; Tili et al., 2011).

miR-155, has emerged as a “master-regulator” of numerous biological processes, most notably those involved in immune function and cancer development (Rodriguez et al., 2007; Thai et al., 2007; Teng and Papavasiliou, 2009; Tili et al., 2009). miR-155 is an oncogenic miRNA product in humans or Bic in mice and is deregulated in a number of different cancers, most of which are of B cell origin. miR-155 downregulates MMR (Valeri et al., 2010a; Chang et al., 2011; Chang and Sharan, 2012; Yamamoto et al., 2012) and induces a mutator phenotype (Tili et al., 2011). Hypoxia and HIF-1 upregulate miR-155 suggesting that hypoxia-induced genomic instability may in part be mediated via miR-155 induction (Babar et al., 2011). Moreover, miR-155 acts to upregulate glycolysis in breast cancer cells (Jiang et al., 2012). miR-155 expression was significantly higher in sexually mature than immature porcine testes (Luo L et al., 2010). miR-743a that plays a role in the oxidative stress response in mitochondria (Shi and Gibson, 2011) was found up-regulated fourfold in mouse spermatogonial cell populations compared to gonocytes (McIver et al., 2012).

miR-18 that belongs to the Oncomir-1 cluster, is an oncogene that is intimately associated with the occurrence and progression of different types of cancer, including B-cell lymphomas, lung, breast, and colorectal carcinomas (Ota et al., 2004; Hayashita et al., 2005; Dews et al., 2006; Volinia et al., 2006; Mendell, 2008). miR-18a impairs the DNA damage response (DDR) through downregulation of ataxia telangiectasia mutated (ATM) kinase which functions as the primary sensor and transducer of DNA damage response signal (Song L et al., 2011). DDR related genes have been reported to play important roles in the maintenance of genomic stability and dysregulations of these genes were frequently found in many cancer types (Weinert and Lydall, 1993; Zhou and Elledge, 2000; Myung et al., 2001; Bartek and Lukas, 2003; Shiloh, 2003; Smith et al., 2010). miR-18 is highly abundant in testis, displaying distinct cell-type-specific expression during the epithelial cycle that constitutes spermatogenesis. Expression of HSF2 and of miR-18 exhibit an inverse correlation during spermatogenesis, indicating that, in germ cells, HSF2 is downregulated by miR-18 (Björk et al., 2010).

Several families of miRNA, dubbed “apoptomirs”, are involved in the regulation of apoptosis (Vecchione and Croce, 2010). miR-29 family members upregulate p53 levels and induce apoptosis in a p53-dependent manner (Park et al., 2009). Myeloid cell leukemia sequence 1 (MCL1), encodes an antiapoptotic Bcl2 family protein that promotes cell survival by interfering at an early stage in a cascade of events leading to the release of cytochrome c from mitochondria (Michels et al., 2005). MCL1 is also known to be the target of miR-29 members sensitizing cells to TNF-related apoptosis-inducing ligand (Mott et al., 2007). miR-29 members are thought to promote apoptosis through a mitochondrial pathway that involves p53, Mcl-1 and Bcl-2 (Xiong et al., 2010). Moreover, the miR-29 family influences de novo DNMTs, DNMT3a and DNMT3b expression directly and possibly maintenance DNMT1 indirectly (Fabbri et al., 2007; Takada et al., 2009; Filkowski et al., 2010; Meunier et al., 2012; Sandhu et al., 2012). DNMTs may also be involved directly and/or indirectly in the apoptotic process, because their alterations induce germ cell death (Doerksen et al., 2000). miR-29 downregulated expression of de novo methyltransferases, DNA methyltransferase 3a and 3b, could directly affect methylation patterns (Fabbri et al., 2007; Filkowski et al., 2010; Meunier et al., 2012) and induce global DNA hypomethylation (Garzon et al., 2009). miR-29 family members are expressed in the male and female germline (Takada et al., 2009; Filkowski et al., 2010; Meunier et al., 2012). In the testes of adut irradiation-exposed mice, miR-29a and miR-29b upregulation was paralleled by a significant downregulation of DNMT3a, profound hypomethylation of transposable LINE1 and SINE B2 sequences and genomic instability (Filkowski et al., 2010). Neonatal exposure to xenoestrogen induced a dose-dependent increase in miR-29a, miR-29b, and miR-29c expression. Increased miR-29 expression resulted in a decrease in DNMT1, DNMT3a, and DNMT3b, and a concomitant increase in transcript levels of DNA methylation target genes in correlation with their pattern of methylation (Meunier et al., 2012). Moreover, increased miR-29 expression decreased antiapoptotic Mcl-1 protein levels and induced adult germ cell apoptosis (Meunier et al., 2012)

11. Sexual mutagenesis-selection cascades (SMSC): there is more than meets the eye


Eldredge and Gould argued that both phases of punctuated equilibrium are inconsistent with natural selection as a single underlying cause….First, natural selection is posited as being effective only in fine-tuning organisms to small-scale changes in their environment. Second, some process other than natural selection must come into play to cause punctuations. The frequent association between punctuations and the appearance of new species suggests to Gould and Eldredge that this process, whatever it is, causes both phenomena…

Travis and Reznick, 2009

Summary

The Modern Synthesis and population genetics orthodoxy hold that (i) germline mutations are random carrying a high risk of being deleterious, (ii) cells within a multicellular individual are closely related, with little expected variation, so that selection between these cells can be ignored and therefore (iii) the gamete bottleneck (out of a huge number of gametes only a small number succeeds to become offspring) is stochastic, (iv) Muller’s ratchet operates in organisms with stochastic bottlenecks. Within this conceptual framework it can be inferred that Muller’s ratchet should operate in sexually reproducing organisms with narrow, stochastic gamete bottlenecks resulting in mutational meltdown. Sexual reproduction solved the “mutator-population size dilemma” by subjecting a huge population of germ cells and gametes to a selection cascade whose pre-selection principle is based on stress resistance (insensitivity to disturbance) and resilience (the rate of recovery after disturbance), competitiveness, and viability of the selected units. Thus, the molecular processes associated with sexual reproduction reveal a complexity that goes far beyond the simple sexual reproduction = meiotic recombination scenario.

The RNA virus quasispecies model can be viewed as a simple framework that contains all the basic ingredients of Darwinian evolution. Genomes are not independent entities due to mutational coupling among variants, and instead, the entire mutant distribution forms an organized cooperative structure which acts like (quasi) a single unit (species). The quasispecies target of natural selection has been proposed to be ‘the mutant distribution as a whole,’ as opposed to individual variants. The level of genetic diversity in a viral population, termed the quasispecies cloud size, is an intrinsic property of the quasispecies. The theoretical advantage of maintaining a diverse quasispecies is that, when the virus is shifted to a new environmental niche or selective regimen, a variant may already be present in the population which will be more fit in the new environment. When the mutation rate is adjusted upwards, the rate of adaptation increases. Quasispecies genetic variation can be reduced by purifying selection and by genetic bottlenecks. What is left by purifying selection and transmission bottlenecks to the action of natural selection is only the proverbial “tip of the iceberg” of quasispecies genetic variation. Muller’s ratchet operates in stochastic bottlenecks resulting in fitness losses and, eventually, in the extinction of the population. On the other hand, selective bottlenecks increase or at least do not attenuate the fitness of the transmitted viruses. The quasispecies concept is also applicable to other biological systems such as prions, bacteria and tumor cells.

Sexual mutagenesis and selection cascades and gamete bottlenecks are mutually interdependent and are finely tuned to each other and to the actions of natural selection. A general pattern can be recognized that allows to distinguish between three broad categories of organisms that differ with regard to body size, population size, size of the quasispecies cloud due to (epi)mutagenesis, stringency of gamete and offspring quality control and importance of natural selection. Sexual reproduction generated a process that, in its consequence, is a beneficial-mutation enrichment system. However, the quality check regime is far from optimal (Hartshorne et al., 2009). This may have several reasons: (i) the quality checks based on cellular stress-resistance cannot reflect the complexity of the environment and its manifold adaptive challenges; (ii) the system cannot check the proper functioning of differentiated cells in a multicellular organism with its division of labor; (iii) although the stochasticity of the gametic bottleneck is attenuated by the manifold quality checks, a residual stochastic, possibly fitness-eroding, element remains and (iv) the system relies on the bet-hedging strategy that must allow for a less than optimal quality check to create e.g. “hopeful monsters”.

Adherents to scientific paradigms self-censor their thinking within the boundaries of their beliefs. The following dogmas of the Modern Synthesis and population genetics made the proponents of these doctrines blind to the evolutionary potential of sexual reproduction: (i) natural selection is “the only direction-giving factor in evolution” (Mayr, 1980); (ii) selection within an individual can be ignored since all cells within a multicellular individual are recently derived from a common single-celled ancestor (the zygote or spore) and are hence closely related with little expected variation (Maynard Smith and Szathmáry, 1995, p. 244); (iii) mutations are accidental; (iv) due to the unfavorable ratio of deleterious and beneficial mutations, mutation rate should be as low as possible; and (v) as a result of the deleterious/beneficial mutation ratio, genetic variation largely depends on slightly deleterious and neutral mutations (Ohta, 1973; 1992; 2011).

So far, evolutionary theories of sexual reproduction focused on, and were limited to, the role of recombination. Recombination results in the exchange of genetic information between non-sister chromatids resulting in crossovers that are essential for the correct segregation of the chromosomes and by combining paternal and maternal alleles creates genetic diversity among individuals within a population (Baarends et al., 2001; Agrawal, 2006b). Real population sizes are generally significantly less than one billion individuals. This finding is important for discussions of the evolution of genetic recombination, especially when the expected numbers of double mutations in natural populations is considered (Ackerman et al., 2010). Maynard Smith (1968, 1978) pointed out that double mutants could be produced by mutation alone provided that the population size is large enough, making the generation of doubly favorable mutants by recombination unnecessary. But for this to happen in practice, the population size would have to be of the order of 1016, and for triple mutants population sizes of 1024 are needed (Ackerman et al., 2010). These numbers are based on the estimate of 10-5 favorable mutations per genome per generation (Perfeito et al., 2007), which translates into a per-gene rate of 10-8, assuming that there are at least 1000 genes per haploid genome. The projected population numbers are clearly unrealistic for the majority of eukaryotic species, especially multicellular eukaryotes (Ackerman et al., 2010). Although it is often impossible to measure population sizes directly, and impossible to measure historical effective population sizes, indirect measures can be obtained because the effective size N of a population is related to the genetic diversity that is maintained in the population (Kimura and Crow, 1964; Frankham, 1996). Such analyses of polymorphism data yield an estimated effective population size of approximately 10,000 for the modern human lineage (e.g., Takahata, 1993; Yu et al., 2001) and less than 100,000 for other primate lineages (Burgess and Yang, 2008). Traditional population genetics suggests that, given that the mutation rate for favorable alleles is orders of magnitude less than the reciprocal of the population size, new favorable mutants in a given gene are essentially unique and are, for all practical purposes nonrecurring. This problem is less acute for prokaryotes and single-celled eukaryotes that can have very large population sizes and that have rapid doubling times (20 min in Escherichia coli vs. 20 years in humans). Consequently, without recombination most of the favorable mutants that occur in finite populations—even very large finite populations—will eventually be lost. Barrick and Lenski, (2009) showed that, in asexual populations, some beneficial mutations can be lost through competition with other beneficial mutations.

Like for recombination, an important feature of “non-accidental” mutagenesis is the “necessity” of large population sizes to account for the randomness of mutagenesis and its large excess of deleterious mutations. In unicellular organisms with their large population sizes this is not a restriction, but it may be a huge hurdle in multicellular organisms and it certainly would not be an adaptive response in small-size populations. Sexual reproduction solved this problem by subjecting a huge population of germ cells and gametes to a selection cascade whose pre-selection principle is based on stress resistance (insensitivity to disturbance) and resilience (the rate of recovery after disturbance), competitiveness, and viability of the selected units (Parsons, 1997; 2005; 2007). As a corollary to the fact that most mutations are deleterious and that the gametes with the least vulnerability are selected, only the gametes survive that are the most robust to mutations. A huge body of literature shows that living systems on all levels of organization—from macromolecules to whole organisms—are to some extent robust to mutations (de Visser et al., 2003; Wilke and Adami, 2003; Kitano, 2004; Stelling et al., 2004; Wagner, 2005a; b; c; Kaneko, 2007; Mihaljev and Drossel, 2009; Draghi et al., 2010; Freilich et al., 2010). Mutationally robust systems are often also robust to environmental change (Ancel and Fontana, 2000; Meiklejohn and Hartl, 2002; de Visser et al., 2003; Milton et al., 2003; Wagner, 2005a; Cooper et al., 2006; Masel and Siegal, 2009; Szollosi and Derenyi, 2009; Lehner, 2010; Hayden et al., 2012; Stewart et al., 2012). Evidence suggests that insensitivity to environmental perturbations and robustness against mutations are generally correlated (Meiklejohn and Hartl, 2002; Remold and Lenski, 2004); hence, selection to promote survival under a large variety of environments might indirectly increase mutational resilience (Papp et al., 2009). Importantly, the ability to detoxify ROS, or resist oxidative stress, must be heritable (Olsson et al., 2008).

Stress tolerance determines competitive properties and outcome of biotic interactions (Liancourt et al., 2005). Julian Huxley (1942, 1947) and later JZ Young (1951) advocated a concept of biological progress according to which broadly increasing is adaptation to life in general, rather than to any particular mode of life. One of the important features of this progress is increasing independence of the environment. This concept was derived from taking into account the characteristics distinguishing dominant groups from both their non-dominant contemporaries and their dominant predecessors. p53 is a key mediator of stress/oxidative stress resistance selection regimes (Derry et al., 2001). Importantly, the modulation of stress resistance by p53 is context-dependent, depending on the level of mitochondrial bioenergetics stress. The C. elegans p53 ortholog cep-1 is required to extend longevity in response to mild suppression of mitochondrial proteins that are involved in electron transport chain-mediated energy production, and also mediates both the developmental arrest and life-shortening induced by severe mitochondrial stress (Ventura et al., 2009; Torgovnick et al., 2010). The context-dependence of p53 family proteins on stress resistance and longevity was shown in a variety of organisms ranging from C. elegans and Drosophila to humans (Feng et al., 2011). Involving p53 signaling pathways that effect mitochondrial homeostasis and function (Matoba et al., 2006; Liu B et al., 2008; Vaseva and Moll, 2009; Galluzzi et al., 2011; Maddocks and Vousden, 2011), the “Mitochondrial Threshold Effect Theory”, holds that it depends on the severity of the mitochondrial dysfunction whether a mutation results in somatic and cellular life span elongation or reduction (Lee SS et al., 2003; Sampayo et al., 2003; Kondo et al., 2005; Ventura et al., 2006).

Stress resilience (i.e. oxidative stress resistence) and energetic efficiency are both key fitness components and pervasive selection criteria in resource-limited habitats (MacArthur and Wilson, 1967; Zotin, 1990; Parsons, 1997; 2005; 2007). Fitness can be approximated to energetic efficiency especially towards the limits of survival. Furthermore, fitness at the cellular level should correlate with fitness at the organismal level, especially for development time, survival and longevity; ‘good genotypes’ under stress should therefore be at a premium (Parsons, 2005). Asking which systems-level aspects of metabolism are likely to have adaptive utility and which could be better explained as by-products of other evolutionary forces, Papp et al. (2009) concluded that the global topological characteristics of metabolic networks and their mutational robustness are unlikely to be directly shaped by natural selection. Conversely, models of optimal design revealed that various aspects of individual pathways and the behavior of the whole network show signs of adaptations, even though the exact selective forces often remain elusive (Papp et al., 2009).

The molecular processes associated with sexual reproduction reveal a complexity that goes far beyond the simple sexual reproduction = meiotic recombination scenario. Sexual reproduction had to integrate a host of different strategies into a single coherent process. I try to trace the evolutionary course of events that may have led to this complex process.

  1. Shuffling genes to bring favorable alleles from different genomes together into one genome is a first approach. But, as countless theoretical models have shown, it appears to be not enough to give sexual reproduction a competitive advantage over asexual reproduction, particularly when recombination load (see chapter 18.1) threatens to offset this advantage. In addition, the advantage may disappear when asexual reproduction uses mitotic recombination (Mandegar and Otto, 2007).
  2. To become an evolutionary success, sexual reproduction had to tap into another resource: mutagenesis. Theoretical modeling has delineated the basic conditions under which this innovation took place. Maynard Smith (1971b) showed that for larger populations, sex may accelerate evolution by a factor very approximately equal to the number of loci at which at any time favorable mutations are possible but have not yet occurred. Otto and Hastings (1998) demonstrated that pre-selection of gametes would have significant effects on the frequencies and types of mutations and alleles in a population. Homologous recombination requires DSBs and oxidative stress. And oxidative stress is the fuel that drives mutagenesis and delivers free of charge the process to select the pearls among the pebbles. In addition, oxidative stress fires the processes related to epimutagenesis and transgenerational inheritance (see chapter 16). Due to the unfavorable ratio between deleterious and beneficial (epi)mutations, a large number of gametes had to be produced to get a reasonable number of quality gametes. The metabolic investment further increased the oxidative stress, a runaway process that brought human spermatogenesis to the brink of error catastrophe.

Sexual reproduction generated a mechanism that, in its consequence, is a beneficial-mutation enrichment system. However, the quality check regime is far from optimal (Hartshorne et al., 2009). This may have several reasons: (i) the quality checks based on cellular stress-resistance cannot reflect the complexity of the environment and its manifold adaptive challenges; (ii) the system cannot check the proper functioning of differentiated cells in a multicellular organism with its division of labor; (iii) although the stochasticity of the gametic bottleneck is attenuated by the manifold quality checks, a residual stochastic, possibly fitness-eroding, element remains and (iv) the system relies on the bet-hedging strategy that must allow for a less than optimal quality check to create e.g. “hopeful monsters” (Chouard, 2010).

11.1 Excursion: RNA virus quasispecies

A hallmark of RNA genome homeostasis is the error-prone nature of their replication and retrotranscription. Viral RNA polymerases exhibit characteristically low fidelity with measured mutation rates of roughly 10-4 mutations per nucleotide copied, which is orders of magnitude greater than those of nearly all DNA-based viruses and organisms (Batschelet et al., 1976; Holland et al., 1982; Steinhauer and Holland, 1987; Lauring and Andino, 2010). The major biochemical basis of the limited replication fidelity is the absence of proofreading/repair and postreplicative error correction mechanisms that normally operate during replication of cellular DNA. In spite of this unique feature of RNA replicons, the dynamics of viral populations seems to follow the same basic principles that classical population genetics has established for higher organisms (Domingo et al., 1996; 2012; Moya et al., 2000). The quasispecies concept was introduced by M. Eigen and co-workers as a formal mathematical model that was initially formulated to explain the evolution of life in the ‘‘precellular RNA world” (Eigen, 1971; Eigen and Schuster, 1977; Eigen et al., 1988). The quasispecies model can be viewed as a simple framework that contains all the basic ingredients of Darwinian evolution. In particular, it captures the critical relation between mutation rate and information transmission (Eigen, 1971; Eigen and Schuster, 1977). According to Jenkins et al. (2001) the quasispecies model is an equilibrium mutation-selection process and describes a heterogeneous distribution of whole genomes ordered around one or a degenerate set of fittest sequences known as ‘‘master’’ sequences (Eigen 1987, 1992; 1993, 1996; Nowak, 1992). Genomes are not independent entities due to mutational coupling among variants, and instead, the entire mutant distribution forms an organized cooperative structure which acts like (quasi) a single unit (species), hence its name. The quasispecies evolves to maximize the average rate of replication of the entire mutant distribution rather than the frequency of the single fittest sequence in the population, and thus the target of natural selection has been proposed to be ‘the mutant distribution as a whole,’ as opposed to individual variants (Eigen 1987, 1992; 1993, 1996). Wilke (2005) pointed out that quasispecies theory is simply a subset of theoretical population genetics, and it is mathematically equivalent to the theory of mutation-selection balance. Quasispecies theory treats multiple loci, whereas early work on mutation-selection balance has focused on one- or twolocus models. On the other hand, most work on population genetics considers finite populations and includes stochastic effects, whereas quasispecies theory is first and foremost a deterministic description of infinite populations. Its basic equations are very similar in structure to mutation-selection equations that are studied in the population genetics literature (Baake and Gabriel, 2000; Bürger, 2000; Baake and Wagner, 2001). The level of genetic diversity in a viral population, termed the quasispecies cloud size, is an intrinsic property of the quasispecies. The theoretical advantage of maintaining a diverse quasispecies is that, when the virus is shifted to a new environmental niche or selective regimen, a variant may already be present in the population which will be more fit in the new environment (Schneider and Roossinck, 2001).

Eigen and Schuster (1977) investigated the relationship between the length of genetic sequences and the rate of error in replication. When the mutation rate is adjusted upwards, the rate of adaptation increases. A quantitative analysis revealed that evolution rates increase linearly with mutation rates for slowly mutating viruses. However, this relationship plateaus for fast mutating viruses (Sanjuán, 2012). But then, all of a sudden, at a critical mutation rate, the whole system collapses in what is referred to in the literature as an “error catastrophe.” This may be generally, but vaguely, circumscribed as a critical mutation rate beyond which mutation can no longer be controlled by selection and leads to genetic degeneration. More specifically, it was originally described for the sharply-peaked landscape and defined as a critical mutation rate above which the fittest genotype is lost from the population (Eigen et al., 1989; Biebricher and Eigen, 2005; Bull et al., 2005). What happens is that the system of short-term memory carriers can no longer maintain a long-term memory—not only can the system not adapt further, it also loses everything that it had previously gained (Andersson, 2011). Evolvable systems have to navigate the space defined by conservation and innovation.

RNA viruses typically have short generation times and large populations, both of which contribute to their observed quasispecies structure (Domingo and Holland, 1997; Duffy S et al., 2008). As RNA virus genetic replicative machinery is generally about a millionfold more error prone than that of organismal DNA, detailed analysis of RNA virus evolution over a 50-year period is essentially equivalent to study of an organismal system over a 50-million-year period (Holland et al., 1982). In different virus species there is a wide range of intrahost quasispecies cloud size (Domingo et al., 1992; 1998; 2012; Plyusnin et al., 1996; Bonneau et al., 2001; Alves et al., 2002; Farci et al., 2002; Wang WK et al., 2002; Beasley et al., 2003; Biek et al., 2003; Davis et al., 2003; Lin SR et al., 2004; Jerzak et al., 2005; Ciota et al., 2009; Brackney et al., 2010). It may be much higher than interhost variation, reflecting the action of purifying selection (Holmes, 2003a; Chain and Myers, 2005; Jerzak et al., 2005; Edwards et al., 2006; Suzuki, 2006; Redd et al., 2012) that is determined by the extent of host-specific adaptation and other selection pressures (Nichol et al., 1993; Schneider and Roossinck, 2001). Since its introduction into the US in 1999, West Nile virus (WNV) remains a relatively homogeneous virus population, with the most divergent strains containing only a few nucleotide and/or amino acid substitutions (Anderson et al., 2001; Ebel et al., 2001; 2004; Lanciotti et al., 2002; Beasley et al., 2003; Davis et al., 2003). However, intrahost genetic diversity in naturally and experimentally infected mosquitoes and birds revealed that WNV populations are structured as quasispecies and documented strong purifying natural selection in WNV populations (Jerzak et al., 2005; 2008).

Quasispecies genetic variation can be reduced by purifying selection and by genetic bottlenecks (Yuste et al., 1999; Lázaro et al., 2003; Li and Roossinck, 2004; Davis CT et al., 2005; Chen and Holmes, 2006; Duffy S et al., 2008; Cuevas et al., 2009; Redd et al., 2012). Genetic bottlenecks are stochastic events that cause strong reductions in the effective population size and take place when only a few individuals of a population found a new one. Therefore, bottlenecks reduce the genetic diversity of a population. When a few, or even a single genome, of a viral quasispecies is randomly chosen to generate a new population, there is a high probability that it carries a deleterious mutation relative to fitter genomes of the parental population. This mutation will be transmitted to all the members of the new population (Manrubia et al., 2005). Bottleneck events are frequent in the course of the life cycles of viruses, not only in the most obvious case of host-to-host transmission (Artenstein and Miller, 1966; Couch et al., 1966; Ali et al., 2006; Chen and Holmes, 2006; Carrillo et al., 2007; Betancourt et al., 2008; Smith et al., 2008; Haaland et al., 2009; Redd et al., 2012), but also during the intrahost spread of virus (Foy et al., 2004; Li and Roossinck, 2004; Scholle et al., 2004; Quer et al., 2005; Ali et al., 2006; Pfeiffer and Kirkegaard, 2006; Betancourt et al., 2008; Smith et al., 2008; Haaland et al., 2009; Bull et al., 2011; Domingo et al., 2012). This balance between the continuous generation of new mutants and the action of positive and negative selective forces and bottlenecks acting on complex ensembles of replicating units leads to a very dynamic, though highly organized population (Domingo et al., 1978, 2001). What is left by purifying selection and transmission bottlenecks to the action of natural selection is only the proverbial “tip of the iceberg” of quasispecies genetic variation (Clarke et al., 1994; Li and Roossinck, 2004; Chen and Holmes, 2006). Despite the fast sequence evolution that is characteristic of viruses, sets of genes are conserved in large groups of viruses (Argos et al., 1984; Kamer and Argos, 1984; Goldbach, 1987; Koonin and Dolja, 1993). Analytical and simulation methods demonstrated that severe bottlenecks of viral transmission are likely to drive down the virulence of a pathogen because of stochastic loss of the most virulent pathotypes, through a process analogous to Muller’s ratchet. Patterns of accumulation of deleterious mutation may explain differing levels of virulence in vertically and horizontally transmitted viral diseases (Novella et al., 1995; 1999; Bergstrom et al., 1999; Lázaro et al., 2003; Novella, 2004; Cuevas et al., 2009).

Population bottlenecks are stochastic events that strongly condition the structure and evolution of natural populations. Considering that most mutations are deleterious, it was predicted that the frequent application of bottlenecks would yield a population unable to replicate. According to theory, in experiments with viruses experiencing a high replication error rate due to the action of mutagens, but in the absence of bottlenecks, real extinctions of infectivity are observed (Sierra et al., 2000; Crotty et al., 2001; Pariente et al., 2001; Grande-Pérez et al., 2002; Severson et al., 2003). However, in vitro as well as in vivo systems evolving through bottlenecks present a remarkable resistance to extinction (Manrubia et al., 2005; Cases-González et al., 2008). Therefore, it is reasonable to assume that there must be mechanisms able to counterbalance the negative effect of bottlenecks. In the case of vertically transmitted viruses, bottlenecks may be particularly severe, since only a small amount of viruses are able to cross the barriers to infect the embryo. Competition among genomes can only take place inside the infected host. Bottlenecks are also very frequent in horizontally transmitted viruses and during replication inside an infected organism (Gerone et al., 1966; Nowak et al., 1991; Pang et al., 1992; Quer et al., 2005; Ali et al., 2006; Pfeiffer and Kirkegaard, 2006; Roossinck and Schneider 2006; Carrillo et al., 2007; Betancourt et al., 2008; Smith et al., 2008; Haaland et al., 2009; Domingo et al., 2012). Under horizontal transfer, each virus can infect every susceptible individual of the host population, and competition can also happen at the inter-host level (Wilson et al., 1992; Chao et al., 2000). The inter- and intra-host competition that takes place in horizontally transmitted viruses should mitigate the effect of bottlenecks (Bergstrom et al., 1999). This means that an intra- and inter-host competition among viruses selects the more virulent forms, which have greater chances of infecting a new host (Manrubia et al., 2005). These findings contain important messages:

(i) Both theoretical (Wagner and Gabriel, 1990; Kondrashov, 1994; Gordo and Charlesworth, 2000) and experimental (Chao, 1990; Coates, 1992; Duarte et al., 1992; 1993; Clarke et al., 1993; Escarmís et al., 1996; 2002; 2006; 2009; Yuste et al., 1999; de la Peña et al., 2000; Lázaro et al., 2003; Novella, 2004; Novella and Ebendick-Corpus, 2004; de la Iglesia and Elena, 2007; Weatherford et al., 2009; Jaramillo et al., 2013) work has shown that Muller’s ratchet operates in stochastic bottlenecks resulting in fitness losses and, eventually, in the extinction of the population.

(ii) On the other hand, bottlenecks associated with selective events increase or at least do not attenuate the fitness of the transmitted viruses (Manrubia et al., 2005; Sagar, 2010; Blish, 2012) that present a remarkable resistance to extinction (Manrubia et al., 2005; Cases-González et al., 2008). Interestingly, several investigations have suggested that transmitted HIV-1 and/or early variants in the recipient are more closely related to the donor’s ancestral sequences suggesting that variants with ancestral features were favored for transmission, with evolution starting all over in newly infected individuals (Zhu et al., 1993; Herbeck et al., 2006; Sagar et al., 2009; Blish, 2012).

The genetic diversity within an RNA virus quasispecies population facilitates adaptation to and improved fitness on existing or changing environments through natural selection, so long as population sizes are large (Holland et al., 1982; 1991; Kurath and Palukaitis, 1990; Fitch et al., 1991; Coffin, 1992; Gorman et al., 1992; Kinnunen et al., 1992; Martinez et al., 1992; Domingo and Holland, 1997; Duffy S et al., 2008). For example, the quasispecies structure has been identified as an important determinant in viral evasion of the host immune response and the development of resistance to antiviral drugs and therapies (Domingo et al., 1998; 2012; Essajee et al. 2000; Farci and Purcell, 2000; Farci et al., 2002). Therefore, the mutational diversity of RNA viruses facilitates their persistence at the cellular, organismal and population levels.

The quasispecies pattern of virus evolution is only recognized by looking at the mutational dynamics of virus replication within the cells (Batschelet et al., 1976; Holland et al., 1982; Steinhauer and Holland, 1987; Fitch et al., 1991; Coffin, 1992; Gorman et al., 1992; Lauring and Andino, 2010). Countless mutations occur, but only a miniscule fraction can survive repeated passages from host to host (Clarke et al., 1994). A rapid intrahost evolution need not be reflected in rapid evolution at the interhost population level (the one resulting from multiple interhost transmissions) (Fryer et al., 2010; Domingo et al., 2012). For the same virus a higher rate of evolution is obtained when the compared sequences correspond to viruses isolated within a short time span than when they correspond to isolates separated by a long time interval (Sobrino et al., 1986; Domingo, 2007). This dynamic could not have been recognized by investigating virus evolution only from transmission bottleneck to transmission bottleneck, treating intracellular virus replication as black box. So far, evolutionary theory approached the evolutionary dynamics of sexual reproduction exclusively from an outcome perspective focusing on recombination and gamete bottlenecks, treating the molecular processes of sexual reproduction within the gonads more or less as black box. And therefore missed its mutation-selection balance with its quasispecies dynamics. Importantly, the quasispecies cloud size in sexual reproduction encompasses both mutated (including recombined) and epimutated variants. The quasispecies theory explained quantitatively the boundaries, within which evolution works best (Eigen et al., 1988, Eigen, 1992). There should have been no evolutionary constraints that could have hindered higher taxa to tap into this evolutionary resource of quasispecies dynamics. Thus, populations of mutated gamete populations have a great deal in common with virus quasispecies: master sequence and cloud size, purifying selection due to sexual selection cascades, and gamete bottlenecks.

The quasispecies concept has also been applied to other biological systems such as prions, bacteria and tumor cells (Más et al., 2010; Ojosnegros et al., 2011; Domingo et al., 2012). In the following, I largely adhere to the excellent treatise of Más et al. (2010) on this topic. Cancer cells may show a mutator phenotype (Loeb, 1998; 2001; Bielas et al., 2006) that may increase the probability of achieving the most advantageous mutation combination for tumor growth. Every tumor harbors high-frequency mutations, usually those resulting in the gain of function of an oncogene or the loss of a tumor suppressor, accompanied by a complex combination of low-frequency mutations. These mutations are thought to drive the global cancer phenotype, and their characteristics resemble those of viral quasispecies with the presence of a dominant “master” clone accompanied by a “clan” of minor forms, the quasispecies cloud (Más et al., 2010). The mutant “clan” generated during cancer cell replication allows the tumor to face diverse challenges, including the immune system and treatment (Más et al., 2010). Cancer can be considered as a complex biological system that evolves through mutations and epigenetic changes following Darwinian principles of competition and selection (Merlo et al., 2006). Theoretical studies have correlated cancer and genetic instability (Maley and Forrest, 2000; Gonzalez-Garcia et al., 2002) with quasispecies models of minimal replicators (Solé et al., 2003; Brumer et al., 2006; Tannenbaum et al., 2006) and even with incursions into error catastrophe (Solé and Deisboeck, 2004).

11.2 SMSC, bottlenecks and natural selection

The Neo-Darwinian idea that evolution is driven by purely random germline mutations followed by independent natural selection on the progeny has become a widely accepted dogma in biology. Moreover, it is conventional wisdom that natural selection is much more efficient in invertebrate species with large effective population sizes, such as C. elegans, than in those with relatively small effective population sizes, such as many vertebrates, as reflected in many aspects of genome evolution (Lynch and Conery, 2003; Denver et al., 2004). Species with small census population sizes routinely go through narrow gamete bottlenecks, i.e. from a large, sometimes huge, number of gametes only a small number of gametes succeed to become offspring. This gamete bottleneck is thought to be stochastic: since all cells within a multicellular individual are recently derived from a common single-celled ancestor (the zygote or spore) and are hence closely related, with little expected variation,it is thought that selection between gametes can be ignored (Maynard Smith and Szathmáry, 1995, p. 244; Otto and Hastings, 1998).Within this conceptual framework, the huge number of gametes, respectively the large number of cell divisions per generations necessary to produce this large number, is a definite burden. From zygote to zygote, the number of cell divisions has been estimated to be: 50 for maize (Otto and Walbot, 1990), 35 for Drosophila (at 18 days for males and at 25 days for females; Drost and Lee, 1998), 25 for female mice (Drost and Lee, 1998), 62 for 9-month-old male mice (Drost and Lee, 1998), 23 for human females (Vogel and Rathenberg, 1975), and 36 for human males at puberty plus 23 per year thereafter (Vogel and Rathenberg, 1975). The number of cell divisions per generation is strongly age dependent in Drosophila and in mammalian males, rising, for example, in human males from 36 for a 13-year-old to over 500 for a 35-year-old (Otto and Hastings, 1998). That the large number of generations is a relevant factor with regard to mutation load is illustrated by the male mutation bias that, at least in part, is replication-dependent. If the gamete bottleneck is stochastic, sexually reproducing organisms are in trouble. Both theoretical (Wagner and Gabriel, 1990; Kondrashov, 1994; Gordo and Charlesworth, 2000) and experimental (Chao, 1990; Coates, 1992; Duarte et al., 1992; 1993; Clarke et al., 1993; Escarmís et al., 1996; 2002; 2006; 2009; Yuste et al., 1999; de la Peña et al., 2000; Lázaro et al., 2003; Novella, 2004; Novella and Ebendick-Corpus, 2004; de la Iglesia and Elena, 2007; Weatherford et al., 2009; Jaramillo et al., 2013) work has shown that Muller’s ratchet operates in stochastic bottlenecksresulting in fitness losses and, eventually, in the extinction of the population. This would mean that sexually reproducing organisms, rather than asexually reproducing, have a huge risk to suffer from mutational meltdown. Estimates of human germline base substitution rates range from 0.97 to 3.8 x 10-8 per base per generation (Haldane, 1935; Kondrashov and Crow, 1993; Crow, 1993b; Nachman and Crowell, 2000; Kondrashov, 2003; Xue et al., 2009; Lynch, 2010b; Roach et al., 2010; The 1000 Genomes Project Consortium, 2010; Awadalla et al., 2011; Conrad et al., 2011; Kong et al., 2012; Sun et al., 2012). A human mutation rate of ~2.5 x 10-8 mutations per nucleotide site corresponds to 175 (range 91–238) mutations per diploid genome per generation (Nachman and Crowell, 2000). Assuming that (i) there is no selective bottleneck between gametogenesis and offspring and (ii) 95 to 100% of all mutations have either neutral (~30% of amino-acid changing mutations in humans, ~16% in Drosophila [Eyre-Walker, 2002], and ~2.8% in enteric bacteria [Charlesworth and Eyre-Walker, 2006]) or deleterious effects on fitness (Eyre-Walker and Keightley, 2007) this ratio would mean that with each generation, fitness of human populations would erode. In addition, the degree of fidelity in epigenetic transmission is about three orders of magnitude lower than that of DNA sequence (an error rate of 1 in 106 and 1 in 103 for DNA sequences and DNA modification, respectively) (Ushijima et al., 2003; Laird et al., 2004; Riggs and Xiong, 2004; Genereux et al., 2005; Fu et al., 2010; Petronis, 2010). Thus, a huge number of (epi)mutations with possibly detrimental phenotypic effects may degenerate germline performance in each new generation, contributing to an almost certain (epi)mutational meltdown.

On the other hand, studies in RNA viruses have suggested that bottlenecks associated with selective events increase or at least do not attenuate the fitness of the transmitted viruses (Manrubia et al., 2005; Sagar, 2010; Blish, 2012) that present a remarkable resistance to extinction (Manrubia et al., 2005; Cases-González et al., 2008). I have presented extensive evidence for the selectivity of the gamete/offspring bottleneck ( see chapter 8). Assuming that bottlenecks are selective, the narrower the bottleneck the more stringent the underlying selection process can be expected to be.

Sexual mutagenesis and selection cascades and gamete bottlenecks are mutually interdependent. The one would not work without the other. Accordingly, they are finely tuned to each other and to the actions of natural selection that provides the environmental arena for the “struggle for survival” of the pre-selected offspring populations. A general pattern can be recognized that allows to distinguish between three broad categories of organisms (see figure 1):

Category 1 (Microorganisms): microscopic body size, large population size, condition-dependent sexual reproduction or asexual reproduction, stress-dependent mutagenesis, “pearls” are selected by natural selection

Category 2 (Invertebrates): intermediate body size and population size. Ectotherms. Sexual reproduction with moderate (epi)mutagenesis and less stringent gamete and offspring quality control resulting in relaxed gamete bottleneck. Intermediate fecundity. Moderate impact of SMSC on preselected variation. Moderate impact of natural selection.

Category 3 (Mammals and birds): large body size, small population size, endotherms. Sexual reproduction with strong (epi)mutagenesis and stringent quality control resulting in narrow gamete bottleneck. Low fecundity. High impact of SMSC on preselected variation. Lower impact of natural selection.

West and colleagues (West and Deering, 1995; Bickel and West, 1998; West and Bickel, 1998) studied the evolution rates of mammalian DNA, based on recent estimates of numbers of nonsynonymous substitutions in 49 genes of human, rodents, and artiodactyls. The rate variations are better described by lognormal statistics, as would be the case for a multiplicative process, than by Gaussian statistics, which would correspond to a linear, additive process. The theoretical explanation of these statistics requires the evolution of different substitution rates in different genes to be a multiplicative process in that each rate results from the interaction of a number of interdependent contingency processes.A multiplicative process is one that will only occur if a number of sub-events take place, so that the probability of the grand event occurring equals the product of the probabilities of each of the sub-events occurring (West and Deering, 1995; Bickel and West, 1998; West and Bickel, 1998). The joint action of the SMSC-natural selection complex could provide the sub-events for this multiplicative process.

Depending on the distribution of fitness effects of new mutations, Gillespie defined three specific model-based domains of molecular evolution (Gillespie, 1999;Jensen and Bachtrog, 2011). In the Ohta domain (Ohta, 1973), patterns of molecular evolution are driven mainly by slightly deleterious mutations (Gillespie, 1999). In this domain, the rate of substitution decreases with increasing effective population size, due to an increase in the efficiency of purifying selection against deleterious mutations. In the Kimura domain (Kimura, 1968), molecular evolution is dominated by mutations with no effect on fitness, and the rate of substitution is independent of the effective population size but simply given by the neutral mutation rate (Gillespie, 1999). Finally, in the Darwin domain, molecular evolution is driven by beneficial mutations, and the rate of substitution is predicted to increase with effective population size (Gillespie, 1999; Jensen and Bachtrog, 2011).If beneficial mutations are independent, rates of adaptation increase linearly with increasing population size. However, if beneficial mutations are common and linked, the rate of substitution will be substantially reduced and eventually become independent of the effective population size of a species (Gillespie, 2000). Within the SMSC regime, a greater fraction of changes will be explored in the larger effective population size of gametes and the role of stochastic processes will be reduced. Thus, the role of the Ohta domain is greatly reduced, as has been suggested by Otto and Hastings (1998)and the Darwin domain is increased.

There is considerable controversy about thefitness effects of mutations in mutation accumulation (MA) studies (Bataillon, 2000; 2003; Keightley and Lynch, 2003; Shaw et al., 2003) and the power of MA studies to detect beneficial mutations (Bataillon, 2000; 2003).These studies involve the use of experiments where mutations are allowed to accumulate in replicate populations under laboratory conditions and in the relative absence of natural selection. At the end of these experiments, the replicate lineages are then assessed for a range of fitness measures against a control lineage and an estimate of the number of new deleterious mutations per zygote per generation is then made (Lynch et al., 1999).Spontaneous MAexperiments in Drosophila (Mukai et al., 1972; Fry et al., 1999; Lynch et al., 1999; Chavarrias et al., 2001), Caenorhabditis elegans (Vassilieva and Lynch, 1999; Vassilieva et al., 2000), Daphnia (Lynch et al., 1998), wheat (Bataillon et al., 2000), yeast (Wloch et al., 2001a; Zeyl and de Visser, 2001), E. coli (Kibota and Lynch, 1996), and Arabidopsis (Schultz et al., 1999; Wright et al., 2002) detected downward trends in the mean fitness of MA line populations relative to control populations as generations accrued.However, equivocal results were obtained in Arabidopsis thaliana (Shaw et al., 2000; Shaw et al., 2002), Drosophila (Gilligan et al., 1997), C. elegans (Keightley and Caballero, 1997), and yeast (Zeyl et al., 2001), although some of these data were questioned (Keightley and Lynch, 2003), [but see Shaw et al. (2003) and Bataillon (2003)].In addition, many animal models are maintained under such unnaturally favorable conditions that accumulation of deleterious mutations and fitness decline is inherent (Bryant and Reed, 1999; Lynch and O’Hely, 2001; Araki et al., 2007; 2009; Hall and Colegrave, 2008), an effect that cannot be prevented by the SMSCs of sexual reproduction. Importantly these animal models and MA experiments were conducted in taxa corresponding to the invertebrate type of the SMSC-natural selection pattern (see figure 1B) with larger census population sizes and less rigorous sexual selection cascades in which natural selection still plays an important role in maintaining population fitness.

In Drosophila, gamete population size during SMSC is small in relation to the population size under natural selection. In humans, the SMSC selects from a much larger gamete population size than processes related to natural selection. Accordingly, in Drosophila species (where population size is about 106), the proportion of nonsynonymous substitutions that have been fixed by positive selection is about 50% while for hominids (with population size around 10,000), this proportion is close to zero. Similarly, the proportion of nonsynonymous mutations that are effectively neutral is less than 16% in Drosophila, whereas it is about 30% in hominids (Eyre-Walker and Keightley, 2007). When selection at the cell and individual levels act in a cooperative manner, increased rather than decreased opportunity for germline selection will be favored by evolution (Otto and Hastings, 1998).

A riddle first pointed out 40 years ago on the basis of comparative allozyme data is the mysteriously narrow range of genetic diversity levels seen across taxa that vary markedly in their census population sizes (Lewontin, 1974; Leffler et al., 2012). As discussed earlier, the effective size of a population (Ne) determines the probability of (and time to) fixation or removal of mutant alleles (Lynch, 2006; Charlesworth, 2009). Extensive information on absolute population sizes, recombination rates, and mutation rates strongly supports the view that eukaryotes have reduced genetic effective population sizes relative to prokaryotes, with especially extreme reductions being the rule in multicellular lineages (Lynch, 2006). According to Kimura (1962), a simple population genetic model relates the probability of fixation of a mutation to effective population size and strength of selection. However, molecular evolution studies showed no conspicuous footprints of population size (Lewontin, 1974; Kimura, 1983; Gillespie, 1991a; b; 1999; 2001). For example, estimates of silent (Nachman, 1997) or amino acid (Nevo et al., 1984) variation among species are distressingly similar, suggesting at face value that the effective sizes of most species–from bacteria to humans–are within one order magnitude of each other (Lewontin, 1974). Accordingly, mathematical models that attempted to correlate polymorphism/molecular substitution rates with effective population sizes (Nevo et al., 1997a; Huzurbazar et al., 2010) faced a problem. Higher taxa that invariably have lower Ne than lower taxa have higher polymorphism/molecular substitution rates than can be explained by their Ne. Thus at face value, with decreasing effective population size the strength of selection increased (Huzurbazar et al., 2010). Alternative explanations for the insensitivity of molecular evolution to population size included: that natural selection is challenged by genetic drift (Ellegren, 2009); that a combination of population bottlenecks and historical effects may blunt the effects of N on evolution (Nei and Graur, 1984); that hitchhiking is analogous to bottlenecks in its ability to reduce genetic variation and render it less sensitive to population size (Maynard Smith and Haigh, 1974); and that linkage can amplify the effects of selection in small effective population size species (Nevo et al., 1997a). Moreover, it  was argued that similarity of obvious mutation rates among lineages with vastly different generation lengths and physiological attributes points to a much greater contribution of replication-independent mutational processes to the overall mutation rate (Kumar and Subramanian, 2002). With the insights gained from the eco-evo theory of sexual reproduction, it can be concluded that the riddle resulted from the perspective restricting evolutionary processes to postnatal populations exposed to natural selection. Relating evolutionary processes to the joint (but presumably not additive) germ cell and postnatal population sizes that are subjected to both SMSCs and natural selection (figure 1) can be expected to resolve the molecular evolution–population size conundrum. That evolutionary processes in populations with vastly differing population sizes, generation lengths and SMSCs yielded surprisingly similar evolutionary rates (that are not dependent on mutation rate, Huang, 2008) points to a general evolutionary principle that may guide all of life.

12. Random trial or educated guess?


The organisms we see today are deeply marked by the selective action of two (now we know it’s almost four, KH) thousand million years’ attrition. Any form in any way defective in its power of survival has been eliminated; and today the features of almost every form bear the marks of being adapted to ensure survival rather than any other possible outcome. Eyes, roots, cilia, shells and claws are so fashioned as to maximize the chance of survival. And when we study the brain we are again studying a means to survival.

Ross W. Ashby (1956, p. 196) An Introduction to Cybernetics

Summary

Evolution is a cybernetic process that has all attributes of a learning automaton. The feedback that is essential for any learning process comes from the selective survival/death of organisms, bestowing “educated guess” characteristics upon evolutionary processes. Educated guess approaches limit the search space and increase the likelihood that some of the variations generated will be useful. The Baldwin effect states that the ability of individuals to learn can guide and accelerate the evolutionary process. Non-stochastic engineering tools are the evolutionary result of learning processes: stress-induced mutagenesis, non-random recombination, mutability of simple sequence repeats, activity of transposable elements, phenotypic plasticity, all of which can act as evolutionary tuning knobs when (epi)genetic innovation is needed. Sexual reproduction, combining these elements in a single process, is the commander in chief of transgenerational information transfer.

The Modern Synthesis holds that (i) mutations occur independently of the environment, (ii) mutations are due to replication errors, and (iii) mutation rates are constant (Lenski and Mittler, 1993; Brisson, 2003). Currently, biologists usually agree that all genetic mutations occur by “chance” or at “random” with respect to adaptation (Miller, 2005) and the novel allele is subsequently selected for or against (Davis, 1989; Rosenberg, 1994). The claim dates back to Darwin’s conception of “spontaneous,” “accidental” or “chance” variation (Darwin, 1859; Darwin and Seward, 1903). The Modern Synthesis later redefined Darwin’s idea as rooted in the phenomenon of genetic mutation following a long period of controversy over the “chance” vs “directed” character of variation (Merlin, 2010). It has been argued that mutations must be random because natural selection cannot “assist the process of evolutionary change,” since “selection lacks foresight, and no one has described a plausible way to provide it” (Dickinson and Seger, 1999). Such an evolutionary strategy was called a raffle or lottery (Stockley et al., 1997; Parker et al., 2010) and would correspond to a “ random trial” approach: genetic change would arise at random, independent of its functional consequences and natural selection would decide which lottery ticket wins. However, even in a raffle approach the number of lottery tickets, as in sperm competition games (Parker, 1990), would increase the chances of a win.

However, it would be highly adaptive for organisms inhabiting variable environments to modulate mutational dynamics in ways likely to produce necessary adaptive mutations in a timely fashion. Metzgar and Wills (2000) argued that adaptively tuned mutation rates do not require any special foresight. Instead, they must have been selected for repeatedly in the past for their ability to generate genetic change. Mutational tuning does not require the specific generation of adaptive mutations (nonrandomness with respect to function) but rather the concentration of mutations under specific environmental conditions or in particular regions of the genome (nonrandomness with respect to time or location) (Metzgar and Wills, 2000).

In 1966, mathematicians, physicists and engineers met in Philadelphia (Moorhead and Kaplan, 1967). The mathematicians argued that neo-Darwinism faced a formidable combinatorial problem. Murray Eden illustrated the issue with reference to an imaginary library evolving by random changes to a single phrase: “Begin with a meaningful phrase, retype it with a few mistakes, make it longer by adding letters, and rearrange subsequences in the string of letters; then examine the result to see if the new phrase is meaningful. Repeat until the library is complete” (Eden, 1967). Would such an exercise have a realistic chance of succeeding, even granting it billions of years? In the view of mathematicians, the ratio of the number of functional genes and proteins, on the one hand, to the enormous number of possible sequences corresponding to a gene or protein of a given length, on the other, seemed so small as to preclude the origin of genetic information by a random mutational search. A functional protein one hundred amino acids in length represents an extremely unlikely occurrence. There are roughly 10130 possible amino acid sequences of this length, if one considers only the 20 protein-forming acids as possibilities. In human codes, M. P. Schützenberger argued, randomness is never the friend of function, much less of progress. When we make changes randomly to computer programs, “we find that we have no chance (i.e. less than 1/101000) even to see what the modified program would compute: it just jams.” (Schützenberger, 1967). Not surprisingly, these arguments were grist to the mill of the advocates of Intelligent Design (e.g. Meyer, 2008).

The term cybernetics stems from the Greek κυβερνητης (kybernetes, steersman, governor, pilot, or rudder). Cybernetics concerns the functioning of self-regulatory systems (Wiener, 1948; Ashby, 1956). Cybernetics requires a closed signal loop: action by the system causes some change in its environment and that change is fed to the system via information (feedback) that enables the system to change its behavior. The “random trial” and “trial and error” approaches differ in an important variable: feedback of outcome. “Random trial” lacks the feedback loop: it either cannot find out whether the trial was a success or failure or it is completely unable to learn from this knowledge. The “trial and error” approach has this feedback loop that identifies “errors”. In evolution this is Charles Darwin’s natural selection (Ackley and Littman, 1991) and Alfred R. Wallace’s elimination of the unfit (Smith CH, 2012a; b). A learning system is able to draw its lesson from the error and make its next trial less random. Thus, in contrast to the “random trial” approach, the “educated guess” approach is less random-driven but uses past experience to navigate future direction: the system is able to learn. “Educated guess” approaches limit the search space and increase the likelihood that some of the variations generated will be useful (Jablonka and Lamb, 2007). There is no great mystery about their evolution: they arose through natural selection as a side-effect or modification of functions that evolved for other purposes (Ackley and Littman, 1991; Jablonka and Lamb, 1995; 2005; 2007; Caporale, 1999; 2003a; b; 2009; Radman et al., 1999; Shapiro, 2011; King, 2012).

Learning is defined as any relatively permanent change in behavior resulting from past experience, and a learning system is characterized by its ability to improve its behavior with time, in some sense tending towards an ultimate goal (Narendra and Thathachar, 1974). A vital component of the learning process is the environment. If the environment were relatively static, there might be little need for learning to evolve. Systems could instead evolve to a state where they have innate mechanisms to handle that environment. But if the environment is diverse and unpredictable, innate environment-specific mechanisms are of little use. Instead, individuals need general adaptive mechanisms to cope with arbitrary environments. In this way, a diverse environment encourages the evolution of learning (Chalmers, 1990). A learning automaton is an adaptive decision-making device that learns the optimal action out of a set of actions through repeated interactions with a random environment. The two characteristic features of learning automaton are that the action choice is based on a probability distribution over the action-set and it is this probability distribution that is updated at each instant based on the reinforcement feedback from the environment (Thathachar and Sastry, 2002).

Often evolution is characterized as cybernetic (e.g. Ashby, 1954; 1956; Schmalhausen, 1960; Waddington, 1961a; Corning, 2005). In his 1858 essay, A:R. Wallace referred to the evolutionary principle “as exactly like that of the centrifugal governor of the steam engine, which checks and corrects any irregularities almost before they become evident….”. For Gregory Bateson (1972, p. 435) “the result will be…a self-corrective system. Wallace, in fact, proposed the first cybernetic model.” However, the first account of how a phenotypic change induced by a change in the environment could lead to a change in the inherited genome was provided by Spalding (1837). Spalding’s driver of evolution comprised a sequence of learning followed by differential survival of those individuals that expressed the phenotype more efficiently without learning (Bateson, 2012). Fitness-related differential reproduction is the feedback control that drives the cybernetic system. With this feedback control adaptation proceeds by trial and error (Ashby, 1954).

The Baldwin effect, independently forwarded by Baldwin (1896), Lloyd Morgan (1896), and Osborn (1896), but largely so called because of Baldwin’s influential book (Baldwin, 1902), states that the ability of individuals to learn can guide and accelerate the evolutionary process (Hinton and Nowlan, 1987; French and Messinger, 1994; Weber and Depew, 2003; Sznajder et al., 2012). Currently, this principle is widely used in evolutionary computing and evolutionary algorithms (Ackley and Littman, 1991; Bull, 1999; Eiben and Smith, 2008; Paenke et al., 2009). The Baldwin effect consists of the following two steps (Turney et al., 1996): In the first step, lifetime learning gives individual agents chances to change their phenotypes. If the learned traits are useful to agents and result in increased fitness, they will spread in the next population. This step means the synergy between learning and evolution. In the second step, if the environment is sufficiently stable, the evolutionary path finds innate traits that can replace learned traits, because of the cost of learning. This step is known as genetic assimilation (Arita and Suzuki, 2000). Mathematical models suggest that learning would speed up the adaptation process by providing more explicit information about the environment in the genotype (Sendhoff and Kreutz, 1999; Arita and Suzuki, 2000). Learning alters the shape of the search space in which evolution operates and thereby provides good evolutionary paths towards sets of co-adapted alleles. Hinton and Nowlan (1987) demonstrated that this effect allows learning organisms to evolve much faster than their non-learning equivalents, even though the characteristics acquired by the phenotype are not communicated to the genotype. However, orthodox evolutionary biologists tend to dismiss the Baldwin Effect. The orthodoxy of evolutionary biologists is strongly reductionist, “which implies that the causes and basic mechanisms of evolution are only to be found at the level of genetics” (Parisi et al., 1992). As a consequence, behavior and learning, both being highly holistic processes, have been largely ignored in attempting to understand evolutionary processes. Another reason for the lack of attention is that evolutionary biologists feel “behavior and learning are the province not of biology but of psychology and ethology” (Parisi et al., 1992; French and Messinger, 1994).

The organization and evolution of molecular and organismal diversity in nature at global, regional, and local scales are nonrandom and structured; display regularities across life; and are positively correlated with, and partly predictable by, abiotic and biotic environmental heterogeneity and stress (Nevo, 2001). Hotspots (and, by implication, coldspots) exist for nucleotide mutation, recombination, insertion/deletion, duplication and chromosomal rearrangements and are hence consequential for evolvability (Maki, 2002; Chen B et al., 2007). As discussed in chapter 2, the stress response (a cascade of internal and external changes triggered by stress) is a component of the ecological stress definition (Parker et al., 1999; van Straalen, 2003). The coping with stress is inherent to cybernetic systems. The necessary information feedback loop for learning systems is provided by Alfred R. Wallace’s elimination of the unfit (Smith CH, 2012a; b) and Darwin’s survival and reproduction of the more stress-resilient organisms. In the following, I will discuss a variety of stress-dependent genetic and epigenetic phenomena that are not stochastic and can be inferred of being selected for by cybernetic processes.

12.1 Stress-induced mutagenesis

Sturtevant (1937) introduced the idea that the mutation rate is an evolvable property, and thus subject to optimization by natural selection. A large body of subsequent theoretical work has delimited the broad parameters under which the mutation rate is expected to evolve (Kimura, 1960; 1967; Leigh, 1973; Holsinger and Feldman, 1983; Kondrashov, 1995b; Dawson, 1998; Drake et al., 1998; Johnson, 1999; Sniegowski et al., 2000; André and Godelle, 2006; Baer et al., 2007; Lynch, 2010a; Shaw and Baer, 2011).

Stress-induced mutation is a collection of molecular mechanisms in bacterial, yeast and animal cells that promote mutagenesis specifically when cells are maladapted to their environment, i.e. when they are stressed. In this sense, stress-induced bacterial and eukaryotic mutagenesis (Achilli et al., 2004; Tenaillon et al., 2004; Galhardo et al., 2007; Robleto et al., 2007; Pybus et al., 2010; Rosenberg, 2011; Shee et al., 2011a; b) is a bet-hedging behavior.

The mechanism of stress-induced mutagenesis understood in greatest detail is classical SOS mutagenesis (Friedberg et al., 1995; 2006; Sutton et al., 2000; Schlacher and Goodman, 2007). Suboptimal environmental conditions can influence the small- and large-scale genomic changes that occur in evolving populations. For example, exposure of E. coli cells to starvation, crowding, or nonoptimal temperature activates the SOS response (Finkel and Kolter, 1999; Storz and Hengge-Aronis, 2000; Hastings et al., 2004). SOS elevates rates of homologous and nonhomologous recombination, interferes with DNA polymerase proofreading, and induces error-prone replication (Abbott, 1985; Humayun, 1998; Napolitano et al., 2000; Storz and Hengge-Aronis, 2000; Yeiser et al., 2002).

A phenomenon called adaptive mutation in microbes indicates that mutation rates can be elevated in response to stress, over an order of magnitude higher than extrapolations from fast-growing cells, producing both deleterious and rare beneficial mutations (Loewe et al., 2003; Sniegowski, 2004). Mutations can arise in apparently static bacterial populations when subjected to nonlethal selective pressures (Ryan, 1959; Ryan et al., 1961; 1963; Shapiro, 1984; Cairns et al., 1988). "Adaptive" or "stationary-phase" mutation is a collection of apparent stress responses in which cells exposed to a growth-limiting environment generate genetic changes, some of which can allow resumption of rapid growth (Loewe et al., 2003; Sniegowski, 2004). A stationary-phase-specific process, the expression of the ‘‘growth advantage in stationary phase’’ (GASP) phenotype, depends on the appearance of new mutations in the population (Zambrano et al., 1993; Zambrano and Kolter, 1996; Yeiser et al., 2002; Corzett et al., 2013). The new mutations confer a competitive advantage to cells, allowing them to take over the population. In the well-characterized Lac system of Escherichia coli, reversions of a lac frameshift allele give rise to adaptive point mutations (Cairns and Foster, 1991; Hastings et al., 2004; Foster, 2005). The otherwise high-fidelity process of double-strand break repair by homologous recombination is switched to an error-prone mode under the control of the RpoS general stress response (Ponder et al., 2005; Gonzalez et al., 2008; Shee et al., 2011a). E. coli encodes five DNA polymerases (Goodman, 2002; Johnson and O’Donnell, 2005; Friedberg, 2006). High-fidelity DNA polymerase (Pol) III performs the majority of DNA replication under vegetative conditions, with Pol I is contributing principally to maturation of Okazaki fragments (Kornberg and Baker, 1992; Friedberg, 2006). Three alternative DNA polymerases (Pol II, Pol IV, and Pol V) perform a vital physiological role by mediating translesion synthesis, enabling efficient replication past DNA damage that would otherwise halt replication, albeit with significantly reduced fidelity (Goodman, 2002; Fuchs et al., 2004; Tippin et al., 2004; Nohmi, 2006; Bichara et al., 2011; Corzett et al., 2013). These error-prone DNA polymerases can be induced under a variety of environmental stresses (Taddei et al., 1995; Yeiser et al., 2002; Layton and Foster, 2003; 2005; Stumpf and Foster, 2005; MacPhee and Ambrose, 2010) and have been characterized most extensively following induction of the SOS regulon in response to DNA damage, leading them to be referred to as SOS-induced polymerases (Courcelle et al., 2001; Goodman, 2002; Friedberg, 2006; Nohmi, 2006; Yang and Woodgate, 2007). These alterations increase the rates of point mutation and genome rearrangement, making it likely that SOS plays a role in evolutionary change (Echols, 1981; Radman et al., 1999).

Also in this system, adaptive gene amplification has been documented as a separate and parallel response that allows growth on lactose medium without acquisition of a compensatory frameshift mutation (Andersson et al., 1998; 2011; Hastings and Rosenberg, 2002; Hastings et al., 2004; Hersh et al., 2004; Roth and Andersson, 2004; Kugelberg et al., 2006; Slack et al., 2006). There are two ways that mutations could be directed – selective generation or selective capture (Hall, 1998). In fact, later evidence revealed that stressed cells also accumulate nonadaptive mutations at a higher rate than nonstressed cells, indicating that cells experiencing stress have an increased mutation rate genome-wide, suggesting a transient, possibly epigenetically regulated, hypermutability in a small fraction of the population (Foster, 1997; Torkelson et al., 1997; Rosche et al., 1999; Godoy et al., 2000; Loewe et al., 2003; Bjedov et al., 2003; Gonzalez et al., 2008).

Hypermutators arise through several mechanisms. For instance, general and specific DNA repair systems, mismatch repair (MMR), and oxidative damage repair have been shown to be repressed or inefficient in cells under conditions of stress in eukaryotic and bacterial systems (Feng et al., 1996; Mihaylova et al., 2003; Pedraza-Reyes and Yasbin, 2004; Hara et al., 2005; Saint-Ruf and Matic, 2006; Robleto et al., 2007; Vidales et al., 2009; Debora et al., 2011) while error-prone polymerases are active in stressed cells (Pham et al., 2001; Yeiser et al., 2002; Kivisaar, 2003; Sung et al., 2003; Duigou et al., 2004; 2005; Tegova et al., 2004; Tippin et al., 2004; Hara et al., 2005; Saint-Ruf and Matic, 2006; Corzett et al., 2013).

The occurrence of stress-induced mutagenesis is now widely accepted (Torkelson et al., 1997; Rosche and Foster, 1999; Godoy et al., 2000; Velkov, 2002; Bjedov et al., 2003; Kivisaar, 2003; Rosenberg and Hastings, 2003; Tenaillon et al., 2004; Wright, 2004; Foster, 2005; 2007; Galhardo et al., 2007; 2009; Sundin and Weigand, 2007; Zhang and Saier, 2009; 2011; Cohen and Walker, 2010; Gibson et al., 2010; Heo and Shakhnovich, 2010; Jayaraman, 2011; Pybus et al., 2010; Quinto-Alemany et al., 2011; Shee et al., 2011a; b; Corzett et al., 2013; Mittelman, 2013; but see Hendrickson et al., 2002, Roth et al., 2006; Andersson et al., 2011; Roth, 2011). In populations under various environmental stresses—stationary phase, starvation or temperature-jump—adaptation most often occurs through transient fixation of a mutator phenotype, regardless of the nature of stress. By contrast, the fixation mechanism does depend on the nature of stress. In temperature jump stress, mutators take over the population due to loss of stability of MMR complexes. In starvation and stationary phase stresses, however, a small number of mutators are supplied to the population via epigenetic stochastic noise in production of MMR proteins (a pleiotropic effect), and their net supply is higher due to reduced genetic drift in slowly growing populations under stressful environments (Heo and Shakhnovich, 2010).

Transcription and replication are associated with mutation bias. Interference between DNA replication and transcription have been suggested as source of genomic instability (Poveda et al., 2010; Lin and Pasero, 2012; Helmrich et al., 2013). Transient DSBs are continuously created during transcription and replication (Costanzo et al., 2001; Dmitrieva et al., 2003; Ohnishi et al., 2009). In bacteria and yeast, proportional to transcription rate, transcription-associated mutagenesis (TAM), including single-base substitutions and insertions/deletions, and transcription-associated recombination (TAR) were demonstrated (Herman and Dworkin, 1971; Savic and Kanazir, 1972; Ripley, 1982; de Boer and Ripley, 1984; Keil and Roeder, 1984; Voekel-Meiman et al., 1987; Davis, 1989; Datta and Jinks-Robertson, 1995; Beletskii and Bhagwat, 1996; 2001; Klapacz and Bhagwat, 2002; 2005; Mokkapati and Bhagwat, 2002; Hudson et al., 2003; Reimers et al., 2004; Ross et al., 2006; Saxowsky and Doetsch, 2006; Kim et al., 2007; Fix et al., 2008; Kim H et al., 2010; Pybus et al., 2010; Lippert et al., 2011; Kim and Jinks-Robertson, 2012). Adaptive mutation during prolonged nutritional stress in cells that are not dividing and in genes whose functions are selected is associated with induction of several affected genes (Wright et al., 1999). Recently, TAM was demonstrated in an inducible allele in stationary cells under selective pressure (Pybus et al., 2010). This association could explain why adaptive mutations often appear directed and how cells under these conditions avoid the accumulation of lethal mutations (Davis, 1989). The expression level of a protein is one of the major determinants of molecular evolution (Sharp, 1991; Green et al., 1993; Duret and Mouchiroud, 2000; Pál et al., 2001; 2006; Krylov et al., 2003; Rocha and Danchin, 2004; Zhang and Li, 2004; Agrafioti et al., 2005; Drummond et al., 2005; 2006; Koonin and Wolf, 2006; Rocha, 2006; Drummond and Wilke, 2008). Studies on the yeast Saccharomyces cerevisiae indicate that the strongest predictor of evolutionary rate is expression level of a protein that explains 30–50% of the variation in the rate of protein evolution (Pál et al., 2001; Drummond et al., 2005; 2006), much more than any other known variable. Likewise, transcription enhances site-directed mutagenesis and recombination in mammalian cells (Nickoloff, 1992; Skandalis et al., 1994; Bachl et al., 2001; Aguilera, 2002; Green et al., 2003; Sohail et al., 2003; Saxowsky and Doetsch, 2006; Gottipati et al., 2008; Da Sylva et al., 2009; Mugal et al., 2009; Hendriks et al., 2010).

During transcription, the nontranscribed strand becomes single stranded and is much more vulnerable to most mutagens and oxidative stress than is the double-stranded DNA (Singer and Kusmierek, 1982; Fix and Glickman, 1987; Skandalis et al., 1994; Wright, 2000; Francino and Ochman, 2001) because it is not protected by pairing (Hoede et al., 2006). Thus, the nontranscribed strand has been found to have greater numbers of mutations than the transcribed strand in E. coli (Beletskii and Bhagwat, 1996; Klapacz and Bhagwat, 2005; Fix et al., 2008) and humans (Skandalis et al., 1994; Green et al., 2003; Mugal et al., 2009). On the other hand, preferential repair of DNA base lesions from transcriptionally active genes compared to inactive regions (Bohr et al., 1985; Tu et al., 1996) and transcribed strand-specific repair (Leadon and Cooper, 1993) have been observed. Stressors such as heat (Maresca and Schwartz, 2006), starvation (Hastings et al., 2004), inflammation (Blanco et al., 2007; Lavon et al., 2007), toxins (Salnikow and Zhitkovich, 2008), free radical injury (Cerda and Weitzman, 1997), or other sources of DNA damage (Ponder et al., 2005) can modify gene transcription and thus alter the rate of mutations affecting fitness (Galhardo et al., 2007). Transcription thus has the potential to modify the genetic landscape by locally altering mutation rates, by stimulating loss of heterozygosity and by generating diverse types of rearrangements that include deletions, duplications, inversions and translocations (Davis, 1989; Kim and Jinks-Robertson, 2012).

Bacteria have specific loci that are highly mutable (Moxon et al., 1994). The coexistence within bacterial genomes of such 'contingency' genes with high mutation rates, and 'housekeeping' genes with low mutation rates, is the result of adaptive evolution, and facilitates the efficient exploration of phenotypic solutions to unpredictable aspects of the host environment while minimizing deleterious effects on fitness (Moxon et al., 1994). Mutation bias is a direction-giving factor, invoking the effects of deletion:insertion bias (Petrov and Hartl, 1998), strand-specific nucleotide biases (Beletskii and Bhagwat, 1996), CpG bias (Fryxell and Zuckerkandl, 2000), GC-biased gene conversion (Marais, 2003; Berglund et al., 2009; Duret and Galtier, 2009; Romiguier et al., 2010; Capra and Pollard, 2011) and GC:AT bias (Lobry, 1997; Singer and Hickey, 2000).

Mutagenesis associated with transcription of simple sequence repeats (see chapter 12.3) and TEs (see chapter 12.4) are major sources of genetic variation. Studies on E. coli reveal a major contribution of transposons and insertion elements to mutagenesis (Arber et al., 1994). Importantly, TAM is not dependent on efficient translation of mRNA (Mokkapati and Bhagwat, 2002). Evidence favoring the inducible (adaptive) evolutionary paradigm over neutrality has been accumulated for bacterial (Ponder et al., 2005; Cirz and Romesberg, 2007), plant (Galloway and Etterson, 2007) and metazoan genomes (Levasseur et al., 2007) and include faster-than-expected rates of phenotype acquisition, close temporal correlation with environmental changes and proof of improved fitness.

12.2 Non-random recombination

Increased recombination has been observed in response to stress (fitness-associated recombination) (Plough, 1917; Grell, 1971; Zhuchenko et al., 1986; Parsons, 1988; Gessler and Xu, 2000; Hadany and Beker, 2003a; b; Schoustra et al., 2010; Zhong and Priest, 2011),including genetic stress (Tedman-Aucoin and Agrawal, 2012; Stevison, 2012). Mitotic recombination might itself be mutagenic owing to errors introduced during double-strand break repair (Lercher and Hurst, 2002). Indeed, the rate of recombination and the rate of substitutions are positively correlated (Lercher and Hurst, 2002). A number of studies have observed that recombination rates are positively correlated with extant levels of genetic variation in a wide variety of organisms, including Drosophila and humans (Aguade et al., 1989; Begun and Aquadro, 1992; Nachman et al., 1998; Stephan and Langley, 1998; Hamblin and Aquadro, 1999; Hardison et al., 2003; Hellmann et al., 2003; Lercher and Hurst, 2003; Roselius et al., 2005; Shapiro et al., 2007; Noor, 2008). Recombination can also lead to gross chromosomal rearrangements (Lambert et al., 2005). Thus mutation and recombination rates involve different kinds of genetic alterations, with mutation conferring gene-level changes, and recombination generating genome-level alterations (Heng, 2009). The consequences of their effect on the genome are radically different, but the selective process acting on the genetic diversity that they generate is the same. Hence, the rate of genetic-variant generation was defined as the joint sum of mutation and recombination rates (Capp, 2010).

Meiotic recombination events are not randomly distributed along the chromosomes, but occur at specialized sites, 1-2 kb long, called hotspots (Steinmetz et al., 1982; Petes, 2001; Pevzner and Tesler, 2003a, b; Kent et al., 2003; Bourque et al., 2004; McVean et al., 2004; Myers et al., 2005; Arnheim et al., 2007). In the human genome there is evidence for extreme local recombination rate variation spanning four orders in magnitude, in which 60% of all recombination events take place in about 6% of the sequence (McVean et al., 2004; The International HapMap Consortium, 2007). Recombination hotspots are a ubiquitous feature of the human genome, occurring on average every 200 kilobases or less, but recombinationoccurs preferentially outside genes (McVean et al., 2004). Gene functions associated with cell surfaces and external functions tend to show higher recombination rates (immunity, cell adhesion, extracellular matrix, ion channels, signalling) whereas those with lower recombination rates are typically internal to cells (chaperones, ligase, isomerase, synthase) (The International HapMap Consortium, 2007). Human hotspots are enriched for a degenerate 13 base pair (bp) motif, suggesting that underlying DNA sequence is a major determinant in hotspot specification (Myers et al., 2008). In yeast and mice, meiosis requires topoisomerase (Spo11)-induced DSBs for initiation of genetic crossovers (Keeney, 2007). Histone methylation marks define the vast majority of mammalian recombination hotspots (Baudat et al., 2010; Grey et al., 2011; Ségurel et al., 2011; Smagulova et al., 2011). A zinc finger protein with histone H3 lysine 4 methyltransferase activity, PRDM9 (Hayashi et al., 2005), is thought to bind to hotspot recognition sites in DNA and trimethylate nearby histones (H3K4me3), thereby activating chromatin and attracting the topoisomerase SPO11, which then catalyzes the initial DSBs. H3K4me3 is a prominent and preexisting mark in budding yeast and mammals for formation of the programmed DSBs (Borde et al., 2009; Buard et al., 2009; Kniewel and Keeney, 2009). Allelic variants of PRDM9 zinc fingers are significantly associated with variability in genome-wide hotspot usage among mammals (Baudat et al., 2010; Myers et al., 2010; Parvanov et al., 2010). However, the spatial correlation between H3K4me3 and DSB hotspots may be attributable to coincident localization of both to gene promoters (Buard et al., 2009; Tischfield and Keeney, 2012). Hotspots are still observed inPrdm9 knockout mice, and as in wild type, these hotspots are found at H3K4 trimethylation marks. Prdm9 knockout mice are proficient at initiating recombination (Hayashi et al., 2005). However, in the absence of PRDM9, most recombination is initiated at promoters and at other sites of PRDM9-independent H3K4 trimethylation. Such sites are rarely targeted in wild-type mice, indicating a role of the PRDM9 protein in sequestering the recombination machinery away from gene-promoter regions and other functional genomic elements (Brick et al., 2012). Recently published surveys of the rapidly evolving sequences of PRDM9 zinc fingers from multiple species of mammals and diverse metazoans (Oliver et al., 2009; Thomas et al., 2009) suggest that PRDM9 may have been involved in hybrid sterility and speciation (Mihola et al., 2009; Sandovici and Sapienza, 2010) in multiple lineages on many occasions. In human peripheral lymphocyte metaphase, chromosomal regions with a high density of 8-oxodG, i.e. a higher susceptibility to oxidative damage, coincide with regions exhibiting a high meiotic recombination rate as well as with those with a high density of SNPs (Ohno et al., 2006).

 12.3 Mutability of simple sequence repeats

Genomic microsatellites (simple sequence repeats; SSRs), iterations of 1 to 6 bp nucleotide motifs, have been detected in the genomes of every organism analysed so far, and are often found at frequencies much higher than would be predicted purely on the grounds of base composition (Tautz and Renz, 1984; Epplen et al., 1993; Toth et al., 2000). The number of repeated units can range anywhere from a few to hundreds or even thousands of copies. Estimates from the human genome reference sequence indicate that microsatellites may account for ~3% of the genome. This contribution, however, is highly approximate and depends strongly on how repeat length and sequence purity thresholds are defined (Kozlowski et al., 2010). On the other hand, approximately 50% of the human genome is occupied by repetitive sequences, including TEs (Jasinska and Krzyzosiak, 2004; Richard et al., 2008). Numerous lines of evidence have demonstrated that genomic distribution of SSRs is nonrandom, presumably due to strong selective pressures because of their effects on chromatin organization, regulation of gene activity, recombination, DNA replication, cell cycle, translation, mismatch repair (MMR) system, etc. (Li YC et al., 2002; 2004). Evolutionarily conserved SSRs are abundant in promoters in both the yeast and human genome (Vinces et al., 2009; Sawaya et al., 2013). On the other hand, numerous lines of evidence show that SSRs located in promoter regions and intronic regions can affect gene activity (Hoffman et al., 1990; Punt et al., 1990; Lafyatis et al., 1991; Sandaltzopoulos et al., 1995; Chen and Roxby, 1997; Meloni et al., 1998; Gebhardt et al.; 1999; 2000; Vinces et al., 2009; Sawaya et al., 2013). SSR repeat number appears to be a key factor for gene expression and expression level (Hamada et al. 1984; Chamberlain et al., 1994; Lanz et al., 1995; Miret et al., 1998; Okladnova et al., 1998; Xu and Goodridge, 1998; Gebhardt et al., 1999; 2000; Liu L et al., 2000b; Li YC et al., 2002; 2004).

An immanent feature of microsatellites is their high mutability, which leads to both sequence and length polymorphism (Kelkar et al., 2008; Madsen et al., 2008; Pumpernik et al., 2008), the latter being at least one order of magnitude greater than the former (Borstnik and Pumpernik, 2002; Pumpernik et al., 2008). The SSR mutation rates (10-2 to 10-6 events per locus per generation) are very high, as compared with the rates of point mutation at coding gene loci (Li YC et al., 2002; Ellegren, 2004; Payseur et al., 2011; Sun et al., 2012). SSRs encode their own mutability through the unit size, length, and purity of the repeat tract (Ellegren, 2004; King and Kashi, 2007; Legendre et al., 2007). For example, repeats consisting of short units are much more common and less stable (Schug et al., 1998) than repeats of longer units. Mutational mechanisms include DNA slippage during DNA replication (Tachida and Iizuka, 1992), recombination between DNA strands (Harding et al., 1992), or an interaction of both (Li YC et al., 2002). Generally, all DNA metabolic processes leading to transient separation of DNA strands such as replication, recombination, repair and transcription have stimulatory effects on repeat instability (Pearson et al., 2005; Wells et al., 2005; Wells and Ashizawa, 2006; López Castel et al., 2010; McIvor et al., 2010). Several lines of evidence, accumulated from prokaryotic, yeast, mammalian cell culture and animal model studies indicate that transcription through repeating sequences is an important factor promoting their instability (Wierdl et al., 1996; Bowater et al., 1997; Mangiarini et al., 1997; Mochmann and Wells, 2004; Lin et al., 2006; 2009; Lin and Wilson, 2007; Soragni et al., 2008; Ditch et al., 2009; McIvor et al., 2010). The persistent interaction between transcription template DNA and nascent RNA (RNA•DNA hybrids, R loops) was shown to stimulate genomic instability, identifying R loops as a common mutagenic conformation (Lin Y et al., 2010; McIvor et al., 2010). Although the mutation process seems to display distinct differences among species, repeat types, loci and alleles, age and sex (Brock et al., 1999; Hancock, 1999; Ellegren, 2000; 2004; Schlötterer, 2000), the instability manifests predominantly as changes in the number of SSR repeats. DNA damage caused by external stresses such as UV irradiation, gamma–irradiation, t-butyl hydrogen peroxide, oxidative damage, etc. can induce slippage mutations and increase genetic instability in SSR sequences (Jackson et al., 1998; Jackson and Loeb, 2000; Chang et al., 2002; Slebos et al., 2002; Fonville et al., 2011). Stress-dependent SSR instability may be modulated by Hsp90 (Mittelman et al., 2010; Mittelman and Wilson, 2010; Fonville et al., 2011). DNA methylation in mammals is associated with repeat stability: demethylation of minor satellites, subtelomeric satellites, microsatellites and selfish repeats appears to lead to increased recombination and mutagenesis and may result in destabilisation of the chromosome on which they reside (Hansen et al., 1999; Okano et al., 1999; Xu et al., 1999; Bourc’his and Bestor, 2004; Guo G et al., 2004; Kazazian, 2004; Kim M et al., 2004; Wang D et al., 2004; Gonzalo et al., 2006; Lees-Murdock and Walsh, 2008). Assaying microsatellite instabilities, a common phenotype associated with mutations affecting DNA replication fidelity/processivity or the MMR functions (Sia et al., 1997; Prolla et al., 1998), in mouse embryonic stem cells with homozygous deletion of the Dnmt1 gene and consequently a hypomethylated genome, the Dnmt1 null embryonic stem cells exhibited high microsatellite instability frequencies (4 to 7-fold increase in the microsatellite slippage rate) in comparison to the wild-type embryonic stem cells (Guo G et al., 2004; Kim M et al., 2004; Wang and Shen, 2004).

Emerging evidence implicates coding and noncoding SSRs as important sources of common, subclinical genetic variation in morphological and behavioral traits in numerous species, including humans, dogs, and flies (Goodman et al., 1997; Sawyer et al., 1997; Fondon and Garner, 2004; Kashi and King, 2006; Mittelman and Wilson, 2010). Coding microsatellites are enriched in genes for transcription factors and other regulatory proteins, and changes in the length of these repeats exert incremental impacts on gene function (Gerber et al., 1994; Wren et al., 2000; Albrecht et al., 2004; Bacolla et al., 2008). Variations in the lengths of noncoding repeats in the promoters of genes have been shown to quantitatively affect transcription and have likely facilitated transcriptional evolution (Vinces et al., 2009). SSRs located in the promoter, intronic, untranslated and coding regions of stress response genes may modulate the evolvability of organisms. In 296 Escherichia coli genes related to repair, recombination and physiological adaptations to different stresses, Rocha et al. (2002) observed a significant high number of short close repeats capable of inducing phenotypic variability by slipped-mispair during DNA, RNA or protein synthesis. The authors suggested that overrepresentation of repeats in stress response genes may be a bacterial strategy to increase versatility under stressful conditions. In haemophilus influenzae, tandem oligonucleotide repeats are far from being randomly distributed in the genome but are substantially overrepresented in genes encoding virulence determinants (Hood et al., 1996) potentially conferring an adaptive advantage in changing host environments. SSRs in promoter regions may serve as transcriptional elements for heat-shock protein genes in Drosophila (Sandaltzopoulos et al., 1995), Aspergillus (Punt et al., 1990), and Phytophthora (Chen and Roxby, 1997). Based on evidence in a variety of taxa from virus and yeast to humans (Treco and Arnheim 1986; Wahls and Moore 1990a; b; Aharoni et al., 1993; Majewski and Ott, 2000; Templeton et al., 2000), SSRs have been proposed as hot spots for recombination (Jeffreys et al., 1998; Templeton et al., 2000). Thus, SSRs as evolutionary tuning knobs may provide an evolutionary advantage of fast adaptation to new environments (Kashi et al., 1997; King et al., 1997; Trifonov, 2003; Kashi and King, 2006; Schmidt and Anderson, 2006; Fondon et al., 2008; Vinces et al., 2009; Gemayel et al., 2010; King, 2012b).

The effectiveness of the MMR system is critical to SSR stability. DNA methylation appears to be closely linked to MMR to maintain SSR (Kim M et al., 2004; Wang and Shen, 2004). In prokaryotic and eukaryotic cells, impaired function of MMR enzymes increases the rate of SSR mutation rates by several orders of magnitude (Levinson and Gutman, 1987a; Strand et al., 1993; Jiricny, 1994; Richards and Sutherland, 1994; Tautz and Schlötterer, 1994; Sia et al., 1997; Strauss et al., 1997; Tran et al., 1997). It is a paradox that the MMR system, which limits mutation in SSR sequences, is particularly vulnerable to mutation by virtue of having SSRs in its own coding regions (Chang et al., 2001). All “minor” components of the human DNA MMR system—MSH3, MSH6, PMS2, and MLH3—contain mononucleotide microsatellites in their coding sequences. The loss of activity in any one of these minor proteins generates a weaker mutator phenotype than occurs with loss of the major MMR proteins, MSH2 and MLH1 (Chang et al., 2001). A study of 10,000 random primate (primarily human) coding sequences (8.6 million base pairs) showed that the average-sized sequence coding for a human gene is expected to contain approximately 0.1, 0.03, and 0.006 mononucleotide runs of length 7 bp or more, 8 bp or more, and 9 bp or more, respectively (Metzgar et al., 2000). Together, the four minor MMR genes contain three 7-bp runs, four 8-bp runs, and one 9-bp run. Controlling for gene length, the probability of finding this many mononucleotide microsatellites by chance in a group of four genes is 1.3 x 10-6 (Chang et al., 2001). The constellation of microsatellites in the coding regions of the minor MMR genes is a general feature among eukaryotes (Chang et al., 2001). SSR instability in humans is associated with such DNA MMR genes as MLH1 (Boyer et al., 1995), MSH2, MSH3, and MSH6 (Boyer et al., 1995; Clark et al., 1999). When these genes mutate or become defective, SSR instability consequently increases. Mononucleotide repeats have a strong influence on the local mutation rate (Levinson and Gutman, 1987b). In yeast, extending a mononucleotide repeat (of length 4) by one nucleotide leads to an increase in the local mutation rate by about a factor of two (Tran et al., 1997). Nontriplet microsatellites, when located in coding sequences, are expected to introduce frameshift loss-of-function mutations at high frequency (Moxon et al. 1994). The authors hypothesized that the exceptional density of microsatellites in the minor MMR genes represents a genetic switch that allows the adaptive mutation rate to be modulated over evolutionary time (Chang et al., 2001).

Trinucleotide repeats (TNRs), whose numbers do not affect the reading frame, tend to be overrepresented in exons (Richard et al., 2008). The relative abundance of TNRs in coding regions and their selective enrichment in genes for transcription factors and other regulatory proteins suggest that TNRs provide a positive evolutionary benefit (Albrecht and Mundlos, 2005; Bacolla et al., 2008; Mittelman and Wilson, 2010). The features of TNRs that suggest their functionality include: (i) widespread occurrence in exons, (ii) formation of stable hairpin or quadruplex structures by some TNRs and (iii) coding for homo-amino acid tracts (Kozlowski et al., 2010). However, at more than 20 loci in the human genome, the lengthening of a TNR tract beyond some threshold causes a neurodegenerative or neuromuscular disorder (Albrecht and Mundlos 2005; Orr and Zoghbi, 2007; Bacolla and Wells, 2009; Orr, 2009; La Spada and Taylor, 2010; López Castel et al., 2010). More than 20 years agoo, increase in length of TNRs was found to be a mutation responsible for spinobulbar muscular atrophy, fragile X syndrome and myotonic dystrophy type 1 (Fu et al., 1991; La Spada et al., 1991; Brook et al., 1991; Mahadevan et al., 1992; Tsilfidis et al., 1992). Since then, more than 20 human syndromes, including Huntington disease, many of the inherited ataxias, and the muscular dystrophies, as well as pathologies in plants and animals, have been attributed to repeat expansions (Wells and Ashizawa, 2006). These unusual mutations, termed dynamic mutations due to their unstable, progressive character and non-Mendelian inheritance occur predominantly at GC-rich repeat sequences (Pearson et al., 2005). Long TNRs tend to become unstable during their transmission to offspring, giving rise to progeny with longer repeat tracts (expansions) or shorter ones (contractions). The typical bias toward expansions leads to progressively more severe disease symptoms in the progeny in each subsequent generation. The extraordinarily high mutation frequency levels of Huntington's disease germline mutations are most consistent with a mutation process that occurs throughout germline mitotic divisions, rather than resulting from a single meiotic event (Leeflang et al., 1999).

12.4 Activity of transposable elements

Barbara McClintock, who is best known for her discovery of transposable elements and their mobilization under stress, was later primarily interested in the sensory and decision-making (that is, cognitive) capacities of cells with damaged genomes (Shapiro, 2010) and concluded her Nobel Prize lecture (1984): “In the future, attention undoubtedly will be centred on the genome, with greater appreciation of its significance as a highly sensitive organ of the cell that monitors genomic activities and corrects common errors, senses unusual and unexpected events, and responds to them, often by restructuring the genome”.

In species such as Drosophila, TEs comprise approximately 10% of heterochromatic (Bartolome et al., 2002) and only 2%–3% of euchromatic DNA (Bartolome et al., 2002; Kaminker et al., 2002) but cause over 50% of de novo mutations (Eickbush and Furano, 2002). While insertions of TEs are responsible for only ~0.1% of de novo mutations in humans, the figure is 100-fold higher in the laboratory mouse (Maksakova et al., 2006). The prevalence of mouse TE activity indicates that the genome of the laboratory mouse is presently behind in the ‘‘arms race’’ against invasion (Maksakova et al., 2006). Throughout phylogenesis, a multitude of abiotic and biotic stressors disrupt epigenetic regulation and mobilize TE (McClintock, 1984; Paquin and Williamson, 1984; Strand and McDonald, 1985; Rolfe and Banks, 1986; Bradshaw and McEntee, 1989; Ratner et al., 1992; Arnault and Dufournel, 1994; Pouteau et al., 1994; Wessler, 1996; Grandbastien, 1998; Capy et al., 2000; Morillon et al., 2000; Ikeda et al., 2001; Chen et al., 2003a; Daboussi and Capy, 2003; Lu and Ramos, 2003; Jorgensen, 2004; Pesheva et al., 2005; 2008; Todeschini et al., 2005; Farkash et al., 2006; Ilves, 2006; McGraw and Brookfield, 2006; de la Vega et al., 2007; Stoycheva et al., 2007b; 2010; Bouvet et al., 2008; Cam et al., 2008; Desalvo et al., 2008; Perez-Hormaeche et al., 2008; Oliver and Greene, 2009a; Zeh et al., 2009; Zhang and Saier, 2009; 2011; Rebollo et al., 2010; 2012; Casacuberta and González, 2013; Feng et al., 2013). Activation of TEs is not always directly triggered by a specific stress but the effects that such stress causes in other cellular mechanisms may allow a rapid activation of some particular TE copies (Dai et al., 2007; Coros et al., 2009). Noncoding and small interfering RNAs are another possible path by which TEs respond to stress (Hilbricht et al., 2008; Mariner et al., 2008; Lv et al., 2010; Yan et al., 2011; McCue et al., 2012). Alu elements function as cell stress genes: different stress conditions cause an increase in the expression of Alu RNAs, which rapidly decreases upon recovery from stress (Häsler and Strub, 2006). Alu RNA has been implicated in regulating several aspects of gene expression such as alternative splicing, RNA editing, translation and miRNA expression and function (Häsler and Strub, 2006; Häsler et al., 2007). The evidence that only some specific TE families, and not all the TEs in the genome, are activated in response to stress and that these TEs respond to some specific stress conditions and not others, strongly suggest that activation of TEs by stress is not only a byproduct of genome deregulation. The consequences of TE activation in response to stress are diverse (Casacuberta and González, 2013). Stress-activated TEs: (i) contribute to major genomic rearrangements (Maumus et al., 2009), (ii) confer nearby genes the capacity to respond to stress (Guo and Levin, 2010; Servant et al., 2012), which may lead to the creation of new regulatory networks (Naito et al., 2009; Ito et al., 2011) and (iii) alter the genome randomly through insertion of the newly generated copies (Dai et al., 2007). Recently, it was found that Tf1, a long-terminal repeat retrotransposon in Schizosaccharomyces pombe, was activated by heat treatment and preferentially integrated into promoters and activated the expression of genes that themselves were induced by heat (Feng et al., 2013). The authors suggested that the integration preference of Tf1 for the promoters of stress response genes and the ability of Tf1 to enhance the expression of these genes co-evolved to promote the survival of cells under stress (Feng et al., 2013). Inserts of Tf1 can enhance the transcription of adjacent genes only if the genes have stress response promoters and are activated by the stressor and/or oxidative stress (Feng et al., 2013). These findings corroborate earlier evidence that TE insertion may be non-random and may be directed to target sites by various specific factors including transcription activators (Devine and Boeke, 1996; Singleton and Levin, 2002; Leem et al., 2008; Zhang and Saier, 2009; 2011; Majumdar et al., 2011; Mularoni et al., 2012).

Integration of TEs has the potential to wreak havoc by destroying coding sequences throughout the genome of the host cell. On the other hand, there is abundant evidence for the adaptive effects of TE activation events in bacteria (Schmidt and Anderson, 2006; Chou et al., 2009; Stoebel et al., 2009; Sun J et al., 2009b; Zhang and Saier, 2009; 2011; Stoebel and Dorman, 2010; Gaffé et al., 2011), plants (Yan L et al., 2006; Lin et al., 2007; Liu et al., 2008; Kanazawa et al., 2009; Chu et al., 2011), and animals (Schlenke and Begun, 2004; Aminetzach et al., 2005; Biemont and Vieira, 2006; Darboux et al., 2007; Santangelo et al., 2007; González et al., 2008; 2009; 2010; González and Petrov, 2009; Naito et al., 2009; Schmidt et al., 2010; Franchini et al., 2011; Magwire et al., 2011; Casacuberta and González, 2013). The particular cases described may represent only the tip of the iceberg (Casacuberta and González, 2013). TEs can have a myriad of effects when they insert into new locations (Feschotte, 2008; Goodier and Kazazian, 2008; Gogvadze and Buzdin, 2009; Casacuberta and González, 2013) and can result in the acceleration of the evolution of genes in a myriad of ways providing a means for rapid species divergences in the affected lineages (Brookfield, 2004; Schmidt and Anderson, 2006; Böhne et al., 2008; Goodier and Kazazian, 2008; Oliver and Greene, 2009a; b; 2011; 2012; Rebollo et al., 2010; Hua-Van et al., 2011).

Some TEs display considerable target site selectivity while others display little obvious selectivity, although none appear to be truly “random.” A variety of mechanisms for target site selection are used: (i) direct interactions between the recombinase and target DNA; (ii) interactions with accessory proteins that communicate both with the target DNA and the recombinase (Craig, 1997; Liao et al., 2000). TEs and host regulatory factors have coevolved in an arms race scenario (McDonald, 1998; Aravin et al., 2007b; Rebollo et al., 2012). A large body of evidence suggests that TEs play a significant role in the host response to an everchanging environment, both in prokaryote and in eukaryote organisms (Rebollo et al., 2010; 2012; Casacuberta and González, 2013). Bioinformatic surveys of TE-gene chimeric transcripts suggest that many gene promoters or alternative promoters are, in fact, derived from TEs (van de Lagemaat et al., 2003; Romanish et al., 2007; Conley et al., 2008; Feschotte, 2008; Faulkner et al., 2009; Rebollo et al., 2012). For example, in mouse and human, 18.1% and 31.4% of total transcription start sites, respectively, are located within TE sequences and are often found to be tissue specific (Faulkner et al., 2009). The preferential insertion of some TEs near genes may facilitate the exaptation of TE sequences as gene regulators (Feschotte and Pritham, 2007). It was postulated that a large fraction of gene promoters and enhancers were donated by ancient TEs, which may retain their cis-regulatory motifs but are so extensively degraded in overall structure that they are no longer recognizable as such (Rebollo et al., 2012). Domestication of TEs is also important in the evolution of genomes, such as the evolution of new protein-coding genes (including regulatory DNA binding factors), and regulatory small RNAs from TEs (Feschotte and Pritham, 2007; Sinzelle et al., 2009; Werren, 2011). TEs are also passive generators of mutations. TEs that belong to the same family of elements and are located in different regions of the genome can act as substrates for ectopic recombination events generating rearrangements such as inversions, translocations or duplications (Schwartz et al., 1998; Hill et al., 2000; Bailey et al., 2003).

TEs can also act as vectors facilitating the horizontal transfer of new genetic content (HGT) (Ochman et al., 2000; Frost et al., 2005). This phenomenon has been extensively demonstrated in prokaryotes. As in prokaryotes, HGT has had an important role in eukaryote genome evolution (Keeling and Palmer, 2008; Syvanen, 2012). The evidence for HGT in diverse eukaryotes is expanding rapidly in organisms such as protists (Loftus et al., 2005), fungi (Fitzpatrick, 2012), nematodes (Haegeman et al., 2011) and in fish (Graham et al., 2008). Horizontal transfer of TEs has been reported in eukaryotic species as diverse as tetrapods, Drosophila, yeast and fungi (Anxolabéhère et al., 1988; Daniels et al., 1990; Maruyama and Hartl, 1991; Robertson and Lampe, 1995; Kordiš and Gubenšek, 1998; Gogolevsky et al., 2008; Lampe et al., 2003; Herédia et al., 2004; Hall et al., 2005; Pritham and Feschotte, 2007; Ray et al., 2007; Loreto et al., 2008; Pace et al., 2008; Gilbert et al., 2010; Schaack et al., 2010a). In eukaryotes, although TEs are capable of capturing and transferring genes at a high frequency within a species (Jiang et al., 2004; Morgante et al., 2005; Schaack et al., 2010a) they have not yet been found to transfer host genes between different species (Schaack et al., 2010a) although some authors predict that it will be soon discovered (Keeling and Palmer, 2008).

12.5 Phenotypic plasticity, genetic assimilation and accommodation

JM Baldwin (1896) noted: “The creatures which can stand the ‘storm and stress’ of the physical influences of the environment, and of the changes which occur in the environment, by undergoing modifications of their congenital functions or of the structures which they get congenitally—these creatures will live; while those which cannot, will not.”

In nature, where environmental conditions are perpetually variable, organisms face the challenge of maximizing fitness under heterogeneous conditions. Selection has solved this problem in numerous ways including environmental canalization, which reduces environmental influence on trait expression (see chapter 10.4), and phenotypic plasticity, where a single genotype can produce multiple phenotypes under different environmental conditions. When a trait is canalized, it may be well-adapted to one environment, but when a trait is plastic it may be well-adapted to many environments (Bradshaw, 1965; Pigliucci, 2001; Ghalambor et al., 2007).

Development is taken fairly broadly as the genotype-phenotype map (Lewontin, 1974; Wagner and Altenberg, 1996; Fontana and Schuster, 1998). Phenotypic plasticity is usually defined as a property of individual genotypes to produce different phenotypes when exposed to different environmental conditions (Pigliucci, 2001; West-Eberhard, 2003; Pigliucci et al., 2006). Phenotypic plasticity can be illustrated with the concept of a reaction norm, which is the representation of values that a trait takes across an environmental gradient (Woltereck, 1909; Waddington, 1942; de Jong, 1990; Weis and Gorman, 1990; Gomulkiewicz and Kirkpatrick, 1992; Via et al., 1995; Flatt, 2005; Aubin-Horth and Renn, 2009). Different genotypes may have different reaction norms, meaning that they respond differently to the same environmental conditions. Indeed, it is possible that one genotype responds to a change in an environmental variable while another does not change, or changes in another direction (Côté et al., 2007). The presence of genetic variation for a trait, along with the force of selection, determines whether selection acting on that trait will result in an evolutionary response. Therefore, if different genotypes in a population produce different reaction norms, and if the slope of the reaction norm is positively correlated with fitness, increased plasticity should evolve in that population for that particular trait (Schlichting and Pigliucci, 1998; Aubin-Horth and Renn, 2009).

Early on Wright (1931) pointed out that plasticity could also inhibit genetic divergence. If a species is plastic such that its phenotype matches the optimum in all environments, there would be no impetus for further evolution, an anti-Baldwin effect (Scheiner and Holt, 2012). Thus, the role of phenotypic plasticity in evolution has historically been a contentious issue because of debate over whether plasticity shields genotypes from selection (e.g. Grant, 1977; Levin, 1988; Falconer, 1989) or generates novel opportunities for selection to act (Robinson and Dukas, 1999; Pigliucci and Murren 2003; Price et al., 2003; West-Eberhard, 2003; Schlichting, 2004; Badyaev, 2005a; Ghalambor et al., 2007). Cumulative evidence suggests that both sides of the coin reflect evolutionary reality. Theoretical models for the evolution of adaptive phenotypic plasticity predict that given genetic variation, selection will favor adaptive plasticity when: (i) populations are exposed to variable environments, (ii) environments produce reliable cues, (iii) selection favors different phenotypes in each environment, and (iv) no single phenotype exhibits superior fitness across all environments (Bradshaw, 1965; Levins, 1968; Via and Lande, 1985; Lively 1986; Gomulkiewicz and Kirkpatrick, 1992; Moran, 1992; Ghalambor et al., 2007). To a certain extent, phenotypic plasticity is at the level of phenotype–namely disturbance-related variation as bet-hedging strategy (Gabriel, 2005; Scheiner and Holt, 2012)–what stress-induced mutagenesis is at the level of genotype. This interpretation also implies that phenotypic plasticity is not necessarily adaptive (Padilla and Adolph, 1996; Ancel, 2000; Ghalambor et al., 2007) due to a considerable element of chance (Price et al., 2003) (like mutations that are rather deleterious than beneficial, see chapter 5.1). Moreover, adaptive plasticity or learning may allow genetically unfit individuals to compensate for the inadequacies of their genotypes and still attain high fitness. As a consequence, the genetic variation for fitness may become reduced and the response to selection may be slowed down (Johnston, 1982; Gordon, 1992; Papaj, 1994; Behera and Nanjundiah, 1995; Mayley, 1997; Ancel, 2000; Huey et al., 2003). However, in conjunction with natural selection in temporally and spatially heterogeneous environments, phenotypic plasticity is adaptive (Bradshaw, 1965; Levins, 1968; Provine, 1971; Schlichting, 1986; Scheiner and Lyman, 1989; de Jong, 1990; 1995; Brönmark and Miner, 1992; Newman, 1992; Scheiner, 1993; Day et al., 1994; Gotthard and Nylin, 1995; Robinson and Wilson, 1996; Blanckenhorn, 1998; Schlichting and Pigliucci, 1998; Agrawal AA, 2001; Pigliucci, 2001; Laurila et al., 2002; DeWitt and Scheiner, 2004; Laforsch and Tollrian 2004; Yeh and Price, 2004; Svanbäck and Eklöv, 2006; Latta et al., 2007; Rando and Verstrepen, 2007;Van Buskirk and Relyea, 2008; Beldade et al., 2011; Snell-Rood, 2012).

Like learning capacity and memory (Mery and Kawecki, 2002; 2003; 2005), phenotypic plasticity has its costs and limits (DeWitt et al., 1998; Reylea, 2002; Merilä et al., 2004; van Kleunen and Fischer, 2007; Van Buskirk and Steiner, 2009; Auld et al., 2010; Snell-Rood et al., 2010; Snell-Rood, 2012). In a meta-analysis of 27 studies (of 16 plant and 7 animal species) that measured selection on the degree of plasticity, plasticity tended to be more costly in treatments with greater stress but this was only true in studies of animals indicating that ecological costs increase in the context of competition, predation risk or resource limitation (Van Buskirk and Steiner, 2009). Because plasticity is a trait which is expressed across environments, hard and soft selection differentially favor plasticity and genetic specialization, respectively (Van Tienderen, 1991, 1997). If the environment changes only very slowly relative to the generation time of the organism, then genetic specialization is favored over plasticity (Orzack, 1985). Obviously, for learning to be useful the environment must not change too fast relative to how fast the animal can learn (Dukas, 1998; Stomp et al., 2008). Since learning is expensive (Mery and Kawecki, 2002; 2003; 2005), in relatively stable environments there is rather a selective pressure for the evolution of instinctive behaviors (Turney, 1996).

Developmental plasticity is linked to epigenetic mechanisms (Bachmann, 1983; Aubin-Horth and Renn, 2009; Bonduriansky and Day, 2009; Beldade et al., 2011; Hallgrimsson and Hall, 2011; Valena and Moczek, 2012). A central question in considering evolutionary change in response to environmental change is whether the inherited epigenetic markers could facilitate genomic change (Johnson and Tricker, 2010; Bateson, 2012). Pál and Miklós (1999) and Price et al. (2003) modeled how phenotypic plasticity can allow for diversification by allowing for individuals to move among peaks on a fitness landscape. In a fitness landscape with multiple peaks and valleys, developmental variability can smooth the adaptive landscape to provide a directly increasing path of fitness to the highest peak. Thus, epigenetic inheritance and developmental variability allow initial survival of a genotype in response to novel or extreme environmental challenge, providing an opportunity for subsequent adaptation. This initial survival advantage arises from the way in which epigenetic inheritance and developmental variability smooth and broaden the landscape that relates genotype to fitness (Pál and Miklós, 1999; Frank, 2011).

The Baldwin effect, an interaction between non-heritable lifetime plasticity (e.g. learning) and evolution, has been shown to be able to guide evolutionary change and ‘smooth out’ abrupt fitness changes in fitness landscapes – thus enabling genetic evolution that would otherwise not occur (Borenstein et al., 2006; Mills and Watson, 2006; Frank, 2011). Phenotypic plasticity is an important mechanism by which populations can respond rapidly to changes in ecological conditions (Scheiner, 1993; Via et al., 1995; Pigliucci, 2001; Latta et al., 2007). Hinton and Nowlan (1987) demonstrated that phenotypic plasticity (specifically, lifetime learning by individuals) can enable an evolutionary system to find optima that would be very difficult to find without plasticity. Thus, lifetime learning in individuals can, in some situations, accelerate evolution by allowing individuals to exploit a novel environment if they possess the capacity to develop a new phenotype when faced with that novel environment (Price et al., 2003; Yeh and Price, 2004). Thus, adaptive plasticity that places populations close enough to a new phenotypic optimum for directional selection to act can enhance fitness and is most likely to facilitate adaptive evolution in new and variable environments (Hinton and Nowlan, 1987; de Jong, 1995; Ancel, 1999; Robinson and Dukas, 1999; Pigliucci and Murren, 2003; Price et al., 2003; West-Eberhard, 2003; Behera and Nanjundiah, 2004; Schlichting, 2004; Yeh and Price, 2004; Badyaev, 2005a; 2009; Ghalambor et al., 2007; Moczek et al., 2011). Implicit in this hypothesis is the idea that genetic variation for plasticity exists (Khan et al., 1976; Jain, 1978; Newman, 1994; Stinchcombe et al., 2004) and plastic individuals will be favored over nonplastic ones in changing environments by natural selection (Nussey et al., 2005).

Lamm and Jablonka (2008) argued that the very same mechanisms may be involved in both plasticity and evolvability. Epigenetic inheritance leads to transgenerationally extended plasticity, and developmentally-induced heritable epigenetic variations provide additional foci for selection that can lead to evolutionary change. The epigenetic mechanisms involved in repatterning can be activated by both environmental and genomic stress, and lead to phylogenetic as well as ontogenetic changes. Hence, the effects and the mechanisms of plasticity directly contribute to evolvability (Lamm and Jablonka, 2008). Heritable plasticity can also respond to (1) artificial selection in which the experimenter selects on a focal trait or trait index, and (2) quasi-natural selection in which the experimenter establishes a set of environmental conditions and then allows the population to evolve (Scheiner, 1993; 2002). Rather than plasticity per se being adaptive, it acts as a form of bet-hedging by producing increased phenotypic variation among offspring (Scheiner and Holt, 2012). Species invading novel or extreme environments often, but not always, display increased plasticity compared to populations from the native range (Chapman et al., 2000; Lee CE et al., 2003; Dybdahl and Kane, 2005; Chun et al., 2007; Cano et al., 2008; Lardies and Bozinovic, 2008; Lombaert et al., 2008). Intriguingly, phenotypic plasticity evolved experimentally in Pseudomonas fluorescens cultures as bet-hedging strategy in a fluctuating environment (Beaumont et al., 2009).

12.5.1 Assimilation and accommodation

Eventually, the new phenotype may become fixed through genetic assimilation, or there may be an evolutionary shift in the elevation or the slope of the reaction norm, a process referred to as genetic accommodation (West-Eberhard 2003; Pigliucci et al., 2006; Suzuki and Nijhout, 2006; Crispo 2007; Aubin-Horth and Renn, 2009). Both Waddington’s assimilation (Waddington, 1953a; b) and genetic accommodation (West-Eberhard, 2003; Braendle and Flatt, 2006) are considered manifestations of the Baldwin effect (Robinson and Dukas, 1999; Crispo, 2007). Phenotypic accommodation is the first step in a process of Darwinian evolution by natural selection, where fitness differences among genetically variable developmental variants cause phenotype-frequency change due to gene frequency change (West-Eberhard, 2005). Given genetic variation in the phenotypic response of different individuals, the initial spread produces a population that is variable in its sensitivity to the new input, and in the form of its response. If the phenotypic variation is associated with variation in reproductive success, natural selection results; and to the degree that the variants acted upon by selection are genetically variable, selection will produce genetic accommodation, a change in gene frequencies under selection (West-Eberhard, 2003; 2005). Genetic assimilation (‘‘phenotype precedes genotype’’) is a common mode of evolution (Waddington, 1961b; Pál and Miklós, 1999; Pigliucci and Murren, 2003; Palmer, 2004; West-Eberhard, 2005). Waddington (1961b) explicitly defined genetic assimilation as ‘‘a process by which characters which were originally ‘acquired characters’ may become converted, by a process of selection acting for several or many generations on the population concerned, into ‘inherited characters.’’ Waddington’s concept of ‘‘acquired characters’’ is clearly equivalent to what we now consider phenotypically plastic traits (Pigliucci and Murren, 2003). During the initial spread, the novel phenotype may increase in frequency rapidly, within a single generation, if it is due to an environmental effect that happens to be common or ubiquitous. In the classical neo-Darwinian genotype-precedes-phenotype mode, mutations initially generate more extreme forms. Alternatively, if it is due to a positively selected mutation, or is a side effect of a trait under positive selection (Müller, 1990), the increase in frequency of the trait may require many generations (West-Eberhard, 2003; 2005). The transgenerational processes that effect assimilation and/or accommodation involve both epigenetic and genetic changes. Within any given individual, a new selective environment is likely to induce a variety of responses in different traits (e.g. Williams et al., 1995; Parsons and Robinson, 2006). Thus, individuals are likely to be made up of both canalized traits that do not respond to novel environmental stimuli (see chapter 10.4) as well as traits that differ in the type of plasticity they exhibit (adaptive and non-adaptive), resulting in individuals that represent a mosaic of traits (Ghalambor et al., 2007).

12.6 Sexual reproduction

Genetic assimilation is the process whereby environmentally induced phenotypic variation becomes constitutively produced (i.e. no longer requires the environmental signal for expression) (Pigliucci et al., 2006). According to this definition, sexual reproduction transformed a process that evolved as a response to environmental disturbance into a proactive bet-hedging tool. Engaging cellular stress for the creation of genetic variability and its purifying selection as a proactive strategy in the face of pervasive environmental change, sexual reproduction can be regarded as the ultimate paragon of “educated guess”.

13. Sexual reproduction: evolvability and robustness


increase of fitness,..., must be brought about largely by changes which increase either genetic stability or variability without bringing about corresponding decrease in the other component. A progressive change is thus one that increases the sum of these components.

J.M. Thoday (1953)

Summary

Genotype (sequence) robustness and evolvability share an antagonistic relationship. On one hand, high robustness implies low production of heritable phenotypic variation. On the other hand, an organism’s capacity to generate heritable phenotypic variation relates to the intrinsic capacity of organisms to evolve. Yet, robustness generally confers evolvability to living systems because it allows them to undergo innovative modification without losing functionality. Estimates of the rates of evolution from contemporary studies range from four to seven orders of magnitude higher than those seen in the fossil record. In particular, there is a striking disparity between spontaneous mutation rates, measured over a small number of generations in studies of pedigrees and the much lower substitution rates measured over geological time frames. Sustained responses to artificial selection do not exhaust genetic variation. The ecological pattern of evolutionary rate supports the notion of a universal dependence of molecular evolution on oxidative stress. A rich body of evidence clearly shows that sex allows for faster adaptation in dynamic environments. Degeneracy, the ability of elements that are structurally different to yield the same output, may act both as a source of robustness and evolvability in biological systems. Sexual reproduction uses several genetic and epigenetic tools that may yield functionally equivalent results. Sexual reproduction succeeded to promote both evolvability and robustness by a process based on oxidative stress: (epi)mutagenesis enhances evolvability while gamete quality control preselects for the most robust and stress resilient gametes.

The ability for an evolving population to adapt to a novel environment is achieved through a balance of robustness and evolvability. Robustness is the invariance of phenotype in the face of perturbation and evolvability is the capacity to adapt in response to selection (McBride et al., 2008). Wagner (2008a) wrote: “Mutational robustness and evolvability, a system’s ability to produce heritable variation, harbour a paradoxical tension. On one hand, high robustness implies low production of heritable phenotypic variation. On the other hand, both experimental and computational analyses of neutral networks indicate that robustness enhances evolvability… [G]enotype (sequence) robustness and evolvability share an antagonistic relationship. In stark contrast, phenotype (structure) robustness promotes structure evolvability. A consequence is that finite populations of sequences with a robust phenotype can access large amounts of phenotypic variation while spreading through a neutral network. Population-level processes and phenotypes rather than individual sequences are key to understand the relationship between robustness and evolvability.”

Degeneracy is the ability of elements that are structurally different to perform the same function or yield the same output. It is a prominent property of gene networks, neural networks, and evolution itself. Edelman and Gally (2001) were the first to propose that degeneracy may act both as a source of robustness and evolvability in biological systems (Whitacre and Bender, 2010). Sexual reproduction uses several genetic and epigenetic tools (see chapters 10 and 12) that may yield functionally equivalent results. Thus, sexual reproduction succeeded to promote both evolvability and robustness by the same process based on oxidative stress: (epi)mutagenesis enhances evolvability while gamete quality control selects for the most robust and stress resilient gametes.

13.1 Evolvability

According to Pigliucci (2008), Flatt (2005) defined evolvability as “the ability of a population to respond to selection,” whereas Griswold (2006) thought in terms of the rate of evolution of a given character (which depends on its heritability, among other things). Houle (1992), however, proposed that the genetic coefficient of variation, rather than heritability, is actually the appropriate quantification of evolvability. Quayle and Bullock (2006) operationally measure evolvability as the time that it takes for a population to hit a given phenotypic target (although this makes the simplifying assumption that the target itself doesn’t shift over time, as a result of environmental changes). Wagner and Altenberg (1996) moved significantly from this ‘classical’ concept of evolvability with their groundbreaking articulation of the distinction between variation and variability as determining evolvability (Pigliucci, 2008). Variation is a measure of the realized differences within a population, whereas variability is the propensity of characters to vary (whether or not they actually do) and depends on the input of new genetic variation through mutation or recombination. For Kirschner and Gerhart (1998) evolvability is an organism’s capacity to generate heritable phenotypic variation that relates to the intrinsic capacity of organisms to evolve (Sniegowski and Murphy, 2006; Pigliucci, 2008).

In her preface to Volume 870 of the Annals of the New York Academy of Sciences: Molecular Strategies in Biological Evolution, Lynn Caporale (1999) wrote: “Most descriptions of mutation have emphasized its negative consequences, and randomness with respect to biological function. This book seeks to balance the discussion by emphasizing mechanisms that both diversify the genome and increase the probability that a genome's descendants will survive… Genomes that encode ‘better’ amino acid sequences are at a selective advantage. Genomes that generate diversity also are at an advantage to the extent that they can navigate efficiently through the space of possible sequence changes.” Volume 870 is a testimony to the manifold “ingenious” processes (e.g. Radman et al., 1999) that organisms evolved to generate genetic diversity as response to environmental challenge, having “learned” since eons that evolution favors the prepared genome.

In his 1905 paper, Einstein proposed that the same random forces that cause the erratic Brownian motion of a particle also underlie the resistance to the macroscopic motion of that particle when a force is applied (Kaneko, 2009; Lehner and Kaneko, 2011). This insight can be generalized to state that the response of a variable to perturbation should be proportional to the fluctuation of that variable in the absence of an applied force (Kubo et al., 1985). In short, the more something varies, the more it will respond to perturbation, irrespective of the precise molecular details. A generalized version of the fluctuation–response relationship can be applied to evolved, dynamical systems (Sato et al., 2003; Kaneko and Furusawa, 2006; Kaneko, 2009). The concept has been confirmed experimentally in unicellular prokaryotes and eukaryotes (Sato et al., 2003; Yomo et al., 2006; Lehner, 2010; Park H et al., 2010). The literature reveals significant effects of genetic diversity on ecological processes such as primary productivity, population recovery from disturbance, interspecific competition, community structure, and fluxes of energy and nutrients. Thus, genetic diversity can have important ecological consequences at the population, community and ecosystem levels, and in some cases the effects are comparable in magnitude to the effects of species diversity (Gamfeldt et al., 2005; Fussmann et al., 2007; Gamfeldt and Kallstrom, 2007; Lankau and Strauss, 2007; Hughes et al., 2008). Moreover, theoretical and empirical studies suggest that diversity at one level may depend on the diversity at the other (Whitham et al., 2003; Abrams, 2006; Crutsinger et al., 2006; Johnson et al., 2006; Vellend, 2006; Lankau and Strauss, 2007).

Many modern theories that attempt to provide an explanation for the advantage of sex incorporate an idea originally proposed by Weismann more than 100 years ago: sex allows natural selection to proceed more effectively because it increases genetic variation through the shuffling of existing genetic material (Weismann, 1889; 1904; Barton and Charlesworth, 1998; Burt, 2000). Fisher’s (1930) fundamental theorem of natural selection holds that the rate of increase in fitness in a population at a time equals the additive genetic variance in fitness at that time.

Mutation is the ultimate source of genetic variation. A large fraction of mutations are deleterious and reduce the fitness of the individuals in which they occur. Thus, spontaneous mutagenesis is in general not adaptive, whereas the direction of evolution depends on natural selection exerted on populations of genetic variants (Arber, 1999). Natural selection reduces genetic variation in a population, eliminating the not-fit-enough. In asexual species, the mutation rate is the main parameter affecting evolvability. Mutation rates can evolve in asexual conditions because an allele that increases the mutation rate (mutator allele) is not eliminated from the population by recombination even if it generates many deleterious mutations (Hurst, 2009). Bacterial mutator alleles provide good examples of systems in which the ability to evolve is selected because of its immediate beneficial consequences (Sniegowski et al., 1997).

Genotypic diversity enhances the evolutionary responsiveness and adaptability of populations (Ayala, 1968a; Abrams and Matsuda, 1997; Yoshida et al., 2003; Gamfeldt et al., 2005; Reusch et al., 2005; Gamfeldt and Kallstrom, 2007; Becks and Agrawal, 2012; Roze, 2012). Genotypic diversity of the seagrass Zostera marina enhances ecosystem stability and therefore buffers against biotic or abiotic disturbances (Hughes and Stachowicz, 2004; Reusch et al., 2005; Reusch and Hughes, 2006; Ehlers et al., 2008). Transects with high genotypic diversity recovered faster from grazing (Hughes and Stachowicz, 2004) and after a high temperature period compared with monoclonal transects (Reusch et al., 2005). Moreover, genotypic diversity increased resistance against parasites in leafcutting ants (Hughes and Boomsma, 2004). Genotypically diverse populations also bear a competitive advantage towards monoclonal populations in Daphnia (Tagg et al., 2005a; b). Genotypic diversity increases competitiveness in invading and resisting invasions of monoclonal populations (Tagg et al., 2005b). Conversely, erosion of genetic diversity impedes adaptive responses to stressful environments (Bijlsma and Loeschcke, 2012; Dierks et al., 2012). In summary, high genotypic diversity serves as a buffer against interfering biotic and abiotic disturbances.

Molecular biologists have long searched for molecular mechanisms responsible for tuning the rate of genetic-variant generation in fluctuating environments. In spite of several bacterial examples, no regulated variation in the genetic-variant generation has been identified in eukaryotic systems (Capp, 2010). Based notably on the example of industrial and pathogenic yeasts, Capp (2010) proposed a nonregulated molecular evolutionary mechanism for the appearance of the transient increase of the genetic-variant generation in eukaryotic cell populations facing challenging environments. The genetic-variant generation in the population was rapidly tuned as a result of a simple Darwinian process acting on the stochastic nature of gene expression (Capp, 2010).

The SMSC as introduced in this work represents the sought-after regulated genetic-variant generation. A faster response to selection is seen as an indicator for an increased genetic variance in fitness created by sex. Sex increases the rate of evolution (Kimura and Ohta, 1971; Kirkpatrick and Jenkins, 1989; Rice and Chippindale, 2001; Butlin, 2002; Poon and Chao, 2004; Goddard et al., 2005; Cooper, 2007; Colegrave and Collins, 2008; Livnat et al., 2008), although evidence of sex constraining genomic and epigenetic variation and slowing down evolution also exists (Kondrashov and Kondrashov, 2001; Futuyma, 2010; Gorelick and Heng, 2011; Melián et al., 2012). There has been evidence that sex allows for faster adaptation in dynamic environments (Hamilton et al., 1990; Howard and Lively, 1994; Keightley and Otto, 2006; Goddard et al., 2005). Indications for faster adaptation has been shown for recombining regions/chromosomes compared with nonrecombining ones in several Drosophila species (McPhee and Robertson, 1970; Rice and Chippindale, 2001; Bachtrog and Charlesworth, 2002; Betancourt and Presgraves, 2002; Rice, 2002; Bachtrog, 2003; Presgraves, 2005). Adaptation to new evolutionary challenges should proceed faster in sexual populations than in asexual populations. Experiments with yeast and the green alga Chlamydomonas reinhardtii (Birdsell and Wills, 1996; Zeyl and Bell, 1997; Greig et al., 1998; Colegrave, 2002; Colegrave et al., 2002b; Kaltz and Bell, 2002; Goddard et al., 2005; Grimberg and Zeyl, 2005) along with work on other taxa (Rice, 2002; Rice and Chippindale, 2001), have largely supported the hypothesis (Albu et al., 2012). Accordingly, sexually reproducing populations are more likely to develop genetic resistance to pathogens, biocides or thermal stress than asexuals (Jaffe et al., 1997; Greig et al., 1998; Rispe and Pierre, 1998; Ooi and Yahara, 1999), at least in part by creating and maintaining genetic variability (Shufran et al., 1997; Bürger, 1999; Robson et al., 1999). Computer modeling indicates that the advantage of sexual reproduction can be substantial in conditions where the mutation rates are higher (Findlay and Rowe, 1990) as is achieved during responses to the selection pressure of a changing environment (Bürger, 1999; Waxman and Peck, 1999). Pre-selection of gametes has been predicted to have significant effects on the frequencies and types of mutations and alleles in a population (Otto and Hastings, 1998), substantially boosting evolvability. In both natural and artificial systems a picture is emerging of populations engaged not in hill-climbing (when they can be trapped on local hilltops) but rather drifting along connected networks of genotypes of equal fitness, with sporadic jumps between networks. These “neutral networks” are of particular significance if they have the “constant innovation” property for this raises the possibility that (given enough time) almost any possible fitness value can ultimately be attained by the population (Barnett, 1998).

Given that there is genetic variation in evolvability, it can be selected for (Jones et al., 2007: Crombach and Hogeweg, 2008; Pigliucci, 2008; Salverda and de Visser, 2011). Turney (1999) suggested the following sufficient (but not necessary) condition for evolvability: If individuals A and B are equally fit but the fittest child of A is likely to be more fit than the fittest child of B, then A is more evolvable than B. The point of this condition is that evolution does not directly select for evolvability, since (by hypothesis) A and B are equally fit. In Turney’s simple computational model, evolvability can increase indefinitely, even when there is no direct selection for evolvability (Turney, 1999). The model shows that increasing evolvability implies an accelerating evolutionary pace (Turney, 1999). For evolvability to increase, environmental change must occur within certain bounds. If there is too little change, there is no advantage to evolvability. If there is too much change, evolution cannot move fast enough to track the changes (Turney, 1999). RNA virus genotypes with similar fitness may differ in their evolvability (Burch and Chao, 2000; McBride et al., 2008). To understand what determines the long-term fate of different clones, each carrying a different set of beneficial mutations, Woods and co-workers (2011) “replayed” evolution by reviving an archived population of Escherichia coli from a long-term evolution experiment and compared the fitness and ultimate fates of four genetically distinct clones. The expected scenario was that eventual winners (EW) clones were already more fit than eventual losers (EL) clones at generation 500, but competition experiments showed that actually the opposite was the case. Surprisingly, two clones with beneficial mutations that would eventually take over the population after 1,500 generations had significantly lower competitive fitness after 500 generations than two clones with mutations that later went extinct. Replaying the experiment many times starting with the 500-generation EWs and ELs showed that the EWs indeed beat the ELs most of the time. Likewise, E. coli strains with larger fitness defects due to deleterious mutations are more evolvable than wild-type clones in terms of both the beneficial mutations accessible in their immediate mutational neighborhoods and integrated over evolutionary paths that traverse multiple beneficial mutations (Barrick et al., 2010).

Mammals, engaging an increased amount of oxidative stress during, particularly male, gametogenesis have evolved at more rapid rates than other organisms (Simpson, 1944; 1953; Romer, 1966; Van Valen, 1974; 1985; Cherry et al., 1982) and humans have evolved extremely rapidly relative to other mammals (Haldane, 1949a; Wyles et al., 1983; Zhang J et al., 2002; Clark et al., 2003; Pollard et al., 2006a; b; Prabhakar et al., 2006; Bird et al., 2007; Hawks et al., 2007; Berglund et al., 2009).

Intriguingly, bet-hedging and evolvability have close links. Variation is the bet-hedging strategy to cover all bases in an often unpredictable environment (Cohen, 1966; Gillespie, 1973; 1974a; Slatkin, 1974; Tonegawa, 1983; Hairston and Munns, 1984; Seger and Brockmann, 1987; Moxon et al., 1994; Danforth, 1999; Meyers and Bull, 2002; Friedenberg, 2003; Balaban et al., 2004; Kussell and Leibler, 2005; Wolf et al., 2005; Venable, 2007; Acar et al., 2008; Ackermann et al., 2008; Beaumont et al., 2009; Olofsson et al., 2009; Childs et al., 2010; Gremer et al., 2012; Morrongiello et al., 2012; Starrfelt and Kokko, 2012). Importantly, variation on virtually every level of biological organization is both the bethedging strategy in variable environments and the firm basis of evolvability.

In evolutionary computation, the Genetic Algorithm (GA) is based on the “survival of the fittest” principle and simulates natural evolution on computer systems to solve complex problems. Individuals are selected and reproduced according to a fitness performance criterion. The fitter the individual, the higher are its chances to produce offspring. Since the process is biased towards the regions of the solution space which enclose the fittest individuals, the evolving population gradually loses diversity and converges. After a population has converged, it is very difficult to readapt to a new optimum when the environment changes (Cobb and Grefenstette, 1993; Simões and Costa, 2002; Bui et al., 2005). Thus, premature convergence is a problem for the GA as it gradually loses its exploratory ability during the evolutionary process under an oversimplified “survival of the fittest” principle. When used as a supplement to the “survival of the fittest” principle, evolvability is able to modify how selection proceeds in the GA and increase its exploratory capabilities (Wang and Wineberg, 2006). Thus, rather than the “survival of the fittest”, the “survival of the more evolvable” is the organizing principle of a dynamic evolutionary process. Sexual reproduction is the fundamental strategy of evolvability set in action.

13.1.1 The ecological pattern of evolvability

Genetic diversity within populations and species diversity within communities are hypothesized to co-vary in space or time because of locality characteristics that influence the two levels of diversity via parallel processes, or because of direct effects of one level of diversity on the other via several different mechanisms (Vellend and Geber, 2005). The plausibility of genetic diversity–species diversity relationships is supported by a variety of theoretical and empirical studies in ecology and evolution (Vellend and Geber, 2005).

Latitudinal gradients in species richness is one of the most universal and oldest known patterns in ecology and biogeography (Rohde, 1992; Brown and Lomolino, 1998; Hawkins et al., 2003a; Willig et al., 2003; Hillebrand, 2004a; b; Roy et al., 2004). Large data sets point towards climate as the major driving force of higher tropical species richness (Fischer, 1960; Stevens, 1989; Gaston, 1996; 2000; Cardillo, 1999; Hawkins and Porter, 2001; 2003; Francis and Currie, 2003; H-Acevedo and Currie, 2003; Hawkins BA et al., 2003a; b; 2007; Willig et al., 2003; Currie et al., 2004; Hawkins and Diniz-Filho, 2004; Cardillo et al., 2005). Faster tropical diversification has been implied by lower average taxon ages or higher rates of first appearances in the tropics (Stehli et al., 1969; Durazzi and Stehli, 1972; Hecht and Agan, 1972; Jablonski, 1993; Flessa and Jablonski, 1996; Allen A et al., 2006; Jablonski et al., 2006; Krug et al., 2007). At least 100 hypotheses have been proposed to explain this biogeographical pattern (e.g., Pianka, 1966; Rohde, 1978; 1992; Rahbek and Graves, 2001; Willig et al., 2003). There is little consensus as to which hypothesis (or combination of hypotheses) is the most likely explanation. Many hypotheses address how ecological processes might allow larger numbers of species to coexist in the tropics (e.g., productivity, energy, stability, spatial heterogeneity, predation, and competition hypotheses) (Pianka 1966; Willig et al., 2003). Several other hypotheses focus (explicitly or implicitly) on potential differences in rates of speciation and extinction between temperate and tropical regions (e.g., evolutionary rates hypothesis) (Willig et al., 2003; Mittelbach et al., 2007). A variety of correlations between, on the one hand, body size, metabolic rate, energy flux, ambient temperature and, on the other hand, variations in rates of nucleotide substitution, and tempo of macro- and micromolecular evolution have been recognized (Martin and Palumbi, 1993; Hawkins et al., 2003a; Pörtner, 2006; Kreft and Jetz, 2007; Welch et al., 2008; Sibly et al., 2012). Hypermutability is one of the most striking features of animal mitochondria: the mtDNA mutation rate is typically one order of magnitude higher than the nuclear one (Brown et al., 1979; Ballard and Whitlock, 2004; Lynch, 2006). This high mutation rate is one of the reasons why mtDNA is a very popular marker for biodiversity studies (Nabholz et al., 2009). There is a strong signal for the effects of metabolic rate on mitochondrial DNA evolution (Avise et al., 1992b; Adachi et al., 1993; Martin and Palumbi, 1993; Rand, 1994; Nunn and Stanley, 1998; Martin, 1999; Welch et al., 2008). Cumulative evidence indicates that both mitochondrial ROS generation (Zar and Lancaster, 2000; Abele et al., 2002; Heise et al., 2003; 2006; Mujahid et al., 2005; 2007) and DNA damage by ROS (Eigner et al., 1961; Greer and Zamenhof, 1962; Lindahl and Nyberg, 1972; 1974; Frederico et al., 1990; Lindahl, 1993; Bruskov et al., 2002) are temperature-dependent. DNA repair capacity and net mutation rate are also temperature-dependent in a wide range of taxa (Muller, 1928; Lindgren, 1972 a; b; Savva, 1982; Raaphorst et al., 1999; Schmidt-Rose et al., 1999; El-Awadi et al., 2001; Lupu et al., 2004; 2006; Xu, 2004b). Thus, ambient temperature and energy flux drive molecular evolution in both plants, exo- and endothermic organisms (Currie, 1991; Bleiweiss, 1998; Abele et al., 2002; Allen A et al., 2002; 2006; Wright et al., 2003; 2006; 2010; 2011; Brown et al., 2004; Davies et al., 2004; Clarke and Gaston, 2006; Pörtner, 2006; Sanders et al., 2007; Hiddink and ter Hofstede, 2008; Jansson and Davies, 2008; Gillman et al., 2009; 2010; McCain, 2009; Stegen et al., 2009; Tittensor et al., 2010). On mountains, temperature decreases monotonically by an average of 0.6°C per 100-m elevational gain (Barry, 1992). Supporting the role of temperature, the diversity pattern is consistent both for temperature gradients due to latitude and elevation (Bleiweiss, 1998; Brown, 2001; Krömer et al., 2005; Kluge et al., 2006; Sanders et al., 2007; Gillman et al., 2009; McCain, 2009). The pattern of increased molecular evolution in species living in habitats with greater biologically available energy may also occur independently of latitudinal clines (Wright et al., 2003). These correlations led to the formulation of a metabolic eco-evolutionary model of biodiversity (Allen A et al., 2002; 2006; 2007; Brown et al., 2004; Stegen et al., 2009; 2012). Endothermic ‘warmblooded’ animals that use metabolism to maintain a constant body temperature (birds and mammals) have higher absolute rates of molecular evolution of mitochondrial and nuclear DNA than poikilothermic vertebrates with environmentally determined body temperatures, such as reptiles and fish (Martin et al., 1992; Adachi et al., 1993; Martin and Palumbi, 1993; Rand, 1994; Rico et al., 1996). The ecological pattern of evolutionary rate supports the notion of a universal dependence of molecular evolution on oxidative stress.

13.1.2 The tempo of evolution

For no bias can be more constricting than invisibility—and stasis, inevitably read as absence of evolution, had always been treated as a non-subject.

Gould and Eldredge, 1993

Mathematical models that mimic biological evolutionary processes have revealed that the traditional view of Darwinian evolution, according to which the most fit of random mutants are selected, faces a major problem (Eden, 1967; Schützenberger, 1967; Bak et al., 1987; 1988; Bak, 1993; 1996; Fernández et al., 1998): It is much too slow to account for real evolution. Bak (1993, cited in Fernández et al., 1998) described the difficulty: “If, for the sake of argument, we imagine the outer world frozen (for a while) and try to construct from scratch an equally fit species by recourse to engineering techniques rather than by evolution, we will be forced to accept that eons are needed. By starting at a random configuration one certainly will reach a wrong and much less fit maximum. It would be necessary to systematically go through all configurations, involving exponentially large times.” According to Kashtan et al. (2007), computer simulations that mimic natural evolution by incorporating replication, variation (e.g., mutation and recombination) and selection, typically observe a logarithmic slowdown in evolution: longer and longer periods are required for successive improvements in fitness (Lipson et al., 2002; Lenski et al., 2003; Kashtan and Alon, 2005). Simulations can take many thousands of generations to reach even relatively simple goals, such as Boolean functions of several variables (Lenski et al., 2003; Kashtan and Alon, 2005).

Both, coevolutive systems and environmental variation may speed up the tempo of evolution. Both theoretical modeling and ecological evidence indicate that the rate of evolution is accelerated by coevolutionary patterns (Dieckmann and Law, 1996; Marrow et al., 1996; Fernández et al., 1998; Sorenson and Payne, 2001; Rice and Chippindale, 2002; Good-Avila et al., 2006). Likewise, temporally varying environments, particularly those that change over time in a modular fashion, such that each new environment shares some of the subproblems with the previous environment, can speed up evolution (Otto and Michalakis, 1998; Earl and Deem, 2004; Kashtan and Alon, 2005; Kashtan et al., 2007). By nature of the mathematical models, they could not identify the molecular processes that would be able to effect the evolutionary speedup.

Estimates of the rates of evolution from contemporary studies range from four to seven orders of magnitude higher than those seen in the fossil record (Gingerich, 1983; 2009; Reznick et al., 1997; Reznick and Ghalambor, 2001). Similarly, more than half a century ago, Kurtén (1959) found an inverse correlation between the rate of morphological evolution and the time interval over which the rate was measured. Specifically, the rate of morphological change between successive generations exceeded macroevolutionary rates by several orders of magnitude. This time-dependent rate pattern was confirmed in a number of subsequent studies, which consistently showed that morphological evolutionary rates appeared faster when measured over shorter timescales (Gingerich, 1983; 2001; Roopnarine 2003). In the words of Gingerich (2009), “At… a rate found commonly in rate studies, a mammal could conceivably change from the size of a shrew to the size of a whale in 103 generations, and with suitable selection do this back and forth as many as 104 = 10,000 times in the geological span of known insectivores and whales.” Indeed, contemporary evolution can be very rapid (Endler, 1986; Thompson, 1998; Hendry and Kinnison, 1999; Yoshida et al., 2003; Reznick et al., 2004b; Hairston et al., 2005; Carroll et al., 2007; Latta et al., 2007; Schoener, 2011). Invasive species may be evolving with extraordinary rapidity in their new environments (Losos et al., 1997; Reznick et al., 1997; Diniz-Filho et al., 1999; Huey et al., 2000; 2005; Mooney and Cleland, 2001; Palumbi, 2001; Drummond et al., 2003; Blair and Wolfe, 2004; Schoener, 2011). A puzzling phenomenon has recently emerged in rates of molecular evolution estimated from DNA sequence data. In particular, there is a striking disparity between spontaneous mutation rates, measured over a small number of generations in studies of pedigrees and laboratory mutation-accumulation lines, and the much lower substitution rates measured over geological time frames (e.g., Parsons et al., 1997; Sigurdardóttir et al., 2000; Howell et al., 2003; Ho et al., 2005; 2008; 2011; Santos et al., 2005; Gibbs et al., 2009). Similarly, for the same virus a higher rate of evolution is obtained when the compared sequences correspond to viruses isolated within a short time span than when they correspond to isolates separated by a long time interval (Sobrino et al., 1986; Domingo, 2007). In a long-term evolution experiment of D. melanogaster over 600 generations, Burke et al. (2010) found little allele frequency differentiation between replicate populations. Adaptation was not associated with ‘classic’ sweeps whereby newly arising, unconditionally advantageous mutations become fixed (Burke et al., 2010). Thus, the substitution rate appears to be much lower than the mutation rate. This pattern is what can be expected from a model presented by Neher et al. (2010). In large populations, when the product of the population size N and the total beneficial mutation rate Ub is large, many new beneficial alleles can be segregating in the population simultaneously.The rate of adaptation in several models of such sexual populations increases linear with NUb only in sufficiently small populations. In large populations, the rate of adaptation increases much more slowly as logNUb. This change from linear to logarithmic dependence on NUb indicates that the rate of adaptation is limited by interference among multiple simultaneously segregating beneficial mutations rather than by the supply of beneficial mutations. As in the asexual case, because of interference between mutations, only a small fraction of the beneficial mutations fix—the rest are wasted. However, this fraction increases with increasing rate of recombination until it saturates at NUbs2, which is the limit of independently fixating mutations (Neher et al., 2010). Chevin and Hospital (2008) modeled the trajectory of an initially rare beneficial allele that does not reach fixation because its selective advantage is inversely proportional to the distance to a new phenotypic optimum, and that optimum is reached, because of other loci, before the variant fixes. This pattern resembles the situation in RNA viruses where evolution rates increase linearly with mutation rates for slowly mutating viruses. However, this relationship plateaus for fast mutating viruses (Sanjuán, 2012).

It is a long-standing issue whether this rapid evolution is based on standing genetic variation or is driven by de novo mutations. Understanding the source of variation for adaptation might tell us a great deal about the factors creating and maintaining genetic variation in natural populations (Hedrick, 1986; 2006; Lynch and Walsh, 1998; Barton and Keightley, 2002). The theory for standing variation assumes that newly beneficial alleles are neutral or deleterious prior to the change of environment and are maintained in the ancestral population through a balance of recurrent mutation, selection and drift (Orr and Betancourt, 2001; Hermisson and Pennings, 2005; Przeworski et al., 2005). Compared with new mutations, adaptation from standing genetic variation is thought to likely lead to faster evolution, the fixation of more alleles of small effect and the spread of more recessive alleles. There is potential to distinguish between adaptation from standing variation and that from new mutations by differences in the genomic signature of selection. Selection on standing genetic variation is ubiquitous in natural populations (Hill, 1982a; Hermisson and Pennings, 2005; Przeworski et al., 2005; Roff, 2007; Barrett and Schluter, 2008; Teotónio et al., 2010) and implicated in the rapid adaptive response to stressful environments (Jarosz and Lindquist, 2010). In artificial selection programs the time scale is usually considered too short for mutations to influence the rates or limits substantially, but this view has been questioned (Frankham, 1980). Hundreds of artificial selection experiments have generated changes in mean phenotypes well beyond the observed range in the base population in just a few dozen generations (Falconer and Mackay, 1996), inspiring the view that quantitative variation is distributed over an effectively infinite number of loci with minuscule effects (Kimura, 1965; Lande, 1975; Bulmer, 1980). Much of the earliest work in theoretical population genetics downplayed the ability of mutation to overcome the force of selection (Fisher, 1930; Haldane, 1932). There have been continued responses over periods of 50 or more generations in some artificial selection experiments (Dudley, 1977; Enfield, 1980; Yoo, 1980a; b; c; Barton and Turelli, 1989; Falconer and Mackay, 1996; Weber, 1996; Weber et al., 1999; Barton and Keightley, 2002; Laurie et al., 2004). The record attained by artificial selection suggests that the response to selection is hardly limited to fine tuning (Reznick and Ghalambor, 2001), as has often been claimed (e.g., Gould and Eldredge, 1993). These sustained responses do not exhaust genetic variation (Frankham, 1980; Yoo, 1980b; Hill, 1982a; b; Dunnington and Siegel, 1996; Eitan and Soller, 2004; Laurie et al., 2004; Carlborg et al., 2006; Burke et al., 2010), arguing for a steady supply of new mutations. The increase in additive genetic variance due to mutation is ~10-3 of the standing variation, which makes a significant contribution after ~50 generations of selection (Hill, 1982a; Lynch and Walsh, 1998). Genetic variation and, more specifically, lack of mutations should not and, it seems, does not limit at least straightforward selection response (Barton and Partridge, 2000).

Many aspects of reproduction are facilitated by proteins that control crucial cellular events such as mating type determination, spermatogenesis, gamete recognition, sperm-egg binding, etc. Contrary to their essential role, reproductive proteins are among the fastest evolving genes known to us (Civetta and Singh, 1995; Lee et al., 1995; Swanson and Vacquier, 1995; 1998; 2002; Metz and Palumbi, 1996; Ferris et al., 1997; Tsaur and Wu, 1997; Metz et al., 1998; Tsaur et al., 1998; Vacquier, 1998; Aguade, 1999; Hellberg and Vacquier, 1999; Gavrilets, 2000; Hellberg et al., 2000; Singh and Kulathinal, 2000; Wyckoff et al., 2000; Armbrust and Galindo, 2001; Swanson et al., 2001). For example, lysin, a sperm protein of marine invertebrates, evolves up to 50 times faster than the most rapidly evolving mammalian gene (Metz et al., 1998). Genome wide comparisons of some 1200 human-mouse orthologous genes have shown that many genes directly related to gamete adhesion rank in the top 5% of the most divergent genes (Makalowski et al., 1996). Likewise, accessory gland proteins, which are part of Drosophila seminal fluids, are twice as diverse between species as non-reproductive Drosophila proteins (Civetta and Singh, 1995).

If evolution can be so fast, then why does it appear to be so slow in the genetic and geological fossil record? Stasis is a prevalent–perhaps the most prevalent–mode of evolution, especially pronounced on long timescales (Merilä et al., 2001; Gould 2002; Eldredge et al., 2005; Hairston et al., 2005; Estes and Arnold, 2007). Eldredge et al. (2005) reviewed how taxa can commonly exhibit both short-term evolutionary dynamics and long-term stasis and concluded that the complex pattern of selection imposed on geographically structured populations by heterogeneous environments and coevolution can paradoxically maintain stasis at the species level over long periods of time. Bell (2010) argued that selection fluctuates strongly over time with large changes in magnitude and frequent reversals in the direction of selection. This could retard the loss of genetic variation, with minimal long-term directional change. He supported this contention by estimates of selection coefficients, or other parameters directly related to selection coefficients, reported in contemporary population time-series (Bell, 2010). Carroll et al. (2007) explained the discrepancy between short-term fluctuating selection and long-term stasis: “Specifically, selection and evolution often fluctuate dramatically in direction through time, presumably tracking fluctuating environments, so that rapid short term changes rarely accumulate into long-term directional trends” (Gibbs and Grant, 1987; Hairston and Dillon, 1990; Ellner et al., 1999; Merilä et al., 2001; Grant and Grant, 2002; 2006; Calsbeek et al., 2012). Traits that are no longer under selective pressure degenerate or are fossilized to pseudogenes (Heininger, 2012). Inferential evidence that “stasis” is a highly dynamic process comes from the maintenance of a variety of processes that mediate adaptive plasticity (Masel et al., 2007). A coalescence of time scales on which ecological and evolutionary processes are acting has been proposed (Hairston et al., 2005; Carroll et al., 2007). For instance, in a prey-predator coevolutionary system (Abrams and Matsuda, 1997), the impact of prey evolution on predator per capita growth rate is 63% that of internal ecological dynamics (Hairston et al., 2005). For Darwin’s finches evolving in response to fluctuating rainfall (Grant and Grant, 2002) it has been estimated that evolutionary change has been more rapid than ecological change by a factor of 2.2 (Hairston et al., 2005). For a population of freshwater copepods whose life history evolves in response to fluctuating fish predation (Hairston and Dillon, 1990) evolutionary change has been estimated about one quarter the rate of ecological change – less than in the finch example, but nevertheless substantial (Hairston et al., 2005).

On the other hand, it appears fairly generally accepted that stabilizing selection is an eminent cause of stasis (Vrba, 1980; Boucot, 1990; Gould, 2002, p. 880-5; Hansen and Houle, 2004; Estes and Arnold, 2007). Evolution is variously constrained on all levels of biological organization, from genome sequence to genome architecture, gene expression, molecular interactions and organismal phenotypes (Kimura, 1983; Clark, 1987; Loeschcke, 1987; Lynch, 2007a; Wolf et al., 2008; Koonin and Wolf, 2010). Viability and potential fitness gains, in a given genetic constraint, build a narrow evolutionary path (Tokuriki and Tawfik, 2009). Experiments showed that knockout of any essential gene confers a lethal phenotype to an organism (Fraser et al., 2000; Herring and Blattner, 2004). The number of such essential genes varies from organism to organism: all genes are essential in viruses, whereas in bacteria, essential genes can reach up to one-third of all genes (Zeldovich et al., 2007). Antagonistic pleiotropy (leading to negative genetic correlations), epistasis, and linkage disequilibrium can all constrain the generation of novel genotypes (Barton and Partridge, 2000). Pleiotropy is expected to constrain the rate of evolution (Otto, 2004), consistent with the observation that broadly expressed genes evolve more slowly (Hughes and Hughes, 1995; Hastings, 1996; Hurst and Smith, 1999; Duret and Mouchiroud, 2000; Hirsh and Fraser, 2001; Jordan et al., 2002; Subramanian and Kumar, 2004; Zhang and Li, 2004; Wall et al., 2005; Zhang and He, 2005; Liao and Zhang, 2006; Liao et al., 2006; Larracuente et al., 2008). Long-term protein evolution is constrained by epistasis: substitutions that are accepted in one genotype are deleterious in another (Weinreich et al., 2005; Camps et al., 2007; Breen et al., 2012). Estimating the prevalence of epistasis in long-term protein evolution by relating data on amino-acid usage in 14 organelle proteins and 2 nuclear-encoded proteins to their rates of short-term evolution in at least 1,000 orthologues for each of these 16 proteins from species from a diverse phylogenetic background, Breen et al. (2012) found that the measured rate of amino-acid substitution in recent evolution is 20 times lower than the rate of neutral evolution and an order of magnitude lower than that expected in the absence of epistasis, indicating that epistasis is pervasive throughout protein evolution. The process of amino acid replacement in proteins is context-dependent, with substitution rates influenced by expression level, local structure, functional role, pleiotropic effects, and amino acids at other locations (Dayhoff et al., 1978; Jones DT et al., 1992; Overington et al., 1992; Koshi and Goldstein, 1995; 1998; Thorne et al., 1996; Jensen and Pedersen, 2000; Whelan and Goldman, 2001; Lartillot and Philippe, 2004; Pagel and Meade, 2004; Siepel and Haussler, 2004; Weinreich et al., 2005; 2006; Lovell and Robertson, 2010; Soskine and Tawfik, 2010; Salverda et al., 2011; Pollock et al., 2012). Often a limited number of trajectories might in fact be possible (Wood et al., 2005; Weinreich et al., 2006). Moreover, evolution in reverse is a widespread phenomenon in biology (Teotónío and Rose, 2000; 2001; 2002; Wiens, 2001; Porter and Crandall, 2003; Estes and Teotónio, 2009; Bell, 2010).

A mathematical model suggested a universal speed limit on the rate of molecular evolution by predicting that populations go extinct (via lethal mutagenesis) when mutation rate exceeds approximately six mutations per essential part of genome per replication for mesophilic organisms and one to two mutations per genome per replication for thermophilic ones (Zeldovich et al., 2007). There may be also upper limits imposed by functional or structural constraints; for example, excessive heterozygosity could interrupt chromosome pairing (Stephan and Langley, 1992) or lead to reproductive incompatibilities between individuals living in distant regions of the species’ range (e.g. Seidel et al., 2008). For instance, MMR proteins actively inhibit recombination between diverged sequences (Chen and Jinks-Robertson, 1998; Kolodner and Marsischky, 1999), thus controlling rates of mutation and evolutionary adaptation.

The fossil record and molecular phylogenies indicate that evolutionary dynamics proceed by punctuational episodes rather than gradual change (Eldredge and Gould, 1972; Gould and Eldredge, 1977; 1993; Sneppen et al., 1995; Messier and Stewart, 1997; Cubo, 2003; Eldredge et al., 2005; Pagel et al., 2006; Hunt, 2007; Mattila and Bokma, 2008; Zeh et al., 2009; Venditti and Pagel, 2010; Laurin et al., 2012; Rabosky, 2012). Thereafter, pronounced decelerations in rates of phenotypic and genotypic evolution have been observed when populations are approaching a local adaptive peak (Lenski and Travisano 1994; Sniegowski et al., 1997; Cooper and Lenski, 2000; Elena and Lenski, 2003; Pagel et al., 2006). The exhaustion of beneficial variants—whether preexisting or potentially accessible by mutation, or the diminishing selective benefit of beneficial mutations may then slow the potential and speed of adaptation (de Visser et al., 1999; Chou et al., 2011) and increase the costs of mutagenesis. Starting from weak functional constraints near the time of origin of a gene there would be a gradual increase in selective pressures with time, resulting in fewer accepted mutations in older versus more novel genes (Albà and Castresana, 2005). In well adapted populations living in stable habitats, conservation may become more important than innovation. Evolutionary models based on the asexual and sexual replication pathways in Saccharomyces cerevisiae suggested that sexual replication can eliminate genetic variation in a static environment, as well as lead to faster adaptation in a dynamic environment (Gorodetsky and Tannenbaum, 2008). But species-wide depletion of accessible beneficial mutations requires a degree of environmental constancy that is not typical of the earth’s history (Lambeck and Chappell, 2001; Zachos et al., 2001; Eldredge et al., 2005).

Pronounced decelerations in rates of phenotypic evolution have been observed over thousands of generations in asexual populations of E. coli founded from a single cell (Cooper and Lenski, 2000). Stasis cannot derive from depletion of preexisting variation, nor from exhaustion of genetic variation more generally. In fact, the amount of genetic variation increased in these populations even as the rate of phenotypic evolution declined (Sniegowski et al., 1997). These populations evidently approached a local adaptive peak or plateau, at which point most potential (i.e., genetically accessible) beneficial mutations were fixed (Eldredge et al., 2005). Consistent with this explanation, the rate of adaptive evolution was re-accelerated by perturbing populations from their proximity to an adaptive peak, either by changing the environment (Travisano et al., 1995) or by introducing deleterious mutations (Moore FBG et al., 2000). In the framework of the fitness landscape described by Wright (1931), populations placed near the top of a fitness peak will experience less beneficial mutations than populations placed far from the adaptive optimum (Fisher, 1930; Lenski and Travisano, 1994; Elena and Lenski, 2003; Lázaro et al., 2003; Silander et al., 2007; Stich et al., 2010). Adaptation is therefore characterized by a pattern of diminishing returns — larger-effect mutations are typically substituted early on and smaller-effect ones later (Barton, 1998; Orr, 1998; 1999; 2005a; de Visser et al., 1999; Barton and Keightley, 2002; Barrick et al., 2009; Chou et al., 2011; Khan et al., 2011; Flynn et al., 2013). In genomes sampled through 40,000 generations from a laboratory population of Escherichia coli, although adaptation decelerated sharply, genomic evolution was nearly constant for 20,000 generations. Several lines of evidence indicate that almost all of these mutations were beneficial. This same population later evolved an elevated mutation rate and accumulated hundreds of additional mutations dominated by a neutral signature (Barrick et al., 2009). Complex species (those having many characters) typically show slower increases in fitness during adaptation than do simple species (Orr, 2000b). Part of the reason is that the distance travelled to the optimum by a beneficial mutation is smaller in a complex than a simple species (this distance decreases with the square root of the number of characters) (Orr, 2000b). Recent work by Welch and Waxman (2003) indicates that this cost of complexity might be a general feature of adaptation (Orr, 2005a).

13.2 Robustness

Robustness can be defined as the tendency for a system to maintain functionality under perturbation. Mutational (or genetic) robustness is defined as the constancy of a phenotype in the face of deleterious mutations (Sanjuán et al., 2007). Waddington recognized decades ago that levels of phenotypic variation in natural populations tend to be small compared to what might be expected given typical levels of genetic and environmental variation (Waddington, 1957). Robustness seems to be the opposite of evolvability. If phenotypes are robust against mutation, it might be expected that a population will have difficulty adapting to an environmental change, as several studies have suggested (Ancel and Fontana, 2000; Carter et al., 2005; Sumedha et al. 2007; Cowperthwaite et al., 2008; Parter et al., 2008; Draghi et al., 2010). However, other studies contend that robust organisms are more adaptable (Bloom et al., 2006; Aldana et al., 2007; Elena and Sanjuán, 2008; McBride et al., 2008; Draghi et al., 2010; Wagner, 2012). Thus, robustness generally confers evolvability to living systems because it allows them to undergo innovative modification without losing functionality (Wagner, 2005a; Wagner, 2008a). High mutation rate is perhaps the most important prerequisite for adaptive genetic robustness (de Visser et al., 2003), so mutational robustness should be strongly selected in biological systems experiencing elevated mutation rates (Wagner et al., 1997; Montville et al., 2005). Evolution in silico demonstrated that faster replicating organisms can easily be out-competed at high mutation rates by slower replicators with greater mutational robustness (Wilke et al., 2001). The impact of deleterious mutations can be reduced when several genes contribute toward a single function, or when there are several copies of a single gene (Tautz, 1992; Wilkins, 1997). In both cases the consequence is genetic redundancy, also called genetic canalization (Gibson and Wagner, 2000). Phenotypic buffering of genomic mutations as provided by gene duplication and diploidy might have evolved to facilitate genetic canalization in higher organisms (Krakauer and Plotkin, 2002; Gu et al., 2003). Biological systems are extraordinarily robust to perturbation by mutations, recombination and the environment. Intriguingly, populations evolved with higher mutation rates show a higher robustness under mutations (Mihaljev and Drossel, 2009). It has been proposed that this robustness might make them more evolvable. In fact, robustness is widely held to be a key factor advancing evolvability (Kirschner and Gerhart, 1998; Rutherford and Lindquist, 1998; Aharoni et al., 2005; Wagner, 2005a; b; 2008a; 2012; Masel and Siegal, 2009; Draghi et al., 2010; Masel and Trotter, 2010). Robustness to mutation allows genetic variation to accumulate in a cryptic state. Cryptic genetic variation is variation that is not normally seen by natural selection. A considerable amount of cryptic genetic variation may accumulate in the genomes of organisms, particularly when they are subjected to stabilizing selection (Hermisson and Wagner, 2004; Félix and Wagner, 2008; Le Rouzic and Carlborg, 2008; Masel and Siegal, 2009). Cryptic genetic variation is uncovered by environmental or genetic perturbations and might be an essential source of physiological and evolutionary potential (Gibson and Dworkin, 2004; Schlichting, 2008). Switching mechanisms known as evolutionary capacitors mean that the amount of heritable phenotypic variation available can be correlated to the degree of stress and hence to the novelty of the environment and remaining potential for adaptation. The best studied, but not only (Takahashi, 2012), putative evolutionary capacitor in a high recombination system is the heat shock protein Hsp90 (Rutherford and Lindquist, 1998; Queitsch et al., 2002; Rutherford, 2003; Debat et al, 2006; Kellermann et al., 2007; Rutherford et al., 20007a; b).

To my knowledge none of the theories attempting to explain the rationale for sexual reproduction acknowledged the implications that phenotypic robustness may have for the evolutionary benefits of sexual reproduction. Phenotypic robustness can be expected to attenuate and delay the immediate benefit of sexual reproduction-related genetic variation and hence, may be regarded as additional liability for the evolution of sexual reproduction.

14. Stress and sex: a double-edged relationship


Summary

The sex-stress/glucocorticoid relationship is non-linear and is described by approximation as inverted “U”-shaped: sex is favored in intermediate stressful environments, while stable stressfree and extreme stressful environments favor asex. This modulation is mediated by a variety of endocrine, paracrine, and autocrine signals and intense cross talk between the hypothalamic–pituitary–adrenal (HPA) and hypothalamic–pituitary–gonadal (HPG) axes. Male reproductive performance is more susceptible to environmental stress, presumably due to the additive effects with endogenous gametogenetic stress. Additionally, in vertebrate males the HPG and HPA axes are linked in a double-negative feedback loop that toggles between sexual reproduction with functional males and asexual reproduction in response to trigger stimuli that impinge upon the feedback circuit. In contrast, the female vertebrate HPG axis is linked by a positive feedback to the HPA axis, while the HPA axis attenuates the HPG axis in a negative feedback loop, a pattern that can trigger oscillations which could explain the female menstrual cycles in vertebrates. There is evidence that the human spermatozoon is a cell in crisis, operating near the threshold of error catastrophe. Exemplarily, three hormonal systems, glucocorticoids, thyroid hormones and melatonin, are discussed that impinge upon the HPA and HPG axes and regulate reproductive activity and gonadal oxidative stress. Overall, HPA and HPG axes cross-talk on a multitude of levels to adapt sexual reproductive activity and (epi)mutagenesis as potential bet-hedging response to environmental conditions.

Sexual reproduction is intimately linked to ecological conditions (Lloyd, 1980). Sexual reproduction and stress are intimately related. Moderate levels of stress and stress mediators may have a positive effect on reproductive processes while greater stress has negative effects (Greenberg and Wingfield, 1987; Sapolsky et al., 2000; Moore and Jessop, 2003; Breuner et al., 2008; Milla et al., 2009; Schreck, 2010). Evolutionary theory predicts that stable environments to which organisms are well adapted would favor low mutation rates (anti-mutator genotypes), constrained only by the costs of error-repair mechanisms (Kimura, 1967; Drake, 1991). In contrast, environments subjected to frequent changes would select for increased mutation rates (mutator genotypes) that permit faster adaptation to the new conditions (de Visser, 2002; Travis and Travis, 2002; Denamur and Matic, 2006). Stressful environmental conditions elicit elevated (epi-)mutation and recombination rates and increased genetic and epigenetic diversity and phenotypic variation. This occurs by at least two mechanisms: (i) the uncovering of pre-existing hidden variation or (ii) the generation of de novo variation, for example by increased (epi-)mutation, transposon activity, or sex/recombination (Plough, 1917; 1921; Neel, 1941; Grell, 1971; 1978; Nevo et al., 1979; 1997b; 1998; Balyaev and Borodin, 1982; Zhuchenko and Korol, 1983; Parsons, 1987; Hoffmann and Parsons, 1991; Hall, 1992; Steele and Jinks-Robertson, 1992; Korol et al., 1994; Abdullah and Borts, 2001; Finnegan, 2002; Lucht et al., 2002; Hadany and Beker, 2003a; b; Kovalchuck et al., 2003; Agrawal et al., 2005; Morgan, 2005; Molinier et al., 2006; Rando and Verstrepen, 2007; Baer, 2008; Boyko and Kovalchuk, 2011; Forche et al., 2011; Zhong and Priest, 2011; Grativol et al., 2012; Steinberg, 2012). Increased recombination in response to stress (fitness-associated recombination) (Plough, 1917; Grell, 1971; Zhuchenko et al., 1986; Gessler and Xu, 2000; Hadany and Beker, 2003a; Schoustra et al., 2010; Zhong and Priest, 2011) is thought to accelerate the rate of adaptation (Hadany and Beker, 2003b). In Arabidopsis thaliana plants treated with short wavelength radiation or flagellin (an elicitor of plant defences), somatic homologous recombination is increased in the treated population and these increased levels of homologous recombination persist in the subsequent, untreated generations (Molinier et al., 2006). In plants, animals, and fungi, TE activity increases in response to extrinsic stress and oxidative stress (McClintock, 1984; Arnault and Dufournel, 1994; Mhiri et al., 1997; Grandbastien, 1998; Capy et al., 2000; Ikeda et al., 2001; Chen et al., 2003a; Daboussi and Capy, 2003; Lu and Ramos, 2003; Jorgensen, 2004; McGraw and Brookfield, 2006; Bouvet et al., 2008; Cam et al., 2008; Perez-Hormaeche et al., 2008; Zeh et al., 2009; Rebollo et al., 2010; 2012; Stoycheva et al., 2010; Casacuberta and González, 2013).

A relationship between stressful conditions and elevated variance of physiological and phenotypic parameters has been documented in ecological studies (Møller and Swaddle, 1995; Orlando and Guillette, 2001; Joyner Matos, 2007; Fraterrigo and Rusak, 2008) that may even be useful as a biomarker (Callaghan and Holloway, 1999). Following environmental stress, both genetic and phenotypic variance of offspring increases (Balyaev and Borodin, 1982; Parsons, 1983; 1988; Nevo, 1988; 1998; 2001; 2011; Hoffmann and Parsons, 1991; 1997; Goho and Bell, 2000; Bubliy and Loeschcke, 2002; Kis-Papo et al., 2003; Imasheva and Loeschcke, 2004; Badyaev, 2005a; b; Swindell and Bouzat, 2006; Theodorakis et al., 2006; Crean and Marshall, 2009) and heritability decreases, at least for morphological traits (Charmantier and Garant, 2005; Wilson AJ et al., 2006). Environmental fluctuations are usually believed to play a "destructive role" in ecosystem dynamics and to act as a source of disturbance. However, noise is also known for its "constructive role", i.e., for the ability to create new ordered states in dynamical systems. Environmental noise may also enhance biodiversity (D'Odorico et al., 2008). An argument that has been repeatedly used to question the adaptive role of condition-dependent mutagenesis is that individuals in poor physiological condition have higher mutation rates for reasons having nothing to do with the possibility of generating a lucky beneficial mutation. Assuring the fidelity of DNA replication is metabolically costly and involves the products of many dozens to hundreds of genes (Wood et al., 2001). Individuals in poor condition will have fewer resources to devote to genomic surveillance, leading to the possibility that individuals in poor condition will suffer an increased mutation rate (Baer, 2008). Organisms apply two strategies to meet environmental challenges: resist or mutate (Heininger, 2001). To resist in periods of environmental harshness, highly efficient stress resilience mechanisms and states of metabolic dormancy (spores, diapause, hibernation) have evolved (Heininger, 2012). On the other hand, a generic thermodynamical analysis of genetic information storage yielded the insight that mutation rate depends on availability/utilization of metabolic resources. A lowered ability to employ metabolic resources in mutation suppression increase the minimum effective mutation rate. This predicts transient mutation rate increases as a response to stress (Hilbert, 2011).

Theoretical and experimental studies across a broad range of taxa demonstrate that sexual reproduction is favored in, and a response to, moderately stressful environments (Bell and Wolfe, 1985; Iglesias and Bell, 1989; Zeyl and Bell, 1997; Greig et al., 1998; Grishkan et al., 2003; Nedelcu and Michod, 2003; Eads et al., 2008; Martins et al., 2008; Denekamp et al., 2009; Steinberg, 2012). Likewise, oxidative stress as final common effector of stress signaling pathways is a general inducer of sexual reproductive activity (Bernstein and Johns, 1989; Heininger, 2001; Nedelcu and Michod, 2003; Nedelcu et al., 2004; Nedelcu, 2005; McInnis et al., 2006). In fungi, deletion or mutagenesis of NADPH oxidases that are used deliberately to produce ROS, specifically block differentiation of sexual fruit bodies, without affecting asexual development (Lara-Ortíz et al., 2003; Malagnac et al., 2004; Takemoto et al., 2007).

Gene swapping evolved as response to environmental variability. Facultatively sexual organisms often engage in sex more often when in poor condition. Aptly, the term condition-dependent sex was coined (Hadany and Otto, 2007; 2009). Like condition-dependent mutagenesis (Baer, 2008) sexual reproduction is fitness-dependent in these organisms. Intriguingly, as a legacy to these shared evolutionary roots, similar neurotransmitters and highly connected nuclei within the hypothalamus of mammals control stress and reproduction (Dobson et al., 2003). Sexual reproduction “institutionalized” the generation of genetic variation as proactive coping strategy to biotic and abiotic variability and unpredictability. Ingeniously, applying a harsh stress-related gamete quality assurance, sexual reproduction introduced increasingly sophisticated selection regimes that selected gametes for their stress resistance, viability and metabolic efficiency.

The coin, however, has its flip side. Requiring a substantial amount of physiological and oxidative stress for its basic functions (Riley and Behrman, 1991; Heininger, 2001; Agarwal et al., 2003; 2005; Lue et al., 2003; Nedelcu, 2005; Aitken and Roman, 2008; Metcalfe and Alonso-Alvarez, 2010), the reproductive system, and particularly gametogenesis, is highly sensitive to additional abiotic and biotic stress. The double-edged relationship between sexual reproduction and stress is epitomized by evidence that pro-inflammatory cytokines both can stimulate (Verhoeven et al., 1988; Warren et al., 1990; Svechnikov et al., 2001) and inhibit (Lin et al., 1991; Hales, 1992; Mauduit et al., 1992; 1998; Xiong and Hales, 1994; Lister and Van Der Kraak, 2002; Diemer et al., 2003b; Bornstein et al., 2004) gonadal hormone biosynthesis in vertebrate testes and ovary cells. NF-kappaB, the target of cytokines, has been implicated both in the activation (Delfino et al., 2003; Zhang L et al., 2004) and repression of the androgen receptor promoter (Supakar et al., 1995; Nakajima et al., 1996). The ambivalent role of stress is also reflected by evidence that, dose-dependently, oxidative and nitrosative stress can both enhance (Peltola et al., 1996; Zirkin et al., 1997; Valenti et al., 1999; Andric et al., 2007; Hwang et al., 2007; 2009; Faes et al., 2009) and impair (Behrman and Aten, 1991; Endo et al., 1993; Del Punta et al., 1996; Hales, 2002; Tsai et al., 2003; Ducsay and Myers, 2011) vertebrate male and female gonadal steroidogenesis. Vertebrate male gonadal hormone synthesis is acutely reduced in a number of conditions associated with ROS production and oxidative/nitrosative stress in the testis (Del Punta et al., 1996; Kostic et al., 1998; Hales, 2002; Allen et al, 2004; Chaki et al, 2005; Murugesan et al, 2005; Hanukoglu, 2006; Luo L et al, 2006; Turner and Lysiak, 2008). Examples are cryptorchidism (Chaki et al., 2005), aging (Zirkin and Chen, 2000; Luo L et al, 2006), and ischemia-reperfusion injury (Turner TT et al., 2005).

There is intense cross-talk between the vertebrate hypothalamic–pituitary–adrenal (HPA) axis response to stressors and the hypothalamic–pituitary–gonadal (HPG) axis that controls reproductive activity (Rivier and Rivest, 1991; Rivest and Rivier, 1995; Viau, 2002; Wingfield and Sapolsky, 2003; Mastorakos et al., 2006; Chand and Lovejoy, 2011). Stress- and/or corticosteroid-induced suppression of reproductive functions has been observed in mammals (Stephens, 1980; Moberg, 1985; Armstrong, 1986; Orr et al., 1994; Tilbrook et al., 2000; Manna et al., 2003; Dhanabalan and Mathur, 2009; Hansen, 2009; Chand and Lovejoy, 2011; Nirupama et al., 2013), including humans (Collu et al., 1984; Rabin et al., 1988; Tremellen, 2008), birds (Siegel, 1980), reptiles (Lance and Elsey, 1986; Elsey et al., 1990; 1991; Moore et al., 1991), amphibians (Licht et al., 1983, Moore and Zoeller, 1985; Zerani et al., 1991) and fishes (Pickering et al., 1987; Campbell et al., 1992; Haddy and Pankhurst, 1999; Pankhurst and Van Der Kraak, 2000; Dufour et al., 2005).

Corticosteroids have been found to have effects at every level of the HPG axis (Thibier and Rolland, 1976; Bambino and Hsueh, 1981; Welsh and Johnson, 1981; Mann et al., 1982; Sapolsky, 1985; Wingfield and Sapolsky, 2003). In general, the effect of corticosteroid is to suppress the secretion or action of the various releasing factors or hormones. Considerable evidence demonstrates the importance of sympathetic nervous, catecholaminergic and glucocorticoid control of Leydig cell function and development (Mayerhofer et al., 1990; Mayerhofer, 1996; Hardy et al., 2005). The deleterious effect of corticosteroid administration on reproductive parameters and functions has been described in mammals (Thibier and Rolland, 1976; Bambino and Hsueh, 1981; Welsh and Johnson, 1981; Mann et al., 1982; Sapolsky, 1985), birds (Wilson and Follett, 1976; Petitte and Etches, 1988; 1991), reptiles (Guillette et al., 1995; Knapp and Moore, 1997; Moore and Jessop, 2003), amphibians (Moore and Zoeller, 1985; Kupwaide and Saidapur, 1987; Moore and Jessop, 2003) and fishes (Carragher et al., 1989).

Particularly well studied are the adverse effects of starvation and caloric restriction on reproduction (Heininger, 2012). Metabolic stress induced by caloric restriction reduces gonadal hormones, delays sexual maturation or inhibits reproductive phases in a variety of multicellular organisms from polychaetes to mammals (Schneider and Wade, 1990, Elman and Breier, 1997).

At the cellular level, there is intense cross-talk between activated nuclear receptors, especially the glucorticoid receptor, the estrogen receptor and the androgen receptor, and the activity of NF-kappaB (De Bosscher et al., 2006), which plays a key role in the control of genes involved in cellular stress responses.

14.1 Male reproduction is more susceptible to a variety of stressors

In both plants, invertebrates and vertebrates, compared to female, male gametogenesis and fertility are more sensitive to a variety of stressors, including temperature, radiation, and toxicants, but also immune and psychosocial stress (Gomes, 1970; Le Grande, 1970; David et al., 1971; 2005; Ash, 1980; McGrady, 1984; Orr et al., 1994; Saini, 1997; Rockett et al., 2001; Sikka, 2001; Chakir et al., 2002; Rivier, 2002; Araripe et al., 2004; Rohmer et al., 2004; Vollmer et al., 2004; Jensen et al., 2006; Jørgensen et al., 2006; Gupta et al., 2007; Podrabsky et al., 2008; Sakata and Higashitani, 2008; Hansen, 2009; Crespo and Shivaprasad, 2010; Hales and Robaire, 2010; Prasad et al., 2011; Nirupama et al., 2013). Already Cowles (1965) noted that excessively high but individually non-lethal temperatures may induce total aspermia or a heightened mutation rate. No observations indicate a similar susceptibility in female gamete formation. Somewhat similar susceptibility appears to occur in plants and animals other than the scrotal mammals. High but non-sterilizing heat may induce mutation rates 5 times for every 10°C rise in temperature. The narrow margin between gametogenic heat damage and optimal somatic temperatures is such that local gonadal and total somatic thermoregulation becomes critical in terms of normal numerical and qualitative multigeneration success (Cowles, 1965).

At least in some insects, there is evidence that spermatogenesis is less developmentally stable than oogenesis (Chung, 1962; Temin, 1966; Lindsley and Tokuyasu, 1980; Saccheri et al., 2005). Mutagenesis experiments in Drosophila suggest a 50% greater potential for mutations affecting male than female fertility (Lindsley and Tokuyasu, 1980). Chung (1962) and Temin (1966) showed that, in D. melanogaster, chromosome II homozygotes causing complete sterility were two to three times as common for males as for females. Wolbachia is a maternally transmitted intracellular symbiont that is mainly localized in the reproductive tissues of arthropods and it is responsible for the induction of feminization, parthenogenesis, male-killing and cytoplasmic incompatibility (Saridaki and Bourtzis, 2010). Intriguingly, Wolbachia infection may increase oxidative stress in infected arthropods thus mimicking the environmental stress-induced, oxidative stress-mediated, male infertility (see chapter 15.1).

Small, marginal or colonizing populations go through periods of inbreeding. Inbreeding has adverse effects primarily on male fertility in many animals, including insects and mammals (Saccheri et al., 2005; Asa et al., 2007; Fitzpatrick and Evans, 2009; Zajitschek et al., 2009; Malo et al., 2010; Okada et al., 2011). Drosophila lines homozygous for the second chromosome displayed a significant reduction in male mating ability as a result of inbreeding (Miller et al., 1993; Miller and Hedrick, 1993; Enders and Nunney, 2010). Moreover, inbreeding generally increases the sensitivity of a population to stress, thereby increasing the amount of inbreeding depression (Miller, 1994; Pedersen et al., 2011; Bijlsma and Loeschcke, 2012). The joint effect of inbreeding and stress was more pronounced for male than for female reproductive performance (Enders and Nunney, 2010; Pedersen et al., 2011).

Intriguingly, disruption of a variety of genes involved in stress responses impairs male more than female fertility (Dix et al., 1996; Yue et al., 1999; Nakai et al., 2000; Celeste et al., 2002; Ng et al., 2002; Hsia et al., 2003; Held et al., 2006; 2011; Wright et al., 2007; Coussens et al., 2008; Burnicka-Turek et al., 2009; Grad et al., 2010). As further evidence that the increased stressor vulnerability is restricted to testicular germ cells, Sertoli cells are extremely resistant to many harsh treatments, often surviving well exposure to treatments causing the complete obliteration of male germ cells (e.g., Oakberg, 1959; Clegg, 1963; Bergh, 1981). Oxidative stress is required for sperm motility and viability, capacitation, acrosome reaction, hyperactivation and the fusion of spermatozoa with the oocyte (Aitken et al., 1995; 1998b; Griveau and Le Lannou, 1997; Rivlin et al., 2004; Kothari et al., 2010) but may, on the other hand, be detrimental for these functions and spermatozoan viability (Fujihara and Howarth, 1978; Wishart, 1984; Mammoto et al., 1996; Griveau and Le Lannou, 1997; Sikka, 2001). Even in birds where egg production appears to be more temperature sensitive than semen production, the male bird appears to contribute more to heat stress infertility than the female due to impaired fertilization success of sperm (McDaniel et al., 1995). In plants, the geographic distribution of sexual and asexual species suggests that more harsh environmental conditions favor asexuals (Peck et al., 1998, Eckert et al., 1999; Dorken and Eckert, 2001). Consistent with these observations, environmental stress, particularly cold and water stress, induces male sterility in a multitude of plants (Sawhney and Shukla, 1994; Sheoran and Saini, 1996; Saini, 1997; Sakata and Higashitani, 2008; Zinn et al., 2010) (see chapter 15.2).

Gonadal activity in the male is subject to variation, as evidenced by marked changes in testosterone release as a function of social and reproductive status and in response to stress. Through a glucocorticoid receptor-mediated process, physiological levels of corticosteroids exert a number of deleterious effects on Leydig cells, including inhibition of testosterone biosynthesis, suppression of luteinizing hormone receptor expression and induction of Leydig cell apoptosis (Bambino and Hsueh 1981; Welsh et al., 1982; Cumming et al., 1983; Stalker et al., 1989; Cooke et al., 1992; Orr and Mann, 1992; Monder et al., 1994a; b; c; Gao et al., 1997; Sharp et al., 2007). On the other hand, males typically show reduced adrenocorticotropic hormone (ACTH) and corticosterone responses to stress (Kitay, 1961; Critchlow et al., 1963; Lescoat et al., 1970; Le Mevel et al., 1978, 1979; Kant et al., 1983) compared with that shown in females, an effect that is attributed in part to the central inhibitory effects of testosterone. Gonadectomized males show increased plasma ACTH and corticosterone responses to several forms of stress (Viau and Meaney, 1991; 1996; Handa et al., 1994; Seale et al., 2004). This effect is reversed with testosterone or dihydrotestosterone (DHT), the reduced nonaromatizable form of testosterone, indicating an androgen receptor-mediated effect (Viau and Meaney, 1991; 1996; Handa et al., 1994). Corticotrophs express few if any androgen receptors (Dubois et al., 1978; Thieulant and Duvall, 1985; Viau and Meaney, 1996) and minimal aromatase activity (McEwen, 1980). Thus, the inhibition of stress-induced ACTH release by androgens is unlikely to be explained by direct actions of testosterone or its estrogen byproducts at the level of the pituitary. Evidence suggests central effects of gonadal steroids on HPA activity; females show higher levels of hypothalamic corticotropin-releasing hormone (CRH) (Hiroshige et al., 1973) and increased levels of CRH mRNA in the paraventricular nucleus (PVN) (Watts and Swanson, 1989). This sex difference reflects a stimulatory effect of estrogen on CRH synthesis and release; acute elevations in plasma estrogen levels in females increase CRH mRNA and content in the hypothalamus (Bohler, 1990; Viau and Meaney, 1992; Seale et al., 2004). The demonstration of increased CRH immunoreactivity (Bingaman et al., 1994) in castrated compared to sham male rats suggests testosterone modulates corticosterone release via HPA axis inhibition.

Positive feedback amplifies a signal, whereas negative feedback attenuates it. The male vertebrate HPG and HPA axes are linked in a double-negative feedback loop (Ferrell, 2002). In this circuit, testosterone (A) inhibits or represses the HPA axis (B) and B inhibits or represses A. Thus, there could be a stable steady state with A on and B off, or one with B on and A off, but there cannot be a stable steady state with both A and B on or both A and B off. Such a circuit could toggle between an A-on state (sexual reproduction with functional males) and a B-on state (asexual reproduction) in response to trigger stimuli that impinge upon the feedback circuit (Ferrell, 2002). Once either state has been established, it could persist indefinitely, being reinforced by the double-negative feedback loop, until some trigger stimulus forces the system to the other state. Positive feedback amplifies the signal, whereas negative feedback attenuates it. The female vertebrate HPG axis is linked by a positive feedback to the HPA axis (Bohler, 1990; Viau and Meaney, 1992; Seale et al., 2004), while the latter attenuates the former in a negative feedback loop (see above). Positive feedback in combination with negative feedback can trigger oscillations (Kholodenko, 2006), which could explain the female menstrual cycles in vertebrates. Due to this gender-differential susceptibility, sexual reproduction has a more narrow ecological distribution than asexual reproduction (geographical parthenogenesis, see chapter 15.3.3).

The poor quality of the human ejaculate sets it apart from that of most, if not all, other mammalian species. There is evidence that the human spermatozoon is a cell in crisis (Aitken, 1999; Aitken and Sawyer, 2003; Joffe, 2007; 2010) and is operating near the threshold of error catastrophe as defined in RNA viruses (Holmes, 2003b; Biebricher and Eigen, 2005; Lauring and Andino, 2010; Christophersen, 2013). Thus, several epidemiological studies show that certain paternal occupations, for example, as a welder, painter, auto mechanic, greenhouse worker, or fireman, involving exposure to metals, combustion products, solvents, or pesticides, are associated with altered sperm quality, and an increase in time to pregnancy, spontaneous abortions, birth defects, or childhood cancer (Savitz et al., 1994; Olshan and van Wijngaarden, 2003; Hales and Robaire, 2010). Even in normal fertile specimens, as much as 50% of the ejaculated sperm population may be abnormally formed and a similar proportion may lack motility (Aitken, 1999). In clinical terms, the impoverished nature of human ejaculate is reflected in the dominant role played by the male factor in the etiology of human infertility, defective semen quality being the most frequently defined cause of this condition in humans (Hull et al., 1985; Irvine, 1998). Not only is human semen quality poor, but there is a growing body of evidence that it is getting poorer. The first indication of this came from the meta-analysis of 14,947 normal men from 61 independent centers (Carlsen et al., 1992). This study revealed an approximate halving of ejaculate sperm concentrations (from 113 to 66 x 106 x ml-1) from 1938 to 1990. This general trend has been independently confirmed in a number of separate data sets (Multigner and Spira, 1997).

14.2 Hormonal modulation of oxidative stress as evolutionary tuning knobs of reproductive activity and genetic variation

As final common effector of a variety of stress signaling pathways, oxidative stress modulates the cellular DNA damage/repair/mutagenesis balance and (epi)genetic variation. Environmental stress has to be relayed to the gonads via a variety of neuroendocrine signals, establishing a soma-germline cross-talk. Here exemplarily three hormonal systems, glucocorticoids, thyroid hormones and melatonin, are discussed that impinge upon the HPA and HPG axes and regulate reproductive activity and gonadal oxidative stress. The focus is here on male reproductive activity since it is more dependent on stress and oxidative stress signaling pathways.

14.2.1 Glucocorticoids

Corticotropin-releasing hormone (CRH) and glucocorticoids (GCs) potentially affect vertebrate gonadal function by acting at any one or more of the following levels within the HPG axis: (i) the hypothalamus (to decrease the synthesis and release of GnRH); (ii) the anterior pituitary gland (to decrease the synthesis and release of LH and/or FSH); (iii) the gonads (to modulate steroidogenesis and/or gametogenesis directly) (Michael and Cooke, 1994; Goos and Consten, 2002). The sex-stress/glucocorticoid relationship is non-linear and is described by approximation as inverted “U”-shaped: sex is favored in intermediate stressful environments, while stable stressfree and extreme stressful environments favor asex (Moore and Jessop, 2003). Intriguingly, the same inverted “U”-shaped relationship has been described for glucocorticoids and cognitive performance and neuronal long-term potentiation and primed burst potentiation (Du et al., 2009). Furthermore, considerable data have shown that low doses of glucocorticoids have trophic actions on neuronal branching and survival (Gould et al., 1990), whereas higher doses are detrimental to neuronal survival (Sapolsky, 1992).

CRH is a 41-amino acid straight-chain peptide (Vale et al., 1981) that, during stress, is released from neurons in the paraventricular nucleus and regulates pituitary secretion of ACTH and pro-opiomelanocortin-derived peptides (Plotsky and Vale, 1984; Antoni, 1986). CRH has an antireproductive action in the brain (Sirinathsinghji et al., 1983; Rivier et al., 1986; Petraglia et al., 1987). Moreover, CRH has also been found in the testis of several mammalian species (Yoon et al., 1988; Audhya et al., 1989) and exerts a major regulatory influence on testicular function (Ulisse et al., 1989; 1990; Fabbri et al., 1990; Tinajero et al., 1992; Huang et al., 1995; 1997). Both stimulatory (Huang et al., 1995; 1997) and inhibitory (Ulisse et al., 1989; 1990; Fabbri et al., 1990; Tinajero et al., 1992; Dufau et al., 1993; Frungieri et al., 2002) actions of CRH on Leydig cell steroidogenesis have been reported, consistent with the inverted “U”-shaped stress-sexual reproduction relationship.

Moreover, studies have identified GC receptors in a range of ovarian and testicular cell types and have clearly shown that GCs can exert direct effects on gonadal steroidogenesis, both in vivo and in vitro (Michael and Cooke, 1994). The increased susceptibility of male reproductive activity seems to have both cellular, peripheral and central HPG and HPA components. GCs target a multitude of genes whose products have manifold effects on mitochondrial function and redox balance. GCs interact with intracellular receptors expressed in almost every tissue. In the absence of related ligands, glucocorticoid receptor (GR)alpha is cytoplasmatic and transcriptionally inactive, because it is associated with several proteins (Pratt, 1992; Pratt et al., 1996). Binding of GCs to their receptor induces dissociation of receptor-associated proteins, with subsequent GRalpha activation and translocation into the cell nucleus (Bamberger et al., 1996). Homodimers of the activated receptor modulate the transcription of various (one to two hundred) responsive genes by binding to specific DNA-associated glucocorticoid responsive elements (GREs). The genomic actions also critically depend on recruitment of coactivators and corepressors that may account for many ligand- and cell-specific effects of GCs (van der Laan and Meijer, 2008). Moreover, the cellular redox state regulates GR ligand affinity, nuclear import of the GR and intracellular hormone potency (Makino et al., 1996; Okamoto et al., 1999; Tanaka et al., 1999; Tomlinson et al., 2004; Agarwal and Auchus, 2005; Kitagawa et al., 2007; Sherbet et al., 2007). GREs mediate tissue-specific pleiotropic actions on redox balance by affecting mitochondrial functions (Sionov et al., 2006; Amat et al., 2007; Bjelakovic et al., 2007), a variety of anti-oxidant enzmes (Pereira et al., 1995; José et al., 1997; Lim and Kim, 2009) and oxidases, such as NADPH oxidases (Marumo et al., 1998; Girod and Brotman, 2004; Hsu et al., 2005; Mitchell et al., 2007; Tobias, 2012), cytochrome P450 mono-oxygenases (Pereira et al., 1998; Honkakoski and Negishi, 2000), and monoamine oxidases (Chen, 2004; Shih and Chen, 2004; Manoli et al., 2005; Ou et al., 2006; Chen K et al., 2011; Grunewald et al., 2012).

Modulation of mitochondrial redox and metabolic activities by GCs is biphasic. Short-term exposure to stress concentrations of GCs is associated with induction of mitochondrial biogenesis and enzymatic activity of selected subunits of the respiratory chain complexes, whereas prolonged exposure to GCs causes respiratory chain dysfunction, increased ROS generation, mitochondrial structural abnormalities, apoptosis and cell death, depending on the target tissue energy requirements and developmental stage of the organism (Orzechowski et al., 2002; Duclos et al., 2004; Lin H et al., 2004; Manoli et al., 2005; Alesci et al., 2006; Halliwell and Gutteridge, 2007). Thus, in a variety of tissues GCs have both anti- and pro-oxidant and anti- and proinflammatory actions. At low to moderate doses GCs stimulate spermatogonial proliferation (Milla et al., 2009) and decrease germ cell apoptosis (Yazawa et al., 2001; Mogilner et al., 2006), but at higher levels can inhibit both spermatogenesis and gonadal steroid production (Milla et al., 2009) and induce apoptosis of germ cells, particularly spermatogonia (Yazawa H et al., 1999; 2000; Sasagawa et al., 2001b; Orazizadeh et al., 2010).

Depending on the cellular milieu, heme oxygenase (HO) activity can be considered as an ambiguous redox modulator with both pro-oxidant and anti-oxidant activities (see chapter 7.2.6). The glucocorticoid element is the only demonstrated functional response element in the promoter sequence of HO-2 (McCoubrey and Maines, 1994; Weber et al., 1994; Raju et al., 1997). Intriguingly, HO-2 levels in the testis are controlled by GCs and developmental and tissue-specific factor(s) determine generation of transcripts unique to the organ (Liu N et al., 2000).

Cytochrome P450 aromatase is a key enzyme in the hormonal pathway catalysing the irreversible conversion of sex steroids, androgens to estrogens, and thus is highly relevant to the process of sex change (Gardner et al., 2005; Kroon et al., 2005). In a variety of vertebrates, including fishes, reptiles and birds, estradiol supplements during early development can result in feminization, whereas the inhibition of estradiol synthesis using aromatase inhibitors can result in masculinization (Yu et al., 1993; Lance, 1997; Chardard and Dournon, 1999; Pieau et al., 1999; Crews et al., 2001; Bruggeman et al., 2002; Devlin and Nagahama, 2002; Crews, 2003). Therefore, it is commonly assumed that the regulation of estradiol synthesis by aromatase plays a key role in the sexual development and differentiation of these vertebrates (Lange et al., 2002). Analysis of the regulatory sequences of the aromatase isoform CYP19A1 from the teleost species Gobiodon histrio, a bi-directional sex changer, revealed a number of cis-acting GREs (Gardner et al., 2003; 2005). Experimental exposure to the unfavorable mating scenario of two males only, usually resulted in the smaller male reverting to the female state (Munday et al., 1998). It was suggested that the mechanism by which this operates involves a positive relationship between increased GC concentration (response to stress) and estrogen (required for sex reversal to female) (Gardner et al., 2005). A similar mechanism may operate in reptile temperature-dependent sex determination (Lance, 2009). GCs have been postulated to be critical regulators of the sex change process (Perry and Grober, 2003). This model is based on the link between GCs and social status in many vertebrates, the frequent suppression of reproductive function in subordinate animals, and GC regulation of steroid synthesis (Perry and Grober, 2003).

14.2.2 Thyroid hormones

In almost all taxa, reproductive activity is timed to optimize the chances of survival for the offspring. Seasons with their variable resource availability exert strong selective pressure for this timing. Evidence from different vertebrate groups strongly suggests that thyroid hormone (TH) is crucially required for the expression of seasonal rhythms, with changes in TH signaling being a key element of circannual timing mechanisms (Hazlerigg and Loudon, 2008). TH is an ancient signalling molecule whose function probably originated well before the divergence between the vertebrate and other deuterostome lineages (Hazlerigg and Loudon, 2008). Evidence linking TH to the control of breeding activity can be found in Echinoderms, as well as in the primitive chordate, Amphioxus (Heyland et al., 2005). Seasonally reproductive birds and mammals engage the TH system in the regulation of the seasonal response. Removal of the thyroid gland dramatically altered the seasonal changes in gonadal growth. Thyroidectomy blocked many of the seasonal responses to photoperiod in a variety of bird species (Benoit, 1936; Woitkevitsch, 1940; Nicholls et al., 1988a; Dawson et al., 2001), that remarkably could be restored by a single injection of thyroxine (T4). Subsequent studies in sheep showed that thyroidectomy overcomes the seasonal (or photorefractory) inhibition of reproductive activity in rams in the spring and supports a concept of a key role for thyroid hormones in the expression of seasonal patterns of breeding activity (Nicholls et al., 1988b; Anderson and Barrell, 1998). THs regulate the duration of Sertoli cell proliferation, stimulate their functional maturation, affecting adult Sertoli cell number, and hence the capacity of the testis to produce sperm (Francavilla et al., 1991; Van Haaster et al., 1992; 1993; Buzzard et al., 2000; Holsberger and Cooke, 2005). TH receptors are located on the Sertoli cells in the seminiferous tubules, and possibly Leydig cells, and it is believed that triiodothyronine (T3) binds directly to these receptors (Singh et al., 2011).

The pro-oxidant action of THs may play an essential role in the induction of seasonal reproductive activity. THs increase the metabolic rate, calorigenesis, and exacerbate oxidative stress due to the acceleration of aerobic metabolism (Wilson et al., 1989; Oppenheimer et al., 1996; Venditti et al., 2003; Fernandez et al., 2005). T4 directly stimulates the production of superoxide anion in neutrophils and alveolar macrophages (Kanazawa et al., 1992; Nishizawa et al., 1998). THs have been associated with mitochondrial ROS generation and the induction of oxidative stress in various tissues such as brain, heart, blood, muscle, kidney and liver (Fernandez et al., 1985; Asayama et al., 1987; Zaiton et al., 1993; Venditti et al., 1997; Huh et al., 1998; Sewerynek et al., 1999; Tapia et al., 1999; Shinohara et al., 2000; Karbownik and Lewinski, 2003; Bednarek et al., 2004; Das and Chainy, 2004; Mogulkoc et al., 2005; Venditti and Di Meo, 2006). The THs triiodothyronine (T3) and L-thyroxine sodium salt (T4) produced DNA damage in human sperm mainly via the production of ROS but retained good cell viability (Dobrzynska et al., 2004). Altered thyroid status has been shown to influence several oxidative stress and enzymatic antioxidant defense parameters in rat testis (Choudhury et al., 2003; Mogulkoc et al., 2005; 2006). For example, hyperthyroidism was associated in the rat testis with increased lipid peroxidation, indicative of oxidative stress, increased levels of reduced glutathione (GSH) and increased levels of mitochondrial hydrogen peroxide (Sahoo et al., 2005; 2008; Mogulkoc et al., 2005; 2006). Increased activity levels of most antioxidant defense enzymes have also been demonstrated (Zamoner et al., 2007). These results indicate that TH treatment caused a high oxidative insult to the testis (Dobrzynska et al., 2004; Mogulkoc et al., 2005; 2006; Wagner et al., 2008; 2009) and are consistent with data showing that hyperthyroid tissues exhibit increased ROS production (Venditti and Di Meo, 2006). Conversely, transient hypothyroidism seems to induce oxidative stress in testis by reducing the levels of testicular enzymatic and nonenzymatic antioxidant defenses (Sahoo et al., 2007; Zamoner et al., 2008).

14.2.3 Melatonin

In vertebrates, primary photoreception is by nonvisual irradiance detectors in the retina, pineal, or hypothalamus (Bradshaw and Holzapfel, 2007). In seasonally breeding mammals changes in the photoperiod are used to time their reproductive cycles; temporal signals to the reproductive system are controlled by the circadian rhythm of pineal melatonin secretion with high levels in the dark period (Tamarkin et al., 1985; Pang et al., 1998; Bromage et al., 2001; Malpaux et al., 2001). Melatonin has a well known function in the regulation of circadian and seasonal rhythms (Cassone, 1990; Gwinner et al., 1997). Melatonin release from brain mediates the influence of environmental photoperiods on planarian asexual reproduction (Morita and Best, 1984; Morita et al., 1987), demonstrating that the melatonin signal is evolutionary old and not necessarily associated with sexual reproduction. In fact, melatonin is one of the phylogenetically oldest biological molecules and ubiquitous in microbes, plants and animals (Hardeland and Fuhrberg, 1996; Roopin and Levy, 2012). Numerous extrapineal sites of melatonin synthesis exist, and in some of them, quantities or concentrations considerably exceed those in pineal and blood plasma (Pandi-Perumal et al., 2006; Hardeland and Poeggeler, 2007). In extrapineal sites, with exception of the retina and, where present, the parietal organ, circadian rhythms may exhibit low amplitudes or even be virtually absent, and the transmission of dark signals seems to be rather unlikely in organs like bone marrow or gastrointestinal tract. The conclusion has to be that melatonin plays a number of different roles (Pandi-Perumal et al., 2006; Hardeland and Poeggeler, 2008), roles that have been changing during evolution (Tan et al., 2010).

Whether the day-light signal is interpreted as anti- or progonadotropic will depend on the species (long-day seasonal breeders including the hamsters, short-day seasonal breeders such as sheep, or nonseasonal breeders like humans), the duration of night melatonin peak (the duration hypothesis), the magnitude of the night melatonin peak (the amplitude hypothesis), and/or the window of sensitivity to melatonin (the internal coincidence hypothesis) (Arendt, 1988; Stankov and Reiter, 1990; Reiter, 1991; Lincoln, 2002). The hypothalamic–pituitary axis plays a key role in the regulation of reproduction by melatonin. In addition to the hypothalamus and pituitary, melatonin receptors in the testis, epididymis, vas deferens, prostate, ovary and mammary gland suggest the concept of multiple sites of melatonin action on the reproductive system (Pang et al., 1998). Melatonin is synthetized in the testes (Tijmes et al., 1996; Kato et al., 1999; Fu et al., 2001; Stefulj et al., 2001) and melatonin receptors are expressed in rodent and avian testes (Ayre and Pang, 1994; Vera et al., 1997; Valenti et al., 2001; Frungieri et al., 2005; Izzo et al., 2010), but both testicular melatonin receptor expression and melatonin synthesis decline during rat aging (Sánchez-Hidalgo et al., 2009a; b).

Melatonin has been shown to alleviate oxidative stress in the testes following the experimental induction of unilateral varicocele and torsion-detorsion (Semercioz et al., 2003; Abasiyanik and Dagdönderen, 2004; Yurtçu et al., 2008) and other stressors (Gavazza and Catala, 2003; Bustos-Obregón et al., 2005; Armagan et al., 2006; Atessahin et al., 2006; Kara et al., 2007; Sönmez et al., 2007; Guneli et al., 2008; Rao and Gangadharan, 2008; Huang et al., 2009; Sarabia et al., 2009a; b; Espino et al., 2010). Pinealectomy aggravates the testicular oxidative damage as a consequence of induced hyperthyroidism (Mogulkoc et al., 2006). Importantly, melatonin may protect germ cells against DNA damage (Sarabia et al., 2009a) as is also indicated by the finding that high endogenous melatonin concentrations enhance sperm quality (Ortiz et el., 2011). Melatonin levels in seminal plasma are depressed in infertile patients exhibiting poor motility, leukocytospermia, varicocele and non-obstructive azoospermia, all of which are conditions associated with oxidative stress in the male tract (Awad et al., 2006).

Melatonin and its metabolites reduce oxidative stress in vitro and in vivo (Tan et al., 1993; 2003; Reiter et al., 2000; 2001; 2009; Rodriguez et al., 2004; Hardeland, 2005). Melatonin has two major attributes that set it apart from most other antioxidants. Firstly, it undergoes a two electron oxidation when acting as antioxidant, rather than the one electron oxidation favored by many free radical scavengers. As a result, melatonin cannot redox cycle and inadvertently generate free radicals (Hardeland, 2005; Aitken and Roman, 2008). Melatonin has been shown to both scavenge free radicals (Tan et al., 1993; 2002), to improve mitochondrial malfunction and H2O2 generation (Hardeland, 2005; Leon et al., 2005), to stimulate a number of antioxidant enzymes including superoxide dismutase, glutathione peroxidase and glutathione reductase and to down-regulate prooxidant enzymes, such as nitric oxide synthases and lipoxygenases (Reiter et al., 2000; Anisimov, 2003; Semercioz et al., 2003; Rodriguez et al., 2004; Hardeland, 2005).

The hormone is important for the synchronization of reproductive response to appropriate environmental conditions in animals. In general, the annual changes in pineal melatonin secretion drive the reproductive responses of photoperiodic mammals (Bronson, 1989). In birds, the data on the photoperiodic action of melatonin are not so clear-cut (Tsutsui et al., 2013), possibly because there has been evidence that the avian hypothalamus can synthesize melatonin de novo (Kang et al., 2007). Yet, there are data available on regulation of seasonal processes by melatonin, including but not limited to that of gonadal activity and gonadotropin secretion (Ohta et al., 1989; Bentley et al., 1999; Bentley and Ball, 2000; Guyomarc’h et al., 2001; Rozenboim et al., 2002). In birds, mammals and other vertebrates, gonadotropin-inhibitory hormone (GnIH) is a neuropeptide that inhibits gonadotropin synthesis and release, indicating a conserved role for this neuropeptide in the control of the HPG axis across species. In birds, melatonin appears to act on GnIH neurons in stimulating not only GnIH synthesis but also its release, thus inhibiting plasma LH concentration (Chowdhury et al., 2013; Tsutsui et al., 2013). Recently, a similar, but opposite, action of melatonin on the inhibition of expression of mammalian GnIH was shown in hamsters and sheep, photoperiodic mammals. These results in photoperiodic animals demonstrate that GnIH expression is photoperiodically modulated via a melatonin-dependent process. Recent findings indicate that GnIH may also be a mediator of stress-induced reproductive disruption in birds and mammals (Tsutsui et al., 2013). Melatonin implants have been shown to delay the onset of clutch initiation in female great tits but without affecting clutch size, body mass, or timing of onset of activity (Greives et al., 2012). In other species, GnIH and GnRH expression are positively correlated (Bentley et al., 2003; Calisi and Bentley, 2009). In some species, GnIH may act to induce a temporary pause in reproductive effort without full HPG axis suppression and regression, which would cause the individual to miss the opportunity to breed in that year. In their natural environments, this temporary cessation of reproduction occurs in response to unpredictable environmental cues, such as stress, or mate/nest availability (Calisi et al., 2011). Thus, hypothalamic GnIH may serve as a short-term modulator of reproductive behaviors in response to social environment (Calisi et al., 2011).

Melatonin up-regulates the expression of GnIH mRNA in starling gonads before breeding. In vitro, melatonin significantly decreases testosterone secretion from LH/FSH-stimulated testes before, but not during, breeding (McGuire et al., 2011). In the frog Rana esculenta, melatonin severely inhibits both control and GnRH-induced Leydig cell testosterone secretion in vitro (d’Istria et al., 2004) and has a direct inhibitory effect on the mitotic activity of primary spermatogonia in the testis (d’Istria et al., 2003). Physiological crosstalk between melatonin and glucocorticoid receptor is of high adaptive significance in wild animals for balancing immunity and oxidative stress during ecologically stressful conditions (Nelson and Drazen, 1999; Gupta and Haldar, 2013). Seasonal changes in plasma melatonin and glucocorticoids modulate the effect of glucocorticoids in the testis of anuran Rhinella arenarum (Regueira et al., 2013).

Overall, HPA and HPG axes cross-talk on a multitude of levels to adapt sexual reproductive activity and (epi)mutagenesis as potential bet-hedging responses to environmental conditions.

15. Asexual reproduction


…there are regimes where the advantages for sex are outweighed by its disadvantages. A complete theory for the existence of sex must be able to identify the regimes where either sexual or asexual replication are respectively dominant, in a manner that is consistent with observation.

E. Tannenbaum, 2008

Summary

Ecological conditions largely determine the evolutionary benefits of asexual and seual reproduction. Populations of clonal organisms are often represented as being evolutionary inert with persistent genetic fidelity. However, molecular data from viruses, prokaryotes and eukaryotes support the argument that asexual taxa can maintain significant levels of genetic diversity and adaptability and can create increased genetic variation in response to stress. Asexuals have a variety of tools that can, often stress-dependently, generate genetic and phenotypic variation without the use of meiotic recombination: transposon activation, mutability of simple sequence repeats, prions, polyploidy, epigenetic modulation, automixis and mitotic recombination. Large population sizes and seed/egg banks facilitate bet-hedging strategies both within and between generations.

Modular organisms have an arsenal of strategies with which they can create non-sexual genotypic/phenotypic plasticity: somatic mutagenesis, polyploidy, seed banks, and epigenetic variation. Importantly, somatic mutations may be transmitted to the offspring both sexually and asexually. As result, plants and sessile aquatic animals often exhibit labile asex/sex expression and rather exhibit a near-continuum between sexuality and asexuality. Environmental stress, particularly cold and water stress, induces male sterility in a variety of plants. Clonal reproduction appears to offer a safe escape route for many plant species under suboptimal environmental conditions. Ancient asexuals violate the expectation that sex and recombination are necessary for long-term survival. The microscopic size and resource-rich habitats of bdelloid rotifers, Darwinulid ostracods and oribatid mites keeps the relative investment in new organisms low and allows genetic exploration of fitness landscapes by trial-and-error expeditions at the population level. Cyclical parthenogens alternate phases of asexual propagation with bouts of sexual reproduction and are of special interest because they allow a direct assessment of the ecological conditions, costs and benefits of sex versus asex. Generally, seasonal and food stress induce sexual reproduction. The phenomenon that parthenogenesis, and not outcrossed sexuality, prevails in harsh, uncertain, disturbed and novel conditions has been termed geographical parthenogenesis. As discussed in chapter 14.1 male reproduction is more susceptible to environmental disturbance and together with the superior colonizing ability of asexuals this explains the ecological distribution of sex/asex.

Traditionally, benefits of sex have been grouped into ecological and genetic explanations (Kondrashov, 1993; West et al., 1999; Hörandl, 2009; Otto, 2009; Hartfield and Keightley, 2012). If the prevalence of sex means it is so beneficial, how can asexuals — provided they are real — survive? By extension it must mean that the selective pressures that cause sex to persist and prevail are somehow less powerful or counteracted by even larger benefits of asexuality in these organisms. So, in an ideal world, a comparison of the genetics and ecology between asexual and sexual species could be expected to yield some hints as to what needs to change within the organism or in its environment to make either asexuality or sexuality the more successful strategy (Maderspacher, 2011; Neiman and Schwander, 2011).

Hereafter, like Vrijenhoek and Parker (2009), I use the term parthenogenesis in a broad sense for all-female clonal reproduction while, in the strict sense, it is the development of an egg without fertilization. The origin of parthenogenesis is polyphyletic in many taxa, suggesting that genetic systems maintaining sexuality are often labile. Asexual lineages are known to arise not infrequently within species, but most of the time perish quite quickly (Simon JC et al., 2003). Strict asexuality is thought to be an evolutionary dead-end in ‘higher’ organisms and that most asexual metazoans only form twigs on the tree of life (Schwander and Crespi, 2009). Evolutionary theory predicts that obligate asexuals have a long-term evolutionary disadvantage, compared with sexuals, owing to a more pronounced ‘Hill-Robertson effect’, a reduction in the efficacy of natural selection that occurs because finite populations accumulate associations of linked genes (haplotypes) that interfere with selection (Hill and Robertson, 1966, Felsenstein, 1974). Absent or very infrequent recombination reduces Ne, and thus the efficacy of selection against deleterious mutations (Ohta and Kimura, 1971), by increasing selective interference from linked loci (Hill and Robertson, 1966; Birky and Walsh, 1988; McVean and Charlesworth, 2000; Marais and Charlesworth, 2003; Charlesworth, 2012). Empirical evidences for such mechanisms were higher rates of non-synonymous mutation accumulation in asexual strains of Daphnia water fleas (Paland and Lynch, 2006) and Potamopyrgus snails (Neiman et al., 2010) compared to their sexual counterparts. This indeed demonstrated inefficient purifying selection associated with clonal reproduction. Thus, it is generally assumed that the higher mutation load in the mitochondrial genome and in the asexual nuclear genomes results from a lack of recombination (e.g. Bell, 1988b; Jansen and de Boer, 1998; Stewart et al., 2008b). A second problem is that the genetic uniformity of the offspring leads to a much lower genetic diversity, which is likely to make it much more difficult to adapt, for example, to changing environments or to parasites; consequently, classical theory holds that asexual species should be slow to evolve (Maynard Smith, 1978a; Bell, 1982).

15.1 Non-sexual ways to increase evolvability

In the framework of the fitness landscape described by Wright (1931), populations placed near the top of a fitness peak will experience less beneficial mutations than populations placed far from the adaptive optimum (Fisher, 1930; Lenski and Travisano, 1994; Elena and Lenski, 2003; Lázaro et al., 2003; Silander et al., 2007; Stich et al., 2010). This dependence of mutation effects on the degree of adaptation of populations led to the theoretical prediction that mutation rates would be reduced in constant environments, in which the population has had enough time to adapt. Once the optimum has been attained, a homogeneous population of individuals with the optimal phenotype is the best adaptive solution, so the generation of further diversity is not necessary. Thus, stable environments would favor low mutation rates (anti-mutator genotypes), constrained only by the costs of error-repair mechanisms (Kimura, 1967; Drake, 1991).

Populations of clonal organisms are often represented as being evolutionary inert with persistent genetic fidelity. If such a biological entity as a ‘clone’ really did exist, it would be a fantastic entity, differing from everything else known in biology, i.e. it would possess a population mean but no variance for any particular trait. It would not be amenable to selection and adaptive variation and would thus be unchanging in time and space (Loxdale and Lushai, 2003) – and due to this evolutionary inertness highly susceptible to extinction. Reproductively favored clones should rapidly eliminate clones with lower fitness, leading to the erosion of genetic (i.e., clonal) diversity (Vrijenhoek, 1978; Fox et al., 1996; Weeks and Hoffmann, 1998). However, natural selection in temporally and spatially heteterogeneous environments is a strong force that shapes dynamics of genotypic diversity (Herbert and Crease, 1980; Harshman and Futuyma, 1985; Browne and Hoopes, 1990; Geedey et al., 1996; Weeks and Hoffmann, 1998; 2008; Niklasson et al., 2004; Vorburger, 2005a). Molecular data from viruses, prokaryotes and eukaryotes support the argument that clones can possess a highly dynamic and adaptive genome (Parker, 1979a; b; Vrijenhoek, 1979; Herbert and Crease, 1980; Ochman et al., 1980; Bell, 1982; Hughes, 1989; Moritz et al., 1989; Browne and Hoopes, 1990; Widén et al., 1994; Fox et al., 1996; Lushai et al., 1998; 2000; 2003; Weeks and Hoffmann, 1998; Wilson et al., 1999; 2003; Kawamura and Fujiwara, 2000; Wolf AT et al., 2000; Delmotte et al., 2002; Lushai and Loxdale, 2002; Loxdale and Lushai, 2003; Schön et al, 2003; Niklasson et al., 2004; Baali-Cherif and Besnard, 2005; Vorburger, 2005a; Castagnone-Sereno, 2006; Terhivuo and Saura, 2006; Hörandl and Paun, 2007; Charaabi et al., 2008; Heethoff et al., 2009; Birky et al., 2010; Bode et al., 2010; Danchin et al., 2011b; Johnson MT et al., 2011; Monti et al., 2012). Thus, asexual organisms can maintain significant levels of genetic diversity and adaptability (White, 1970; Angus and Schultz, 1979; Turner et al., 1983; Good and Wright, 1984; Densmore et al., 1989; Hedges et al., 1992; Quattro et al., 1992; Haddal et al., 1994; Tinti and Scali, 1996; Vrijenhoek, 1998; Johnson et al., 1999; Halkett et al., 2005; Loxdale, 2008; 2010; Janko et al., 2011; Johnson MT et al., 2011). A group of asexual animals, the bdelloid rotifers, has diversified into distinct species broadly equivalent to those found in sexual groups and displays two fundamental properties of species, independent evolution and ecological divergence by natural selection. Thus, sex is not a necessary condition for creating phenotypic and genetic diversity and speciation (Oliver and Herrin, 1976; Atchley, 1977a; b; 1978; Parker, 1979a; b; Ellstrand and Roose, 1987; Fox et al., 1996; Mitton and Grant, 1996; Gorokhova et al., 2002; Barraclough et al., 2003; González et al., 2003; Birky et al., 2005; 2010; Fontaneto et al., 2007; 2009; Heethoff et al., 2007; Hillis, 2007; Strasburg et al., 2007; Kaya et al., 2009). Like sexually reproducing organisms, asexual taxa can create increased genetic variation in response to stress (Goho and Bell, 2000; Doroszuk et al., 2006; van Oppen et al., 2011; Berman and Hadany, 2012). A wide range of asexual taxa have been shown to undergo rapid genetic changes, including root-knot nematodes (Castagnone-Sereno, 2006), crustaceans (Schön et al, 2003) and insects (Wilson et al., 2003). In obligate parthenogenetic New Zealand Sitobion spp. aphid populations, genotypes were found highly heterozygous, whilst there was extensive turnover of genotypes, such that 36% were not sampled again but 42% were new after two years, suggesting clonal selective sweeps to be operative with some local persistence (Wilson et al., 1999).

In genetically clonal populations, phenotypic diversity in fluctuating environments is generated by stochastic phenotype-switching mechanisms (Soll and Kraft, 1988; Moxon et al., 1994; Pérez-Martín et al., 1999; Bayliss et al., 2001; Lachke et al., 2002; Bonifield and Hughes, 2003; Balaban et al., 2004; Kearns et al., 2004; van der Woude and Bäumler, 2004; Kussell and Leibler, 2005). There are a variety of tools that can, often stress-dependently, generate genetic and phenotypic variation without the use of meiotic recombination, such as transposons (Nevers and Saedler, 1977; Biel and Hartl, 1983; Chao et al., 1983; Modi et al., 1992; Wilke and Adams, 1992; Bowen and Jordan, 2002; Lankenau and Volff, 2009; Maumus et al., 2009; Oliver and Greene, 2009a; Zhang and Saier, 2009; Upton et al., 2011), simple sequence repeats, also called microsatellites and minisatellites (Kashi et al., 1997; von Sternberg, 2002; Kashi and King, 2006; King and Kashi, 2007; King, 2012a; b), and prions (True and Lindquist, 2000; Masel and Bergman, 2003; Tyedmers et al., 2008; Halfmann and Lindquist, 2010; Lancaster et al., 2010). Evolutionary theory predicts that ancient asexuals should contain fewer functional retrotransposons than sexuals or even none at all, because sex is thought to be necessary to spread these transposons within populations (Hickey, 1982; Schön and Martens, 2000; Arkhipova and Meselson, 2005; Schön et al., 2009). Except for bdelloid rotifers (Arkhipova and Meselson, 2000) and yeast (Zeyl et al., 1996), however, this pattern could not be confirmed. Schön and Arkhipova (2006) found one group of degenerated retrotransposons, Daphne, in Darwinula stevensoni, and a second transposon group, Syrinx, that was still functional and had obviously recently been active. Other asexuals such as the fungi Candida albicans (Mattews et al., 1997; Goodwin and Poulter, 2000), Entamoeban protozoans (Pritham et al., 2005) and Foraminifera (Maumus et al., 2009) have also been shown to harbor functional retrotransposons in high numbers.

At the edge of life the adaptive road is becoming increasingly narrower. Thus, while moderate stress increases genetic diversity, extreme stressful environments decrease genetic diversity (Kis-Papo et al., 2003; Sonjak et al., 2007; de los Ríos et al., 2010; Vinogradova et al., 2011). Likewise, spontaneous mutation rates in extremophiles (Battista, 1997; Grogan et al., 2001; Grogan, 2004; Mackwan et al., 2007) and sexual reproduction of fungi (Kis-Papo et al., 2003) are reduced in extremely stressful habitats. In H. volcanii and the polyploid methanogenic archaeon M. maripaludis, gene conversion leads to a fast and efficient equalization of genome copies (Hildenbrand et al., 2011; Lange et al., 2011; Soppa, 2011).

Different ways have been described, in which parthenogenetic lineages may originate from sexual species (Simon JC et al., 2003). The majority of the cases of parthenogenesis studied today are associated with bacterial infections (Lorenzo-Carballa and Cordero-Rivera, 2009). Three different microorganisms have been found associated with parthenogenesis in arthropods: Wolbachia (Stouthamer and Werren, 1993; Stouthamer et al., 1993; 2010; Werren, 1997; Arakaki et al., 2001; Weeks and Breeuwer, 2001) and Rickettsia (Hagimori et al., 2006), both members of the a-proteobacteria group; and Cardinium, a member of the Cytophaga–Flexibacter–Bacteroides group (Zchori-Fein et al., 2001; Weeks et al., 2003; Zchori-Fein and Perlman, 2004; Provencher et al., 2005). In addition, an endosymbiont from the Verrucomicrobia group has been found associated with parthenogenesis in a nematode species (Vandekerckhove et al., 2000). Most of the cases of Wolbachia-induced parthenogenesis in haplodiploid Hymenoptera cause diploidization of unfertilized haploid eggs, which develop into females. This occurs through different forms of gamete duplication, which result in the production of fully homozygous progeny (Suomalainen et al., 1987; Stouthamer and Kazmer, 1994; Gottlieb et al., 2002; Pannebakker et al., 2004). Intracellular Wolbachia induces oxidative stress in insect cell lines and in vivo (Brennan et al., 2008; Pan et al., 2012). Wolbachia may have a key role in the infected insect’s ferritin expression and iron metabolism modulating their regulation of oxidative stress (Kremer et al., 2009; Saridaki and Bourtzis, 2010). Intriguingly, Wolbachia infection is associated with a higher amount of the base 8-oxo-deoxyguanosine in DNA and strand breaks in meiotic spermatocytes suggesting that ROS-induced DNA damage in sperm nuclei may contribute to the modification characteristic of cytoplasmic incompatibility, a form of male-derived zygotic lethality (Brennan et al., 2012). Thus, Wolbachia infection may mimick the environmental stress-induced, oxidative stress-mediated, male infertility (see chapter 14.1). On the other hand, the increased oxidative stress may increase oocyte genetic variation or expose otherwise cryptic genetic variation due to the concomitant downregulation of heat shock proteins (Xi et al., 2008). This increased genetic variation may make asexual reproduction in infected insects a viable strategy.

15.1.1 Large population size

In large populations, Muller’s ratchet occurs more slowly, and even lower rates of recombination will effectively arrest mutation accumulation (Charlesworth et al., 1993). Obviously, a risk spreading, bet-hedging strategy is more promising in larger populations: the more lottery tickets are “bought” the higher are the chances to have a “winner”. That’s why unicellular organisms with their large population sizes are the ideal players for these evolutionary games (see chapter 4.2). For multicellular organisms, the large investment of scarce resources into mature organisms limits population sizes and makes this approach only feasible for microscopic multicellular organisms, e.g. Bdelloid rotifers, Darwinulid ostracods, orbatid mites, in less resource limited habitats. On the other hand, large populations of extremely old clonal plants have been reported (Steinger et al., 1996; Lynch AJJ et al., 1998; Reusch et al., 1999; Arnaud-Haond et al., 2012).

For small asexual populations and low mutation rates, the speed of adaptation is primarily limited by the availability of beneficial mutations: a mutation has the time to reach fixation before the next mutation occurs. Therefore, in this case the speed of adaptation increases linearly with population size and mutation rate. By contrast, for large asexual populations or high mutation rates, beneficial mutations are abundant. In this case, the main limit to adaptation is that many beneficial mutations are wasted: when arising on different genetic backgrounds, they cannot recombine and thus are in competition with each other. The recent works can be broadly categorized into two classes: (i) so-called ‘‘clonal interference models’’ and (ii) models in which all mutations have the same effects. The clonal interference models (Gerrish and Lenski 1998; Orr, 2000a; Campos and de Oliveira, 2004; Campos et al., 2004, 2008; Wilke, 2004; Rosas et al., 2005; de Visser and Rozen, 2006; Park and Krug, 2007; Campos and Wahl, 2010; Sniegowski and Gerrish, 2010; Good et al., 2012) emphasize that different beneficial mutations have different-sized effects and that mutations with large beneficial effects tend to outcompete mutations with small beneficial effects. In the clonal interference models (Gerrish and Lenski, 1998; Wilke, 2004), the fixation probability of a beneficial mutation decreases monotonically with increasing N as a result of competition among beneficial mutations leading to an everincreasing advantage of sex. Clonal interference also drags out fixation events, providing time for further beneficial variants to arise from competitors before any of them have fixed (Desai et al., 2007). There is a small conceptual flaw in this derivation which is that the possibility that other beneficial mutations were segregating before the initial appearance of the focal individual was neglected (Patwa and Wahl, 2008). If many mutations are segregating simultaneously, the focal beneficial mutation is likely to have arisen on the background of a previously segregating beneficial mutation. Thus mutations may sweep in groups, the ‘multiple mutation’ regime (Tsimring et al., 1996; Kessler et al., 1997; Rouzine et al., 2003, 2008; Beerenwinkel et al., 2007; Desai and Fisher, 2007; Desai et al., 2007; Zeyl, 2007; Brunet et al., 2008; Park SC et al., 2010). The latter type of models, also dubbed traveling-wave theory, emphasizes that in large populations, multiple beneficial mutations frequently occur in quick succession on the same genetic background. The speed of adaptation in these asexual populations is determined by the emergence and subsequent establishment of mutants that exceed the fitness of all sequences currently present in the population. Rather than approaching a limit for large N, the speed grows as lnN in the regime of practical interest, reflecting the increasing spread of the population distribution along the fitness axis (Park SC et al., 2010). The models can be tuned to account for finite population sizes and determine how quickly populations adapt as a function of population size and mutation rates (Hallatschek, 2011). Conceptually, the multiple mutation regime lies on a continuum between clonal interference and quasispecies dynamics. The ‘‘clonal interference models’’ that focused on two loci only were implicitly assuming monomorphism at all the other loci in the genome. The calculation of the waiting time until the occurrence of a second mutation at a second locus, considered possible beneficial mutations only at a given second locus (Christiansen et al., 1998), whereas it would be more reasonable to think of the chance of a second mutation at any one of thousands of other loci. This would increase by several orders of magnitude the probability that beneficial mutations would be selected concurrently, making the conditions for an advantage of recombination much less stringent than is often assumed (Albu et al., 2012). Although there will still be rapid changes in allele frequencies as expected for periodic selection, now there will be many lineages transiently rising and falling as they continue to mutate and compete. There is growing evidence from multiple experimental approaches for exactly these sorts of ‘multiple mutation’ dynamics (Rozen et al., 2002; Desai et al., 2007; Kao and Sherlock , 2008; Barrick and Lenski, 2009; Betancourt, 2009; Kvitek and Sherlock, 2011; Lang et al., 2011; Herron and Doebeli, 2013; Lee and Marx, 2013; Marx, 2013). On the other hand, the clonal interference model has also experimental support (Atwood et al., 1951; de Visser et al., 1999; Miralles et al., 1999; Shaver et al., 2002; de Visser and Rozen, 2005; Hegreness et al., 2006; Cooper, 2007; Kao and Sherlock , 2008; Lee and Marx, 2013). Possibly, both models can predict the evolutionary dynamics in asexual populations dependent on the mutation rate, population size and the mutation’s selection coefficient (Bollback and Huelsenbeck, 2007; Kao and Sherlock , 2008; Sniegowski and Gerrish, 2010). Kim and Orr (2005) modeled the dynamics of adaptation in sexuals versus asexuals at 2 beneficial sites to determine the effect of recombination on the rate of adaptation. Importantly, (i) the relative difference in fixation time of the second beneficial mutation, between sexuals and asexuals, is small when populations are large with a high mutation supply; and (ii) as the mutation rate increases asexual populations behave progressively more like sexual populations. Consequently, there is no advantage to sex in an infinite population (assuming no epistatic fitness interaction between loci) (Maynard Smith, 1968; Eshel and Feldman, 1970) or small populations; instead, sex should have the greatest effect in populations of intermediate size (Otto and Barton, 2001; Kim and Orr, 2005).

15.1.2 Polyploidy

Genome doubling, or polyploidy, is a major factor accounting for duplicate genes found in eukaryotic genomes. Polyploidy can be regarded as another mechanism to increase both evolvability and robustness (see chapter 13). Polyploid giant cells can be formed by parasexual behavior in bacteria, algae, yeasts and other protozoans under various stress conditions (Rink and Partke, 1975; Urushihara, 1992; Mares et al., 1993; Akerlund et al., 1995; Pecoraro et al., 2011). Polyploidy is another source of genetic novelty that doesn’t depend upon sex, although when it is passed on sexually it can create new organisms that are much more genetically unique, than after regular recombination alone (Niklas, 1997). Polyploidy can create organisms that are so genetically different from their parents that they are unable to breed with organisms from their parent species (Clarke, 2012). Polyploids often show novel phenotypes that are not present in their diploid progenitors or exceed the range of the contributing species (Ehrendorfer, 1980; Levin, 1983; Ramsey and Schemske, 2002). Polyploids exhibit progressive heterosis. Hybrid vigor or heterosis is evolutionarily defined as that the heterozygotes have a higher fitness in a population than the homozygotes and refers to the phenomenon that progeny of diverse varieties of a species or crosses between species exhibit greater biomass, speed of development, and fertility than both parents (Tunner and Nopp, 1979; Wetherington et al., 1987; Comai, 2005; Birchler et al., 2010; Chen, 2010). Genome polyploidization may result in unique gene combinations, genome rearrangements, altered gene expression patterns (including silencing and up- or downregulation of one of the duplicated genes), enlarged body, or cell size and elevated levels of heterozygosity and genetic diversity (Song et al., 1995; Parker and Niklasson, 2000; Wendel, 2000; Soltis and Soltis, 2000; Osborn et al., 2003; Soltis et al. 2004; Luttikhuizen et al., 2007; Leitch and Leitch, 2008). Genetic and epigenetic changes such as sequence loss, methylation-dependent regulation of duplicated genes and re-activation of epigenetically silenced transposons, are common consequences of polyploidization across a wide range of species (Comai et al., 2000; Wendel, 2000; Lee and Chen, 2001; Shaked et al., 2001; Kashkush et al., 2002; Madlung et al., 2002; 2005; Osborn et al., 2003; Wang J et al., 2004; Wang YM et al., 2004; Lukens et al., 2006; Chen, 2007; Paun et al., 2007; Doyle et al., 2008; Hegarty and Hiscock, 2008; Leitch and Leitch, 2008; Soltis and Soltis, 2009; Verhoeven et al., 2010b; Xiong et al., 2011).

Estimates are that at least 50%, and perhaps even 90%, of flowering plants have undergone one or more episodes of chromosomal doubling in their evolutionary history (Masterson, 1994; Otto and Whitton, 2000; Cui et al., 2006; Jiao et al., 2011). Gene duplication events were intensely concentrated around 319 and 192 million years ago, implicating two polyploidization events in ancestral lineages shortly before the diversification of extant seed plants and extant angiosperms, respectively. Significantly, these ancestral whole-genome duplications resulted in the diversification of regulatory genes important to seed and flower development, suggesting that they were involved in major innovations that ultimately contributed to the rise and eventual dominance of seed plants and angiosperms (Jiao et al., 2011). Especially harsh conditions or periods of climatic change might affect the rate of formation, establishment, persistence and long-term evolutionary success of polyploids in angiosperms (Fawcett and Van de Peer, 2010). Hybridization and polyploidization can trigger DNA methylation changes in plants (Adams and Wendel, 2005; Dong et al., 2006; Chen, 2007; Paun et al., 2007). Epigenetic mechanisms may play a key role for gene expression, phenotypic variation and instability in plant polyploids (Comai, 2005; Chen, 2007).

The impact of deleterious mutations can be reduced when there are several copies of a single gene (Tautz, 1992; Wilkins, 1997; Vorburger, 2001; Krakauer and Plotkin, 2002). Masking of deleterious mutations provides an immediate advantage to higher ploidy levels (Mable and Otto, 2001; Vorburger, 2001; Gerstein and Otto, 2009). Moreover, deleterious mutations can be eliminated in a genome copy number-dependent manner by gene conversion that is biased for the wild type (Khakhlova and Bock, 2006; Gaeta and Pires, 2010; Soppa, 2011). At higher male to female mutation ratios, and sufficiently large population sizes, hybridogenetic populations can carry a lower mutation load than sexual species. In smaller populations, the same mechanism reduces the speed of Muller's ratchet. Schultz (1969) proposed that hybridogenesis may act as a transition state in the formation of new species, and Vrijenhoek et al. (1989) found some evidence for such an event in a sexual species of Poeciliopsis with supposed hybridogenetic ancestry. Lower mutation accumulation in hybridogenetic populations opens the possibility that hybridogenetic species can develop into new sexual species once recombination is re-established and reproductive isolation from sexual ancestors has occurred (Som and Reyer, 2007; Som et al., 2007).

The emergence of new phenotypic and molecular variation shortly after polyploid formation has been documented, offering unique avenues for phenotypic response to selection (Orr and Otto, 1994; Song et al., 1995; Jiang et al., 1998; Soltis and Soltis, 1999; Osborn et al, 2003). It has been argued that polyploidy makes clones more tolerant to abiotic stress and gives them wider ecological tolerances and higher fitness than their sexual ancestors possess (Lewis, 1980; Weider, 1993a; Kearney et al., 2005), i.e. as general-purpose genotypes (Lynch, 1984; Weider, 1993a). Likewise, theory suggests that ploidy level can profoundly influence parasite resistance and host–parasite coevolution (Nuismer and Otto, 2004; Oswald and Nuismer, 2007). Host–parasite interactions have also been shown to exert strong selection on the underlying genes that modulate species interactions, e.g. favoring changes in ploidy level (Oswald and Nuismer, 2007). However, empirical data to support theory are scarce (Stover, 1986; Nuismer and Thompson, 2001; King et al., 2012). Higher ploidy may even be harmful (D’Souza et al., 2005).

In natural strains of Saccharomyces cerevisiae isolated from several natural populations at the “Evolution Canyon” microsite (Nahal Oren, Mt. Carmel, Israel), tetraploid strains were more tolerant to all DNA-damaging agents than their neighboring diploid strains (Lidzbarsky et al., 2009). Wider ecological tolerances also allow wider geographic distribution of polyploids (Adolffson et al., 2010) and invasion and occupation of new and harsh environments (Stebbins, 1950; Levin, 1983). In line with this hypothesis, examples of ploidy elevation along a south-north cline are common (e.g., Beaton and Hebert, 1988; Ward et al. 1994; Little and Hebert 1997; Luttikhuizen et al. 2007) supporting the view that polyploidy is selected for in extreme environments (Beaton and Hebert, 1988; Ward et al., 1994). In the diploid–tetraploid species pair of gray treefrogs, Hyla chrysoscelis and H. versicolor (Anura: Hylidae), the tetraploid species almost exclusively occupied areas of higher elevation, where climatic conditions were relatively severe (colder, drier, greater annual variation). In contrast, the diploid species was restricted to lower elevations, where climatic conditions were warmer, wetter and exhibited less annual variation (Otto et al., 2007).

However, an unequivocal test of the direct advantage of polyploidy in harsh environments has proven difficult to produce because of other factors that are closely associated with polyploidy. For example, polyploids often originate through hybridization, which may lead to ecological differences between ancestral and derived lineages (Avise et al., 1992a). Hence, alternative explanations for the dominance of elevated ploidy in high latitudes include: (1) asexual polyploids simply colonize northern habitats more efficiently than sexual diploids, as discussed by Kearney (2005) and Thompson and Whitton (2006), or (2) polyploids are ecologically different from their ancestral lineages and are therefore found in the high-latitude habitats.

One important confounding factor of asexuality and geographical distribution is the frequent association of apomixis and polyploidy (Vandel 1928; Suomalainen et al., 1987). All apomictic angiosperms are polyploid (Asker and Jerling, 1992) but not all polyploid plants are asexual. In the animal kingdom, the association between polyploidy and parthenogenesis is present in two-thirds of taxa, mostly insects and reptiles (Suomalainen et al., 1987; Otto and Whitton, 2000). There is an almost perfect match between the geographic distribution of reproductive modes and the geographic distribution of ploidy levels (Bierzychudek, 1987; Suomalainen et al., 1987; Jokela et al., 2003; Stenberg et al., 2003). Previous attempts to contrast the importance of asexual reproduction and elevated ploidy as explanations for geographic parthenogenesis have been complicated by this almost perfect match (Suomalainen et al., 1987; Jokela et al., 2003; Stenberg et al., 2003). Whereas most polyploid animals are parthenogenetic, only a small fraction of the polyploid plants are apomictic. In fact, more than 99% of the polyploid plants are sexual. In all unisexual vertebrates (some 90 species) (Vrijenhoek et al., 1989; Vrijenhoek, 1989) that have been carefully examined parthenogenesis arose in association with hybridization (Dawley, 1989; Avise et al., 1992a; Avise, 2008). According to Arnold (2007), hybridization underlies the origin of many parthenogenetic fish taxa. “Geographical polyploidy” has been coined by zoologists to emphasize the role played by polyploidy in geographical parthenogenesis (Little et al. 1997; Stenberg et al., 2003). It has been argued that elevated polyploidy alone may be a sufficient explanation for geographical parthenogenesis (Vandel, 1940; Lundmark and Saura, 2006) even though asexuality and polyploidy need not be directly related to each other (Lundmark and Saura, 2006). As an example of the latter, the wider geographic distribution of asexual triploids compared to asexual diploids of Eucypris virens ostracods is due to elevated ploidy rather than to asexuality (Adolffson et al., 2010).

Endothermic animals such as mammals and birds are particularly sensitive to polyploidy (Lampert and Schartl, 2008). Even partial aberrations from the regular diploid cell stage usually result in severe developmental defects (FitzPatrick et al., 2002) and are an important factor in cancer formation (Krämer and Ho, 2001). Fishes, amphibians and reptiles on the other hand cope very well with polyploidy (Pandian and Koteeswaran, 1998). In this group polyploidy is considered an important driving force in evolution as it increases the genetic material on which mutation and selection can act (Ohno, 1970). Transitions to higher ploidy often require special cytogenetic processes that circumvent the problems occurring during normal meiosis when uneven number of chromosomes attempt to pair. It is therefore not surprising that the great majority of uneven ploidy groups lack meiosis and reproduce asexually (Otto and Whitton, 2000). In addition to merely having more DNA, advantages of polyploidy may result from frequent genome rearrangements (Soltis et al., 2004) and alteration of gene expression. In microorganisms, more DNA buffers against the adverse sequelae of mutagenesis and enables a bet-hedging evolutionary lifestyle. In adverse ecological habitats that constrains sexual reproduction, multicellular organisms may resort to DNA buffering by polyploidy. More DNA, however, has its resource costs (Neiman et al., 2013), particularly with regard to the limited resource phosphorus (Heininger, 2012). Asexual Potamopyrgus antipodarum, a New Zealand snail, have markedly higher bodily phosphorus and nucleic acid content per unit mass than sexual counterparts. These differences coincide with and are almost certainly linked to the higher ploidy of the asexuals (Neiman et al., 2009b; 2013).

15.1.3 Automixis and mitotic recombination

A whole array of processes intermediate between mitosis and meiosis are known. Some of these are grouped under the term automixis. Some automictic processes lead to homozygosity in offspring, others retain heterozygosity (Asher, 1970; Birky, 1996). Most thelytokous hymenopterans reproduce by some form of automixis. Early stages of meiosis to form a haploid ovum are normal. Diploidy is restored by fusion of two haploid nuclei from a single dividing oogonium. The restoration of diploidy may occur in different ways, each with different consequences for genetic variation among offspring (Lamb and Wiley, 1987; Suomalainen et al., 1987; Haccou and Schneider, 2004; Beukeboom and Zwaan, 2005; Mateo Leach et al., 2009). Importantly, the strength of Muller’s ratchet is reduced considerably for several forms of automictic thelytoky (Haccou and Schneider, 2004).

Parasex can produce genetic diversity via independent chromosomal assortment, mitotic recombination, and the ability of the diploid state to act as a capacitor for evolution by enabling the accumulation of recessive mutations that are deleterious individually but beneficial in combination (so-called reciprocal sign epistasis) (Pontecorvo, 1956; Schoustra et al., 2007; Calo et al., 2013). Mitotic recombination significantly hastens the spread of beneficial mutations in asexual populations. Indeed, given empirical data on mitotic recombination, adaptation in asexual populations proceeds as fast as that in sexual populations, especially when beneficial alleles are partially recessive (Mandegar and Otto, 2007). Candida albicans, the most prevalent pathogen of humans, has no known meiotic cycle but undergoes mitotic recombination during somatic growth. C. albicans undergoes a parasexual cycle wherein two diploid cells mate, resulting in cell fusion and a ploidy increase (2N to 4N), and the tetraploid cells then undergo mitosis and random chromosome loss to return to the diploid state with no recognized meiosis (Noble and Johnson, 2007; Forche et al., 2008; 2011). C. albicans generates increased amounts and different types of genetic diversity in response to a range of stress conditions that arises either by elevating rates of recombination and/or by increasing rates of chromosome missegregation (Forche et al., 2011). Likewise, biotic and abiotic stress increased plant somatic homologous recombination (Filkowski et al., 2004; Molinier et al., 2006; Boyko et al., 2010). Schoustra et al. (2007) examined the growth fitnesses of both haploid and diploid Aspergillus nidulans strains, and they found that diploid strains attained higher fitnesses than isogenic haploid strains (i.e., the diploid strains’ progenitors) after ~3,000 mitotic generations, and invariably, these faster-growing isolates evolved from a diploid progenitor that had undergone a parasexual reduction to return to the haploid state. Thus, mitotic recombination occurring during the parasexual cycle can accelerate adaptation under laboratory conditions (Schoustra et al., 2007). The higher fitness obtained is due to “sign epistasis” effects (see chapter 18.1.1), where mutations occurring in diploid nuclei could be neutral or deleterious on their own in a haploid but are instead advantageous when combined. This study revealed that the parasexual cycle can serve as a capacitor for evolution and might generate genotypic diversity de novo rather than admixing genetic differences from two divergent parental isolates (Lee SC et al., 2010).

Gene conversion and mitotic crossing-over have the potential to both purge the genome of mutations and increase allelic richness of the species by bringing together advantageous mutations (between alleles within individuals or lineages) (Gandolfi et al., 2001) and escape ‘Muller’s ratchet’ (Lehman, 2003). In mutation-accumulation lines of asexual Daphnia it was shown that the rate of loss of nucleotide heterozygosity by ameiotic recombination is substantially greater than the rate of introduction of new variation by mutation (Omilian et al., 2006). But also gene conversion between different genomes within one cell of an oligoploid or polyploid microorganism would be a mechanism to escape ‘Muller’s ratchet’ (Khakhlova and Bock, 2006). In H. volcanii and the polyploid methanogenic archaeon M. maripaludis, gene conversion indeed operates and leads to a fast and efficient equalization of genome copies (Hildenbrand et al., 2011; Lange et al., 2011; Soppa, 2011). Predicting the occurrence of recombination events by analyzing aligned sequences of a given region of DNA that all originate from one species, some evidence for recombination was found particularly in arbuscular mycorrhizal fungi, and less in Darwinula stevensoni and the bdelloid rotifers (Gandolfi et al., 2003), although the method could not resolve the question whether recombination was mitotic or meiotic. Another possible mechanism to escape Muller’s ratchet is horizontal gene transfer in the population, and for several natural populations it has indeed been found that recombination is so frequent that it resembles sexual reproduction (Papke et al., 2004).

Many asexual taxa are thought to be particularly efficient in DNA repair, which would allow them to reduce the accumulation of deleterious mutations (Castonguay and Angers, 2012). There is evidence for this in asexual taxa such as asexual weevils (Tomiuk and Loeschcke, 1992), aphids (Normark, 1999), darwinulid ostracods (Schön et al., 1998), Daphnia (Omilian et al., 2006), and oribatid mites (Schaefer et al., 2006). This led Schaefer et al. (2006) to ask “why not more taxa are ancient asexuals if the long-term disadvantages of parthenogenetic reproduction can be defeated.”

15.1.4 Phenotypic plasticity

In addition to their versatile genome, asexually reproducing organisms may display a high degree of phenotypic plasticity (Stibor, 1992; Bruno and Edmunds, 1997; Scheiner and Yampolsky, 1998; Hollingsworth, 2000; Amsellem et al., 2001; De Waal, 2001; Negovetic and Jokela, 2001; Mitchell and Read, 2005; Stelzer, 2005; Geng et al., 2007; Hoogenboom et al., 2008; Tully and Ferrière, 2008; Svanbäck et al., 2009; Gorelick et al., 2011; Castonguay and Angers, 2012; Massicotte and Angers, 2012). Bet-hedging, indicating phenotypic plasticity in varying environments, has been observed in various asexual taxa (Ricci, 1991; Gilbert and Schreiber, 1995; 1998; Gilbert, 1998;  Orsenigo et al., 1998; Altiero et al., 2006; Pinto et al., 2007; Rossi and Menozzi, 2012). Based on epigenetic mechanisms (Verhoeven et al., 2010a; Gorelick et al., 2011; Castonguay and Angers, 2012; Harris et al., 2012; Massicotte and Angers, 2012; Robichaud et al., 2012) this phenotypic plasticity is triggered particularly by environmental stress (Finnegan, 2002; Boyko and Kovalchuk, 2011; Grativol et al., 2012). Epigenetic variation could be particularly important to the evolutionary potential of asexual species that harbor little genetic variation (e.g. Wilson et al., 2003; Richards et al., 2008; Vogt et al., 2008; Verhoeven et al., 2010a).

15.1.5 Seed/egg banks and dormancy as bet-hedging strategies

Sexual reproduction relies onan“educated guess” bet-hedging strategy both within and between generations. Dormancy is a bet-hedging strategy used by a wide range of taxa, including microorganisms. It refers to an organism’s ability to enter a reversible state of low metabolic activity when faced with unfavorable environmental conditions (Lennon and Jones, 2011). Soil and sediment banks of dormant propagules are a means to conserve genetic variation between generations and rely on temporal bet-hedging (Ellner, 1985; Gómez and Carvalho, 2000; Evans and Dennehy, 2005; Simons and Johnston, 2006; Evans et al., 2007; Gremer et al., 2012). These resting stages accumulate in the sediments of their habitats forming resting propagule banks (Hairston, 1996; Thompson K et al., 1997; Cáceres and Hairston, 1998; Gómez and Carvalho, 2000; De Meester et al., 2002; Brendonck and De Meester, 2003; Evans and Dennehy, 2005; Simons and Johnston, 2006; Evans et al., 2007; Gremer et al., 2012). Plants, rotifers, daphnids, copepods and Artemia are well known examples where dormancy is developmentally programmed in the form of seeds, resting eggs, ephippia or cysts (Gómez and Carvalho, 2000; Brock et al., 2003; Gyllström and Hansson, 2004; Denekamp et al., 2010; Clark MS et al., 2012).So far, at least to my knowledge, none of the treatises on the relationship between sexual and asexual reproduction appreciated the role of seed/egg banking. But I think that it is no freak of nature that organisms that set seed/egg banks are also particularly prone to indulge in asexual reproduction.

A seed/egg bank has the effect of overlapping the generations of the population, integrating the effects of selection and genetic variation over long periods of time (Templeton and Levin, 1979; Eriksson, 1996; Cáceres, 1997; Ehrlén and Lehtilä, 2002; Nunney, 2002; Brendonck and De Meester, 2003; Gyllström and Hansson, 2004; Honnay et al., 2008). When a population goes through a genetic bottleneck, recruitment of below-ground/sediment genotypes conserved in the soil can quickly restore above-ground genetic diversity as soon as the habitat conditions become suitable again (McCue and Holtsford 1998, Brendonck and De Meester, 2003; Uesugi et al. 2007, Honnay et al., 2008). The costs and benefits of prolonged dormancy are context dependent. Forgoing one or more seasons of growth and reproduction in favor of remaining dormant can confer fitness advantages in a variable environment (Gremer et al., 2012).

The below-ground genetic reservoir is not limited to propagules. Many widespread ecosystems of temperate, arid and polar regions such as grassland, steppe, desert and tundra have >50% of plant production or biomass below-ground (Jackson et al., 1997; Steinaker and Wilson, 2005; Mokany et al., 2006; Pärtel et al., 2012; Poorter et al., 2012). Below-ground richness generally exceeds that above-ground, because of greater dispersion of plant parts below- than above-ground in both time and space. Most perennial plants have persistent below-ground storage organs and meristems that can survive during unfavorable seasons and years in the absence of above-ground biomass (Eissenstat and Yanai, 1997; Wells and Eissenstat, 2001) and allow short- or long-term dormancy for up to decades in the absence of above-ground biomass (Klimesova and Klimes, 2007; Reintal et al., 2010).

15.2 Asexual reproduction in sessile organisms

So far, at least to my knowledge, scientific treatises on the evolutionary rationale of sexual reproduction took no account of the fundamentally different mode of inheritance between modular and unitary organisms (see Heininger, 2012). Their different lifestyles, sessile vs. mobile, shaped their competitive selection pressures resulting in different germline segregation strategies and bauplans (Heininger, 2012). Sessile, modular organisms, plants and benthic aquatic animals, violate Weismann’s doctrine (Weismann, 1892), lacking germ-soma sequestration (Jerling 1985; Sutherland and Watkinson, 1986; Buss, 1983; 1987; Hughes, 1989). Although genetic variants commonly arise within somatic genomes, organelles and cells, such a variant will not be heritable if it is (i) denied access to the formation of gametes or (ii) denied the capacity to asexually form an independent new organism capable of further propagation, or denied both (Buss, 1983).In modular organisms,the adult body is itself a reproductive unit thanks to the presence of totipotent somatic cells (e.g. meristems, interstitial cells) that can form both gametes and somatic tissues and allow for variants arising in clonal growth to contribute to heritable evolutionary change (Monro and Poore, 2004; 2009). Thus, the ease of switching between sexual and asexual reproduction and the genetic and epigenetic consequences of this switch should be fundamentally different for modular and unitary organisms.

Asexual reproduction is common in plants. Harper (1977) estimated that over 80% of perennial plants have some form of clonal reproduction. A later survey of plant species from Central Europe estimated that close to 67% of the species considered could potentially reproduce clonally (Klimes et al., 1997). Clonal reproduction can be considered as an alternative life cycle loop that allows persistence of a species in the absence of the ability to complete the normal life cycle (i.e. seed production, germination and recruitment). Clonal reproduction appears to offer a safe escape route for many plant species under suboptimal environmental conditions (Kudoh et al., 1999; Erikson and Ehrlén, 2001; Honnay and Bossuyt, 2005). Population dynamics of perennial forest herbs, for example, are strongly influenced by successional stage through the degree of canopy closure: low light conditions can suppress sexual recruitment and trigger clonal growth (Verburg and Grava, 1998; Kudoh et al., 1999; Lezberg et al.. 2001). Apomixis in plants may take two general forms: the seeds or spores may be derived from altered developmental sequences that bypass normal meiotic divisions and omit syngamy, or apomixis may be based upon vegetative reproduction (Klekowski, 2003). Stolons, runners, rhizomes, tubers, bulbils, corms, layering, fragmentation, and apomixis are among the myriad alternative modes of asexual reproduction exhibited among angiosperms (Richards, 1997). Morphological plasticity on an ecological time scale is common in forms with somatic embryogenesis. Three general types of shoot apical meristems occur in vascular plants: single tetrahedral apical initial in pteridophytes, unstratified with impermanent initials in gymnosperms and stratified with impermanent initials in angiosperms and some gymnosperms (Klekowski, 2003). Many plants are constructed by iteration at both modular and clonal scales, which produces a hierarchical organization. Each level of the hierarchy is a level at which replication or copying occurs, at which births and deaths can be counted. The units at each level can carry mutations and transmit them to future generations (Clarke, 2012). As a plant ages, the number of mutations per apical initial increases (Klekowski, 1988; Klekowski and Godfrey, 1989). Thus, long-lived plants accumulate somatic mutations and become genetic mosaics as they grow. Moreover, under abiotic or biotic stress increased mutagenic changes in plant genomes have been identified (McClintock, 1984; Cullis, 1987; 2005; Ries et al., 2000; Lucht et al., 2002; Kovalchuk et al., 2003; Molinier et al., 2006). Stress-related epigenetic variation may also increase phenotypic plasticity (Verhoeven et al., 2010a; b). Environmental stresses can trigger DNA methylation changes in plants (Chinnusamy and Zhu, 2009). Stress-induced methylation changes may be targeted specifically to stress-related genes. Alternatively, methylation changes may generate nonspecific (random) differences between individuals, which may have adaptive significance during times of stress (Rapp and Wendel, 2005), because they increase the range of variation that natural selection can act upon. There are indications that apomictic dandelions may have compensatory mechanisms to generate heritable variation, for instance via increased transposon activity or somatic recombination (Richards, 1989; King and Schaal, 1990).

Intraorganismal selection has been acknowledged by numerous authors (Whitham and Slobodchikoff, 1981; Buss, 1983; Antolin and Strobeck, 1985; Klekowski, 1988; 2003; Sutherland and Watkinson, 1986; Gill et al., 1995; Otto and Orive, 1995; Clarke, 2011) and may be a way both of eliminating deleterious somatic mutations (Klekowski and Kazarinova-Fukshansky, 1984; Michod, 1995; Otto and Orive, 1995; Otto and Hastings, 1998) and to fix advantageous mutations (Otto and Orive, 1995; Fagerström et al., 1998; Otto and Hastings, 1998; Pineda-Krch and Fagerström, 1999; Pineda-Krch and Lehtilä, 2004; Clarke, 2011). This variability may be transmitted to the offspring both sexually and asexually (Whitham and Slobodchikoff, 1981). Genetic variance for effects caused by somatic mutations will be greater among spores than among offspring produced by budding or fission, because the variance of items is greater than the variance of means, and spore production therefore enhances the effect of selection in reducing mutational load (Bell and Koufopanou, 1991). In sponges, species capable of somatic embryogenesis uniformly lacked a repeated colony morphology, whereas species incapable of somatic embryogenesis always form characteristic colony morphologies (Korotkova, 1970; Buss, 1983). Angiosperms, the phylogenetically newest and most successful of all plants, have evolved stratified meristem regions, which make intraorganismal selection maximally effective in three ways: They ensure that deleterious mutations are rapidly purged, they allow high fitness mutations to displace the wild type, and they preserve genetic variance over the long term (Pineda-Krch and Lehtilä, 2002; 2004; Clarke, 2011).Somatic mutation and selection can cause ramets to become genetically distinct from each other, even if they are mitotic descendants from a common zygote.Thus, asexual plant populations can diversify more rapidly than sexual populations because they are free from the homogenizing effects of sexual recombination and segregation (Johnson MT et al., 2011). Although asexual reproduction may often constrain adaptive evolution (Hill and Robertson, 1966; Barton and Otto, 2005; Johnson et al., 2009; Hersch-Green et al., 2012), the loss of recombination and segregation need not be an evolutionary dead end in terms of diversification of lineages (Johnson MT et al., 2011). The parthenogens of Chara canescens (Charophyceae), multicellular green algae that superficially resemble land plants, occupy broader geographical and ecological ranges than their sexual counterparts. An experimental study showed that parthenogenetic C. canescens individuals from two neighbouring populations are locally adapted to light and differed in their capacity to acclimate to irradiance and salinity suggesting clonal diversity with differentially adapted clonal ‘‘microspecies’’ (Schaible et al., 2012).Artificial selection of intraclonal genetic variation in the branching red seaweed Asparagopsis armata demonstrated that intraclonal genetic variation may potentially help clonal organisms to evolve adaptively in the absence of sex and thereby prove surprisingly resilient to environmental change (Monro and Poore, 2009).

Modular organisms have an arsenal of strategies with which they can create non-sexual genotypic/phenotypic plasticity: somatic mutagenesis, polyploidy, seed banks, and epigenetic variation. As result, plants and sessile aquatic animals often exhibit labile asex/sex expression and rather exhibit a near-continuum between sexuality and asexuality (Barrett, 2002; Goodwillie et al., 2005; Whitton et al., 2008) but may turn to sexual reproduction under environmental challenges (Bell and Wolfe, 1985; Harvell and Grosberg, 1988; Kimmerer, 1991; Romme et al., 1997; Korpelainen, 1998; van Kleunen et al., 2001). Most plants combine sexual and clonal reproduction, and the balance between the two may vary widely between and within species (Eckert, 2002; Clarke, 2011). For plants both somatic and germ-line mutations can be passed to the gametes or vegetatively produced organisms. Asexual reproduction may have the advantage that such selected genotypes, unlike in sexuals, are not broken up each generation by recombination. Obviously, stable, undisturbed environments should favor such strategies (Silvertown, 2008).An analysis of more than 2,000 plant populations suggested that the ultimate clonal plant would be a rare, aquatic, alien apomict living in an undisturbed (very long timescales free from disturbance are needed), geographically marginal habitat (Silvertown, 2008). This was considered such a restrictive set of ecological conditions that it is perhaps better regarded as a recipe for the failure of sexual reproduction than as clonal success (Silvertown, 2008). On the other hand, the geographic distribution of sexual and asexual plant species suggests that more harsh environmental conditions favor asexuals (Peck et al., 1998, Eckert et al., 1999; Dorken and Eckert, 2001; Eckert, 2002).Consistent with these observations, environmental stress, particularly cold and water stress, induces male sterility in a variety of plants (Sawhney and Shukla, 1994; Sheoran and Saini, 1996; Saini, 1997; Knight et al., 2005; Sakata and Higashitani, 2008; Zinn et al., 2010). Moreover, near the limits of species’ ranges, factors that severely limit sexual recruitment (e.g., lack of pollinators, short growing season) may tip the balance that would otherwise favor retention of sexuality, facilitating vestigialization of sexual traits in asexual plants(Eckert, 2002; Dorken et al., 2004). Asexual organisms are thought to have fewer disadvantages than sexuals in environments with few biotic interactions, such as in high altitudes, high latitudes and arid habitats (Bell, 1982; Hörandl, 2006).

In the Mojave desert, along a stress gradient of light intensity and moisture availability from understory to intershrub microsites, males of the dioecious desert moss Syntrichia caninervis are clustered near the less stressful shrub canopy line and are less stress tolerant than females. Sex expression declined from 76% under the shrub canopy to 5% in the intershrub region (Stark LR et al., 2005). Sexual reproduction in an extremophile bryophyte system, as measured by the number of sporophytes per shoot, decreases with extreme environmental stress at geothermal hot-springs. The number of sporophytes per shoot was found positively correlated with distance from geothermal features. Sporophytes were most common in non-geothermal sites, but very rare in high temperature geothermal sites suggesting that sperm viability may be a significant factor limiting sporophyte formation in geothermal bryophyte communities (Eppley et al., 2011). The reproduction of mosses Pleurozium schreberi and Pohlia nutans were compared in the surroundings of copper smelters at Harjavalta, Finland (Huttunen, 2003). The production of gametangia decreased in P. schreberi, and near the smelters most of the shoots were sterile. The numbers of spores per capsule in P. nutans decreased at polluted sites and the proportion of aborted spores increased (Huttunen, 2003). Harsh environmental conditions are associated with strong selection for the target trait but low genetic variance of this target trait. Consequently, the potential for microevolution is constrained by either a lack of heritable variation (in poor environments) or by a reduced strength of selection (in good environments) (Wilson AJ et al., 2006).

15.3 Asexual reproduction in mobile animals

15.3.1 The “evolutionary scandals”: microscopic organisms inr-selected habitats

In chapter 4.2 it has been argued that it is a unicellular bet-hedging strategy to create mutants and let natural selection decide on their viability. In large multicellular organisms this would mean a large investment into possibly poorly viable mutants, a strategy that in resource-limited environments would not pay off. Obviously, there are two liabilities, resource limitation and stress, that balance the switch from sexual to asexual reproduction and back. Resource limitation of reproduction is determined by resource availability and investment of resources into offspring. In unicellular microorganisms the investment in new organisms is small and hence genetic exploration of fitness landscapes by trial-and-error expeditionsat the colony population level is a viable strategy. In multicellular organisms, the balance between resource availability and resource investment into new organisms determines whether this bet-hedge approach at the population level is evolutionarily stable. Only microscopic multicellular organisms in less resource limited habitats can be expected to comply with these requirements.

There is a long-standing debate over the existence of ancient asexual organisms (Martens et al., 2003; Martens and Schön 2008; Birky, 2010). The term ‘ancient asexual’ implies persistence of an asexual lineage longer than expected under the various hypotheses for sex (Neiman et al., 2009a). Because truly ancient asexuals would violate the expectation that sex and recombination are necessary for long-term survival, they have been considered an “evolutionary scandal” (Maynard Smith, 1986; Judson and Normark, 1996; Schön et al., 2009).Various molecular- and organismal-based approaches were used for recognizing signs of (cryptic) sex (Ramesh et al., 2005; Malik et al., 2008; Schurko et al., 2009). Some scientists do not accept the existence of any long-term asexuality at all (Little and Hebert, 1996), or only accept this evidence for the bdelloid rotifers (Hayden, 2008). Without taking sides in this controversy, I delineate the framework within which asexual reproduction may have endured for millions of years.

Given the differences regarding speciosity, individual genetic variability, genetic mechanisms underlying the formation of gametes (apomixis versus automixis), incidence of rare males and prevalence of homogenizing mechanisms between the few putative ancient asexual groups (bdelloid rotifers, darwinulid ostracods, certain lineages within oribatid mites) it is clear that there is no single scenario that will explain the origin and persistence of all three groups (Schön et al., 2009). Yet, some unifying ecological principles can be outlined. At the phenotypic level, relatively old asexuals appear to exhibit a relative preponderance of taxa with traits that may reduce extinction risk, including some combination of high dispersal abilities, dormant resting stages, broad geographical distributions, and large population sizes (Schwander and Crespi, 2009). Most darwinulids and bdelloids live in what could generally be termed “marginal habitats”. One possibility is that competition (with sexual species) would be lower, as there are simply fewer species occurring in such highly fluctuating habitats. Another reason could be that asexuals are able to survive in very low densities over many generations, since they do not need to find a mate in order to reproduce (Van Dijk, 2007; Hörandl, 2008). For all sexual populations, there is a density threshold (mediated by size and mobility of animals, amongst other biological characteristics) below which the probability of finding a mate is too low to ensure sufficient reproduction for the population to remain viable. In marginal habitats, such as semi-terrestrial ones, conditions may vary widely and asexuals would have the advantage over sexuals. If asexuals are sufficiently small so that effective population sizes still remain large in spite of low densities (and all three putative asexual groups have very small body sizes), such low densities may not necessarily lead to genetic bottlenecks as sufficient genetic variability might survive (Schön et al., 2009).

15.3.1.1 Bdelloid rotifers

Bdelloid rotifers are abundant invertebrate animals that can be found on every continent in almost any freshwater environment and moist-terrestrial habitats over a wide range of temperature and pH, and are often among the most common microinvertebrates, particularly in ephemeral (transitory) aquatic environments (Ricci, 1987; Wallace and Snell, 2001). Individuals range from about 0.1 to 1 mm in length and have muscles; ganglia; tactile and photosensitive sensory organs; structures for feeding, swimming, and crawling; digestive and secretory organs; and ovaries. Their ability to maintain cosmopolitan metapopulations is related to their long-distance dispersal abilities, their ability to initiate populations with single females, and their capacity to survive unfavorable periods via the presence of dormant resting stages (Finlay and Fenchel, 2004; Fontaneto et al., 2008). Although individual biomass is minute, large population size, coupled with high turn-over rates make rotifers an important component of food webs (Herzig, 1987; Wallace and Snell, 2001). Adding to their importance is the fact that rotifers are eaten by invertebrate predators and are also the first food of fish fry, thereby making their energy available to higher trophic levels. Bdelloid rotifers constitute the largest, oldest, most diverse animal taxon for which there is morphological, cytological, and molecular evidence for long-term parthenogenesis. Their closest relatives are in the class Monogononta, which reproduce mostly by apomictic parthenogenesis, with an occasional one-generation sexual cycle. The bdelloids are believed to descend from a parthenogenetic female monogonont that lost the ability to enter a sexual cycle (Birky, 2004). The putative role of resource availability in this change of reproductive strategy has been shown experimentally (Boraas, 1983; Bennet and Boraas, 1989; Fussmann et al., 2003, see chapter 3). Fossil evidence shows that the bdelloids are at least 35 to 40 million years old while their genetic diversity suggests they are more than twice that age. In this period, the descendants of the first bdelloid diversified into at least 360 species in three families (Poinar and Ricci, 1982; Waggoner and Poinar, 1993; Mark Welch and Meselson, 2000; 2003; Wallace and Snell, 2001; Birky, 2004; 2010; Mark Welch et al., 2004a; 2004b; 2009; Rice and Friberg, 2007; Neiman et al., 2009a). Two bdelloid features, the ability to survive desiccation at any life stage and obligate parthenogenesis allow an unusual lifestyle (Ricci, 1987; 1998). These features qualify the bdelloids as colonizing organisms in r-selected environments. Anhydrobiosis capacitates bdelloids to stably inhabit desiccation-prone habitats and also provides a means of dispersal: when desiccated, a bdelloid contracts into a small flat ellipsoid called a tun, which adheres firmly to substrates such as soil particles, mud, or moss fragments that may be passively transported over long distances (Tunnacliffe and Lapinski, 2003). During quiescence, bdelloids appear to suspend respiration, metabolism, and aging. Thus, anhydrobiosis is the ultimate strategy for eggs or other stages of the life cycle to survive extended periods of environmental stress (Guppy and Withers, 1999; Radzikowski, 2013). On recovery, bdelloids resume reproduction and other life history traits consistent with their age at the start of dormancy ignoring the time spent in anhydrobiosis (Ricci et al., 1987; Ricci and Covino, 2005). Adults, eggs and embryos of 15 bdelloid species, representing four families and six genera (inhabiting both water bodies and water retained by mosses), were desiccated and kept dry for 7 days. After this treatment, higher recovery rates were observed for the moss species, i.e. those frequently subjected to desiccation events (Ricci, 1998). In a variety of organisms the ability to withstand desiccation confers cross-tolerance to various other extreme environmental stressors, including different types of radiation (Alpert, 2006; Watanabe, 2006; Daly et al., 2007; Watanabe et al., 2007; Gladyshev and Meselson, 2008; Jonsson et al., 2008; Hengherr et al., 2009; Slade et al., 2009; Gusev et al., 2010). During anhydrobiosis, DNA breaks take place (Gladyshev et al., 2008; Gladyshev and Arkhipova, 2010; Gusev et al., 2010) that are repaired by homologous recombination (Rice and Friberg, 2007; Slade et al., 2009; Gusev et al., 2010). Bdelloids may be unusually susceptible to DSBs during anhydrobiosis because, unlike most other organisms known to be capable of anhydrobiosis, they do not produce the disaccharide trehalose, which promotes tolerance to desiccation (Lapinski and Tunnacliffe, 2003; Rice and Friberg, 2007). Bdelloid populations maintained in a hydrated state over many generations show a decline in fitness compared to populations that were cyclically desiccated (Ricci et al., 2007). Repair processes associated with recovery from desiccation may have a beneficial effect beyond desiccation tolerance. Repair of desiccation-induced DNA damage would require the presence of a homologous template, maintaining colinear pairs in gene-rich regions and selecting against insertion of repetitive DNA that might cause chromosomal rearrangements (Gladyshev and Arkhipova, 2010). Mothers who have been through desiccation, produce daughters of increased fitness and longevity (Ricci and Covino, 2005) and desiccation stress is a necessary condition for enhanced fitness in bdelloids (Ricci et al., 2007). This feature is reminiscent of the germline rejuvenating action of mitochondrial oxidative stress in sexually reproducing organisms (Isaeva and Reunov, 2001; van Werven and Amon, 2011). Parthenogenesis allows bdelloids to rapidly re-occupy a habitat after it changed from unfavorable to favorable conditions or single tuns to found a new population and colonize new habitats (Birky et al., 2005).

The rate of molecular evolution in bdelloid rotifers is higher than that of close sexual relatives and points toward mutation accumulation in the absence of sex (Barraclough et al., 2007).However, such accumulation of mutations can be handled without deleterious effects over long time-frames (Mark Welch and Meselson, 2000; 2001). In bdelloid rotifers many genes were found that appear to have originated in bacteria, fungi, and plants, concentrated in telomeric regions along with diverse mobile genetic elements. The capture and functional assimilation of exogenous genes resembles a bacterial transformation-like type of “parasexual” behavior and may represent an important force in bdelloid evolution (Gladyshev et al., 2008).

The rotifer Habrotrocha elusa is able to survive long periods of desiccation which, combined with aerial dispersal, allows this invertebrate to escape a deadly fungal parasite, Rotiferophthora angustispora, an organism less tolerant to desiccation (Wilson and Sherman, 2010; Leung et al., 2012). If anhydrobiotic dispersal enables bdelloid species to escape temporally and spatially from some or many natural enemies, their coevolutionary burden would be substantially reduced. Therefore, bdelloids may have evaded parasites and pathogens over evolutionary time, without incurring the costs of sexuality, by playing a never-ending game of “hide-and-seek” (Ladle et al., 1993; Judson, 1997; Wilson and Sherman, 2010; Wilson, 2011).

It has been suggested that the absence of meiosis in asexual lineages should lead to higher interallelic divergence at any given locus within an individual (i.e., allelic sequence divergence) compared to sexual populations (i.e., Meselson effect; Mark Welch and Meselson, 2000). However, most studies in putative asexual lineages failed to show high levels of neutral allelic divergence (Kuhn et al., 2001; Schön and Martens, 2003; Pawlowska and Taylor, 2004; Hijri and Sanders, 2005; Schaefer et al., 2006), with the exception of the bdelloid rotifers (Mark Welch and Meselson, 2000; Pouchkina-Stantcheva et al., 2007), Meloidogyne root knot nematodes (Lunt, 2008; Danchin et al., 2011b),and Timema stick insects (Schwander et al., 2011), while they revealed sex in the Placozoa (Signorovitch et al., 2005). However, as was pointed out (Butlin, 2000; Schön et al., 2008), the Meselson effect is asymmetric: its presence may confirm long-term asexuality, but its absence does not necessarily refute itsince ‘unsexy’ processes, such as DNA repair, mitotic recombination, and allelic gene conversion may lead to homogenization between alleles and to the loss of mutations that arose anew on different chromosomes.But the story is even more complicated: part of the high allelic divergence seen in bdelloids probably derived from an initially diploid lineage that underwent whole-genome duplication followed by massive gene loss and karyotype restructuring(Maderspacher, 2008; Mark Welch et al., 2008; Hur et al., 2009). On the other hand, the level of divergence between ancient alleles was markedly lower (ca. 3%) and not very different from the range observed in sexually reproducing species such as Ciona savingnyi (Small et al., 2007). Subsequent observations in Adineta vaga (another bdelloid species) confirmed this result (Hur et al., 2009).

In Meloidogyne root knot nematodes, the observed Meselson effect is the result of previous hybridization between species (Lunt, 2008). Thus, the sequence divergence between duplicated chromosome pairs (i.e., paralogous loci) could be misinterpreted as high allelic divergence and the Meselson effect in these lineages (Mark Welch et al., 2008; Schurko et al., 2009).

Taken together,their minute individual biomass (minimizing investment into new individuals), large population size, colonizing lifestyle in r-selected habitats, egg bank, anhydrobiosis and associated extreme stress resistance, DNA repair capacity and functional assimilation of exogenous genes by a type of “parasexual” behaviorappear to make asexuality an evolutionarily stable strategy in bdelloid rotifers.

15.3.1.2 Darwinulid ostracods

Ostracods are small, bivalved crustaceans, which occur in almost all aquatic and moist (semi-) terrestrial habitats. Like rotifers, ostracods reachlarge population sizes (Van Doninck et al., 2003b) and are an important component of food webs (Evans and Stewart, 1977; Pihl, 1985). The fossil record indicates that the ostracod family Darwinulidae (adult size 0.6–0.8 mm; Vestalenula mathilda <0.4 mm) has reproduced fully asexually for 200 Myr (Martens et al., 2003) and neither sexual nor mixed extant populations or close sexual relatives are known (Martens, 1998).Rare males in a single darwinulid living species, Vestalenula cornelia, have recently been described (Smith RJ et al., 2006), but they may be non-functional, as they seem to have rudimentary reproductive organs; in addition, no spermatozoa have been observed in either males or sympatric females.

Most darwinulids can be found in what could generally be termed “marginal habitats”, such as moist mosses and other (semi-) terrestrial environments, much like bdelloid rotifers.The exceptions to a distribution in marginal habitats are the darwinulid species with General Purpose Genotype (GPG), such as Darwinula stevensoni and Penthesilenula brasiliensis: they are ubiquitous, or nearly so, and can also be found in lakes, rivers, interstitially, etc. (Schön et al., 2009). All darwinulids are brooders, and 4 out of 5 genera have the posterior part of the carapace inflated so as to create a brood pouch (Horne et al., 1998a; Schön et al., 2009). With an average of 6–8 offspring per female in Penthesilenula brasiliensis (Pinto et al., 2007), and 11–15 offspring in Darwinula stevensoni (Van Doninck et al., 2003a), darwinulids generally have low fecundity as compared to other ostracods (Geiger, 1998) and, especially in higher latitudes, rather long life cycles. D. stevensoni takes 4 years in Canada (McGregor, 1969) and in Finland (Ranta, 1979) to complete its life cycle, although further South, e.g. Belgium, it takes only 1 year (Van Doninck et al., 2003a), which is still considerably longer than in most other Cypridoidea (e.g. Martins et al., 2008; Schön et al., 2009). Darwinulids do not have resting eggs like all of the Cypridoidea and some Cytheroidea (Geiger, 1998), and they do not swim but only crawl slowly. The low fecundity and long life cycles of Darwinulids compared to sexual ostracods argue against the twofold advantage of asexuality vs. sexuality (Maynard Smith, 1971a) and beg for other explanations for their long-term asexuality. Horne and Martens (1999) argued that the absence of sexual freshwater ostracods in northern Europe is not just a consequence of the superior colonization abilities of clones. Fossil evidence indicates that sexual ostracods also inhabited northern Europe during post-glacial times and were replaced as climates gradually became more stable. Consequently, they argue that modern climatic stability favored the replacement of sexual lineages by competitively superior clones.

Use of allozyme markers and direct sequencing of nuclear genomes revealed low genetic diversity between D. stevensoni populations that cannot be explained by recent selective sweeps (Rossi et al., 1998; 2004; Schön et al., 1998; 2000, 2003; 2009; Gandolfi et al., 2001). Differences in genetic diversity between embryos and adults furthermore indicate that up to half of the observed genetic changes in adults can be caused by somatic mutations (Schön and Martens, 2003). It has been argued that the presence of homogenizing mechanisms such as gene conversion, DNA repair and mitotic recombination are most likely the most important factors that may reduce mutation rates in darwinulids (Schön and Martens, 1998; Schön et al., 1998; 2008; 2009).Possibly due to these mechanisms the Meselson effect is not observed in Darwinulid ostracods (Schön and Martens, 2003).

The unusually wide tolerance range for both salinity (0-30 g/l) and temperature (10°C, 20°C and 30°C) of the freshwater species D. stevensoni, supports the hypothesis that it has indeed developed a GPG (Van Doninck et al., 2002) (see chapter 15.3.3). It has been argued that a GPG can only originate and persist in fully asexual lineages as recombination will almost certainly break-up the allele-combinations required for a GPG. On the other hand, the chance that a sexual lineage would have a GPG at the time clones originate from this sexual root (and can thus freeze the GPG in the clonal lineage) is very small indeed (Van Doninck et al., 2003b), which might explain why few taxa seem to have evolved a real GPG (Schön et al., 2009).As a further means to conserve the GPG, darwinulids may be able to exert embryo selection. It was shown that Darwinula stevensoni and Penthesilenula brasiliensis have the ability to detect, select and eject unwanted material such as eggs or embryos from the brood pouch (Horne et al., 1998b; Pinto et al., 2007). In the experiments of Pinto et al. (2007) the majority of the ejected eggs remained viable and hatched and the juveniles moulted to later instars. They survived for several months, but none ever reached adulthood.

The freshwater ostracod Eucypris virens, is commonly found in European temporary pools, where its long-term persistence completely relies on the build-up of resting egg banks. Resting eggs of both sexual and parthenogenetic E. virens showed extreme tolerance to a variety of environmental stressors (Vandekerkhove et al., 2013) making the egg banks of dormant propagules a source of genetic variation between generations and for temporal bet-hedging. Patterns of allozyme diversity among unisexual and sexual ostracods Cypricercus and Eucypris virens suggest that the success of unisexual ostracods may also be linked to their ability to recruit clonal diversity through interbreeding with close sexual relatives (Turgeon and Hebert, 1995; Schön et al., 2000). Data derived from a survey of ostracod valves preserved in 34 Holocene freshwater sediment cores suggest that environmental fluctuations during a period of less than 5000 years were sufficient to provide an advantage to sexually reproducing ostracods over parthenogenetic forms (Griffiths and Butlin, 1995).

In summary, their minute individual biomass (minimizing investment into new individuals), large population size, marginal habitats, maintenance of a possible General Purpose Genotype due to embryo selection and DNA homogenizing mechanismsappear to make asexuality an evolutionarily stable strategy in darwinulid ostracods.

15.3.1.3 Oribatid mites

About 10% of all known oribatid mite species reproduce by female parthenogenesis, a rate much higher than in other invertebrate and vertebrate taxa (Cianciolo and Norton, 2006; Avise, 2008).One reason for this high proportion may be the permanent availability of homogeneously distributed dead organic matter in soil systems (Scheu and Drossel, 2007) that allows densities of up to 400,000 individuals per square meter in forest soil (Schatz and Behan-Pelletier, 2008). In acidic boreal forests, they reach densities of up to 400,000 ind/m2 whereas in calcareous forests, densities are usually somewhere between 20,000 and 40,000 ind/m2. There is little seasonal fluctuation of oribatid mite densities, indicating that the communities are in equilibrium conditions (Heethoff et al., 2009). Overall, the lower density of oribatid mites in calcareous forest soils presumably is mainly due to macrofauna activity, whereas their density in more acidic forest soils probably is limited by the availability of high quality resources (bottom-up control; Salamon et al., 2006). In parthenogenetic oribatid mite species males are rare (spanandric) and sterile. Spermatophores are non-functional, as spermatogenesis is incomplete (Taberly, 1988), and there is no evidence for recombination or incorporation of paternal genetic material into the offspring (Palmer and Norton, 1992; Heethoff et al., 2009). Constant conditions favor asexuality (Bürger, 1999) which may explain the high incidence of parthenogenesis in environments such as stable forest soils (Cianciolo and Norton, 2006, Domes et al., 2007). Intriguingly, the distribution of reproductive modes of soil fungi follows a similar pattern (Grishan et al., 2003). On the other hand, oribatid mites that dominate on the bark of trees are predominantly sexual taxa (Cianciolo and Norton, 2006; Erdmann et al., 2006; Fischer et al., 2010). In contrast to oribatid mites in soil, communities on bark appear to be less sensitive to disturbances (Erdmann et al., 2006). Densities on the bark of trees are also low as compared to soil (Erdmann et al., 2006). Bark oribatid mites feed on lichens. Lichens grow slowly and may be in limited supply but also defend themselves, e.g. by producing usnic acids (Emmerich et al., 1993). It has been shown in laboratory experiments that parthenogenetic taxa suffer more than sexual taxa from resource limitations (Domes et al., 2007), suggesting that sexual taxa are better adapted to resource limitation.

Divergence levels in the mitochondrial cytochrome c oxidase subunit 1 gene between and within clades of Platynothrus peltifer suggest that in this species asexual reproduction is at least 100 million years old (Heethoff et al., 2007). Another fully asexual species in this group, Mucronothrus nasalis, is thought to be 200 million years old (Hammer and Wallwork, 1979). There is low genetic divergence and no evidence for the Meselson effect in parthenogenetic oribatid mites even with the probable absence of genetic recombination (Schaefer et al., 2006).

Oribatids are relatively long-lived and have numerous morphological defensive adaptations.Predation (top-down control) is likely to be of little importance as a regulatory factor for adult oribatid mites that have been proposed to live in ‘enemy-free space’ (Peschel et al., 2006). Most also possess a pair of large exocrine oil glands that produce species-specific mixtures of hydrocarbons, terpenes, aromatics, and alkaloids with presumably allomonal functions. When the oil glands of the model oribatid species, Archegozetes longisetosus,are discharged and the ‘disarmed’ individuals are offered as prey to polyphagous Stenus beetles (Staphylinidae), Stenus juno fed on disarmed mites with behavioral sequences and success rates similar to those observed when they prey on springtails, a common prey. In contrast, mites from the control group with full glands were almost completely rejected; contact with the gland region elicited a strong reaction and cleaning behavior in the beetle. Thus, oribatid mite oil gland secretions have an adaptive value for chemical defense (Heethoff et al., 2011). On the other hand, mites suffer a variety of infections (Poinar and Poinar, 1998; van der Geest et al., 2000), so that the coevolutionary value of sexual reproduction may apply.

In summary, their minute individual biomass, stable habitats with abundant resources allowing large population sizes and enemy-free spaceappear to make asexuality an evolutionarily stable strategy in oribatid mites.

15.3.2 Cyclical parthenogenesis

In some organismsphases of asexual propagation are alternated with bouts of sexual reproduction, called cyclical parthenogenesis, holocycly or heterogony.Cyclical parthenogenesis is generally rare. It occurs in only ~15,000 animal species, spread over seven taxonomic groups (Hebert, 1987). From an evolutionary point of view cyclical parthenogenesis is fundamentally sexual (Vrijenhoek, 1998). The cyclically parthenogenetic life-cycle is believed to retain the advantages of recombination while minimizing the cost of sex. However, this life cycle is also thought to be unstable due to periodic loss of sexual reproduction by directional selection. Yet, cyclical parthenogens are of special interest because they allow a direct assessment of the ecological conditions, costs and benefits of sex versus asex.

15.3.2.1 Monogonont rotifers

Monogonont rotifers are small, aquatic invertebrates that normally reproduce by cyclical parthenogenesis, an alternation between an extended phase with ameiotic parthenogenesis and short sporadic sexual episodes (Arndt, 1993; Nogrady et al., 1993; Armengol et al., 2001; Wallace et al., 2006). Explaining the evolutionary dynamics of the monogonont rotifer life cycle is important for understanding how cyclical parthenogenesis is maintained, and for comparing monogononts with their close relatives, the bdelloid rotifers, which are ancient obligate asexuals. Sex is initiated with the production of sexual females, whose oocytes undergo meiosis and develop into haploid males (if not fertilized), or diploid diapausing eggs (if fertilized), which show a remarkable tolerance to unfavorable conditions that typically occur every year (Serra and Snell, 2009)and remain viable for decades (Marcus et al., 1994; Kotani et al., 2001; Clark MS et al., 2012).Diapausing eggs contain multinucleate diapausing embryos encased in a three-layered shell that protects them from external stressors, like desiccation and temperature extremes (Wurdak et al., 1978). Highly instructive are the dormancy patterns among rotifers: short-lived monogononts produce resting eggs with long-lasting diapause commonly after switching to mictic phase; bdelloids, living 3 times as long, enter anhydrobiosis with short-lasting quiescence at any time during their life cycle as a direct response to changing environment. The two dormancy forms of the rotifers are alternative and mutually exclusive and appear to be related to the temporal variation of their habitats (Ricci, 2001). Molecular characterisation of processes associated with the formation of resting eggs revealed that small heat-shock proteins and some antioxidant genes were upregulated in resting eggs, therefore suggesting that desiccation tolerance is a characteristic feature of resting eggs even though they do not necessarily fully desiccate during dormancy (Denekamp et al., 2009).Screening polymorphic microsatellite loci in populations of the rotifer Brachionus plicatilis in a temporary pond, Gómez and Carvalho (2000) analysed: (i) the genetic structure of the resting egg bank; (ii) the changes in the genetic structure of rotifer populations during the parthenogenetic phase; and (iii) the population structure after its initiation from resting eggs. The last sample in the parthenogenetic phase showed evidence of clonal selection, as indicated by a low observed clonal diversity and the appearance of linkage disequilibria. On the other hand, the resting egg bank was in Hardy–Weinberg and linkage equilibrium, and contained a high genotypic diversity. Unexpectedly, the resting egg bank differed from the planktonic population in its allelic composition, suggesting that resting egg hatching is biased (Gómez and Carvalho, 2000).

At the interface between sexual and asexual reproduction, the role of resource availability for the reproductive mode and the offspring quantity/quality trade-off becomes evident (see chapter 3). There are several documented cases of Brachionus strains that have permanently lost the ability to reproduce sexually (Boraas, 1983; Buchner, 1987; Bennett and Boraas, 1988; Fussmann et al., 2003; Stelzer, 2008; 2011; Stelzer et al., 2010). Boraas (1983) found that newly established cultures of Brachionus calycifloru collected from the field produced 40% mictic (sexual) females when induced. After 2–3 months in a chemostat, i.e. with unlimited resources, that percentage was reduced to 0 in similarly inducing environments. This work has been confirmed (Bennet and Boraas, 1989; Fussmann et al., 2003; Becks and Agrawal, 2010), establishing the costs of sex under unlimited resources (Stelzer, 2011).Loss of sex has been found to be stable for years in strains used in different laboratories (Stelzer, 2008). Bennet and Boraas (1989) founded their cultures with a single female, and hence genetic variation causing sex loss had to arise during experimental culture. In other experiments not initiated with one female, genetic variation for investment in sex might be present in the founder populations. Nevertheless, the observed evolutionary dynamics did result in selection for sex loss (Serra and Snell, 2009). Resource abundance has been considered to be the most significant factor determining population size. Under various food levels, Brachionus plicatilis flexibly changes its reproductive patterns and lifespans (Yoshinaga et al., 2000; 2003). In a food-rich environment, B. plicatilis produces approximately 30 offspring during its lifespan of approximately 10 days. In contrast, when fed for only a few hours daily, it suppresses active reproduction and produces less than 10 offspring, while surviving for nearly a month. Long-lived individuals resulting from reproductive suppression are likely to obtain a second chance for reproduction in future. Besides life history parameters, offspring quality (starvation resistance) also increases when B. plicatilis reproduces in a food poor environment(Yoshinaga et al., 2001; 2003).

In Brachionus and several other monogonont rotifers, the production of sexual females is induced at high population densities by a chemical that is produced by the rotifers themselves (Stelzer and Snell, 2003; Snell et al., 2006; Timmermeyer and Stelzer, 2006), a process analogous to quorum sensing in bacteria (Kubanek and Snell, 2008). Obligate parthenogens are unable to produce sexual females, thus they also lack males and diapausing eggs. This inability is caused by a loss of responsiveness to the chemical signal that induces sex (Stelzer, 2008). As a consequence, populations of obligate parthenogens can grow to extremely high population densities, without ever inducing sex, whereas cyclical parthenogens readily induce sexual reproduction as soon as population densities exceed one female per ml (Gilbert, 2004; Stelzer, 2012). On the other hand, obligate parthenogenesis may be associated with small body size (Bennet and Boraas, 1989; Stelzer et al., 2010). The linked features of adaptive dwarfing and inflation of population size reflect a bacteria-like reproduction strategy, replacing K-selected sexual reproduction by r-selected asexual reproduction when facing high-resource conditions.In most cases, however, obligate parthenogens are of similar size, or even larger than their closest sexual relatives – the latter can usually be attributed to polyploidy (Suomalainen, 1950).

15.3.2.2 Daphnia

Members of the genus Daphnia are small (1–5 mm), largely transparent crustaceans that are found in most still freshwater bodies around the world. They are planktonic filter feeders, eating mainly planktonic algae and are themselves the food for planktivores, such as fish and some invertebrates. Most members of the genus Daphnia are cyclic parthenogens. Under most conditions, a female Daphnia will produce daughters by apomictic parthenogenesis. Induced by changes in the external environment, the same female can produce asexual sons or haploid eggs that need fertilization by males (Eads et al., 2008).Ephemeral populations of Daphnia tend to follow boom-and-bust cycles, whereby high densities often precede steep declines (Dudycha, 2004). Crowding may also indicate decreasing water levels as a result of the pond drying up. Either of these situations should favor investment in diapause as opposed to neonate production. Accordingly, Daphnia have been found to produce more diapausing eggs and fewer neonates in response to crowding (Berg et al., 2001; Lürling et al., 2003). These eggs are also resting stages (=ephippia) and can outlive harsh conditions for many decades. Diapausing eggs may hatch the following year, or may remain buried and hatch many years later (Kerfoot and Weider, 2004). Certain Daphnia are even able to produce these resting eggs without meiosis and thus do not need males (obligate parthenogenesis). Genotype x environment interactions affect responses to sex induction cues (Deng, 1996). These cues include high population density, food stress, photoperiod and temperature (Carvalho and Hughes, 1983; Larsson, 1991; Kleiven et al., 1992). Moreover, exposure to parasites (Duncan et al., 2006) and predators (Sluzarczýk, 1995) may induce the sexual formation of dormant stages to ensure persistence in the habitat. Offspring sex ratios, and the very direction in which they changed in response to crowding, differed significantly among genotypes with some genotypes producing more and others fewer males in response to crowding (Fitzsimmons and Innes, 2006). Obligately parthenogenetic genotypes seemed to respond to the crowding stimulus in similar ways as the facultatively parthenogenetic genotypes (Fitzsimmons and Innes, 2006). The frequency of sex, calculated using the proportion of males and sexually reproducing females, may span an ~30-fold difference between high- and low-sex populations (Cáceres and Tessier, 2004). Both the costs and benefits of sex, as measured by changes in means and variances in life-history traits, increased substantially with decreasing frequency of sex (Allen and Lynch, 2012). Brood sizes may vary enormously between maternal environments—from large broods under adult good conditions to very small broods and sexual resting eggs under adult stress conditions (Mitchell and Read, 2005). The cost of males was evident when sexual and asexual females were raised separately: sexuals produced fewer female offspring. However, there was no cost of males when reproductive modes were raised in pairs, as sexuals won the competition with asexuals (Wolinska and Lively, 2008).

The burst of clonal variation immediately after a sexual generation, can be lost rapidly during the asexual phase of the life cycle. Typically, it is assumed that selection at least contributes to this loss, which is supported by studies demonstrating that clones of cyclical parthenogens that are distinguishable by molecular markers also differ significantly in ecologically relevant traits (Weider, 1993b; Epp, 1996). Quantitative traits were influenced by three factors: (1) clonal selection significantly changed the population mean phenotype during the course of the growing season; (2) sexual reproduction and recombination led to significant changes in life-history trait means and the levels of expressed genetic variation, implying the presence of substantial nonadditive genetic variation and genetic disequilibrium; and (3) egg-bank effects were found to be an important component of the realized year-to-year change (Pfrender and Lynch, 2000). DNA transposons constitute approximately 0.7% of the D. pulex genome. Differences between lineages where sex was prohibited or promoted indicate that recombination has significant effects on TE dynamics (Schaack et al., 2010b).

15.3.2.3 Aphids

Aphids comprise worldwide ~4,400 known species of small sap-sucking insects of the order Hemiptera and are highly specialized plant feeders (Blackman and Eastop, 1994; 2000; 2006). Their life cycles are unusual among arthropods because they can include obligate shifting between unrelated host plant taxa, elaborate polyphenisms (‘‘the ability of organisms with the same genotype to develop two or more distinctly different alternative phenotypes without intermediates’’; Nijhout, 1999), and variation in reproductive strategy within a single species (Moran, 1992). Aphid life cycles can encompass cyclical parthenogenesis, obligate parthenogenesis, obligate parthenogenesis with male production and an intermediate ‘bet-hedging’ strategy where an aphid genotype will over-winter by continuing to reproduce by parthenogenesis and by investment in sexually produced eggs (Dedryver et al., 1998; 2001; Wilson et al., 2003). Under ideal conditions (climatic and with a dearth of predators, parasitoids and pathogens), a single parthenogenetic female can typically give rise to 30–90 offspring (Blackman, 1971) and, due to its short generation time (10 days), can potentially result in billions of individuals in a single growing season (Dixon, 1989; 1998; Harrington, 1994; Loxdale, 2009). Thanks to their aerial dispersal, aphids occur globally(Loxdale, 2009) and can be found in arctic (Strathdee et al., 1993) and sub-antarctic regions (Hullé et al. 2003), have been found far out at sea (Hardy and Cheng, 1986), on mountain tops (Loxdale, 2009), and infest a huge range of plants, usually monophagously. Although, some species are polyphagous within plant genera or families (e.g., Brassicas), and in the case of the ubiquitous and highly polyphagous Myzus persicae, this species attacks over 40 plant families worldwide (Blackman and Eastop, 2000).

Aphids are model organisms for the elucidation of phenotypic plasticity (Simon et al., 2011; Srinivasan and Brisson, 2012). In the pea aphid, Acyrthosiphon pisum, females respond to specific environmental cues by transmitting signals that have the effect of altering the development of their offspring. The production of alternative morphs, e.g. wing polyphenism (consisting of winged and unwinged females) and reproductive polyphenism (consisting of asexual and sexual individuals), by genetically identical individuals involves epigenetic mechanisms (Srinivasan and Brisson, 2012). These adaptations (sexual versus asexual, winged versus unwinged) have evolved in response to environmental changes that are predictable (seasons) and unpredictable but common (population density, host plant quality, and predation). Importantly, the developmental response for the morphs is separated by at least one generation from the triggering cue (Srinivasan and Brisson, 2012). Aphids express the wing phenotype to limit predation and competition for resources. The production of sexual morphs coincides with predictable, seasonal changes in photoperiod and temperature.Typically, the tropics and subtropics harbour only obligate parthenogens and cold climates only cyclical parthenogens, while both reproductive modes occur in temperate climate zones because their coexistence is facilitated by temporal variation in winter severity (Rispe and Pierre, 1998; Rispe et al., 1998). Sexual lineages are favored in regions with regular harsh winters because they produce cold-resistant eggs, while asexual lineages do not resist frost but take advantage of their faster multiplication rates in mild winter areas (Rispe and Pierre, 1998; Rispe et al., 1998; Simon et al., 2002). Clonality is associated with viviparity while sexuality is linked with oviparity (Le Trionnaire et al., 2008). With its large size (517 Mbp) and large number of predicted genes (35,000 genes) that is due to a large number of gene duplications, the pea aphid genome possesses one of the largest gene repertoires among animals (The International Aphid Genomics Consortium, 2010; Srinivasan and Brisson, 2012), rivaling that of Daphnia pulex, another polyphenic arthropod (Colbourne et al., 2011, see chapter 15.3.2.2). Reproductive polyphenism appears to depend on epigenetic modifications such as DNA methylation, chromatin remodeling by histone modificationsand regulation by small RNA pathways (Legeai et al., 2010; Walsh et al., 2010; Simon et al., 2011; Srinivasan and Brisson, 2012). Notable among the gene duplications are those of genes involved in DNA methylation, small RNA pathways, and chromatin modifications and remodeling. From an ecological point of view, plasticity of reproductive mode is dependent on epigenetic modifications. As in Daphnia, environmental change and hardiness favor sexual reproduction. Aphids present frequent transitions from cyclical parthenogenesis to permanent asexuality. Anholocyclics have abandoned sexual reproduction, reproducing all the year round by sustained parthenogenesis (Blackman, 1971; MacKay, 1989; Dedryver et al., 1998). Both nuclear and cytoplasmic markers of Rhopalosiphum padi clearly showed that many asexual lineages have hybrid origins between R. padi and an unknown sibling species, and are of recent origin (Delmotte et al., 2003). Mitotic recombination, although rare in asexual aphids (Sunnucks et al., 1996; Wilson et al., 1999), almost certainly occurs at a low rate or at specific locations in the genome (Blackman and Spence, 1996). In addition, patterns of recombination and segregation at sexual reproduction might be complex (Simon et al., 2002). For example, the aphid Myzus persicae shows high levels of recombination at oogenesis, but essentially no recombination during spermatogenesis (Sloane et al., 2001).

Clonal diversity may reflect the balance between the influx of clones and their elimination through drift and/or selection. That clonal diversity can erode quickly has been shown in holocyclic aphids. The burst of clonal variation immediately after a sexual generation can be lost rapidly during the asexual phase of the life cycle (Rhomberg et al., 1985). Typically, it is assumed that selection at least contributes to this loss, which is supported by studies demonstrating that clones of cyclical parthenogens that are distinguishable by molecular markers also differ significantly in ecologically relevant traits (Sunnucks et al., 1998; Turak et al., 1998; Vorburger et al., 2003; Vorburger, 2004). Accordingly, selection has a strong influence on the genotypic composition of aphid populations (Sunnucks et al., 1997; Guillemaud et al., 2003; Llewellyn et al., 2004; Zamoum et al., 2005). In semi-natural habitats at least, particular aphid clonal lineages are not that abundant, new genotypes being continually created by sexual recombination (annual) and mutation (ongoing) and are then mainly eliminated by natural selection/drift and competition for limited resources. Hence the retention of sex, even rare sex, is important in generating variation, more than is apparently possible by mutation alone within asexual lineages (Loxdale and Weisser, 2011). In sharp contrast to their southeast Asian and European counterparts, Sitobion miscanthi and S. near fragariae aphids in Australia and New Zealand exhibit a complete absence of sexual reproduction. A genetic analysis revealed stepwise mutation of microsatellite alleles and also karyotypic change, representing rare evidence of evolution and clonal selection within wild-living parthenogenetic lineages (Wilson et al., 1999; 2003). Strong clonal selection has also been demonstrated in the aphid Myzus persicae (Vorburger, 2005a) possibly based on strong fitness differences between clones (Vorburger, 2005b). In one aphid species, the pea aphid (Acyrthosiphon pisum), genetic variation for resistance to aphidiidine parasitoids has been demonstrated (Henter and Via, 1995; Hufbauer and Via, 1999; Ferrari et al., 2001; Stacey and Fellowes, 2002). Thus, it is possible that more resistant clones are under positive selection when parasitoids are abundant, but – if resistance incurs a cost – under negative selection when parasitoids are scarce (Vorburger, 2005a).

Most parthenogenetic individuals are eliminated by temperatures below –5 to –10°C, depending on species (Williams, 1980; Powell and Bale, 2004) and only fertilised eggs can resist long periods of intense frost (Sømme, 1969; Bale et al., 2007). A result of these selective patterns is that geographic parthenogenesis in aphids is “upside-down” compared to most organisms which have more parthenogenesis northwards (Hughes, 1989; Dedryver et al., 2001). In spring, holocyclic populations were in Hardy-Weinberg equilibrium at individual loci and had a relatively high genotypic diversity. Conversely, anholocyclic populations deviated from Hardy—Weinberg equilibrium and often consisted of a single clone (Simon et al., 1996). These findings argue for the primacy of ecological, and not genetic determinants, for the reproductive mode of aphids (Simon et al., 2002).

15.3.3 Geographical parthenogenesis: when males are too stressed to reproduce

Various quotations taken from Darwin (1859) by Hoffmann and Parsons (1991) indicate that Darwin already appreciated the impact of climatic stress at the extremes of species distributions. Often geographic ranges end at seemingly arbitrary points in space (Kirkpatrick and Barton, 1997; Holt and Keitt, 2005; Gaston, 2009). Historically, ecologists and biogeographers have correlated range boundaries with climate to identify environmental determinants of range boundaries (Griggs, 1914; Good, 1931; Dahl, 1951). Subsequent analyses have shown that range limits are associated with abiotic variables such as temperature or precipitation (Root, 1988; Cumming, 2002), biotic factors such as competitors (Terborgh and Weske, 1975; Bullock et al., 2000), or complex interactions between biotic and abiotic variables (Randall, 1982; Taniguchi and Nakano, 2000). A continuum emerges whereby competition is a factor of consequence in relatively benign environments but less so in more marginal environments where environmental stress is more important as selective regime (Parsons, 1993). Several hypotheses for the evolutionary stability of range limits propose that populations at range boundaries do not have sufficient genetic variation to respond to natural selection (Bradshaw and McNeilly, 1991; Hoffman and Blows, 1994; Gaston, 2003; Pujol and Pannell, 2008). Other hypotheses focus on other factors that may prevent populations from adapting to the environment at the range margin, such as genetic trade-offs among fitness-related traits in the marginal environment (Antonovics, 1976), genetic trade-offs between fitness in central and border environments (Holt, 2003), or gene flow from populations adapted to the range center (Haldane, 1956; Garcia-Ramos and Kirkpatrick, 1997; Kirkpatrick and Barton, 1997). These hypotheses are not necessarily mutually exclusive, and may act synergistically to constrain range expansion. All of these hypotheses are united by the assumption that populations are maladapted at a range boundary and unfit beyond the current range (Angert and Schemske, 2005). Range equilibrium is suggested by transplant experiments beyond species’ current range boundaries; many species have low fitness and exhibit negative population growth in areas beyond present distribution limits (e.g., Angert and Schemske, 2005; Geber and Eckhart, 2005; Griffith and Watson, 2006; Sexton et al., 2009; but see Van der Veken et al., 2007).There are empirical examples of high levels of mortality at range limits (Dekker and Beukema, 1993; Ungerer et al., 1999; Edwards and Hernández-Carmona, 2005; Angert, 2006; Gaston, 2009), and in transplants beyond those range limits (Angert and Schemske, 2005; Bird and Hodkinson, 2005; Geber and Eckhart, 2005). Greater levels of physiological stress at the range limits appear to be causally involved (Parsons, 1991; Tomanek and Somero, 1999; Harley, 2003; Sorte and Hofmann, 2004; Joyner-Matos, 2007; Joyner-Matos et al., 2007; Normand et al., 2009). In many cases, these high rates of mortality occur during or shortly after extreme environmental events. It is thus noteworthy that, conversely, most (especially ‘correlative’) modelling of the distributions of individual species employs spatial variation in mean not extreme environmental conditions (Gaston, 2003). It may not always, however, be the extreme events per se that are significant but the interaction between these events and resource availability, such that insufficient resources are available to enable organisms to cope with the demands of those events. In this sense, physiological and resource limitations on species distributions may be strongly confounded (Gaston, 2009). Inadequate levels of successful reproduction are probably one of the most common demographic explanations for range limits. There is much evidence for reduced or outright failure of reproduction at range limits, particularly among plants (where, if taking place, it is more readily observed), and a sense that this may often be more important in range limitation than subsequent survival of offspring (Gaston, 2009). Depending on taxon, changes in reproduction at range limits can variously include levels of sexual reproduction (Dorken and Eckert, 2001; Tremblay et al., 2002; Beatty et al., 2008).

Taking into account the pattern of stress-sex relationship as discussed in chapter 14 it can be expected that extreme environments shift the sex-asex balance. The sexual forms have a central, or limited, distribution while clonally reproducing all-female lineages of plants and animals most often surround the central area or arefound towards e.g. higher altitudes or latitudes. In a majority of the cases the clones also have a much wider distribution and ecological tolerance than the sexual forms they originate from (Beaton and Hebert, 1988; Parker and Niklasson, 2000; Schön et al., 2000; Stenberg et al., 2003). This spatial distribution of clones has been attributed to reproductive mode, elevated ploidy level or hybrid origin (Kearney, 2005) or all of these.

The great diversity of sex determination mechanisms in animals and plants ranges from genetic sex determination (e.g. mammals, birds, and most dioecious plants) to environmental sex determination and includes a mixture of both, for example when an individual’s genetically determined sex is environmentally reversed during ontogeny. Environmental sex determination and environmental sex reversal can lead to widely varying and unstable population sex ratios (Stelkens and Wedekind, 2010). Many reptiles, amphibians, fish, and insects are capable of undergoing either environmental sex determination or sex reversal at early life stages (Stelkens and Wedekind, 2010). There are many different triggers for sex reversal. Most of them are abiotic (e.g. temperature, pH, endocrine-disrupting hormones, photoperiod, hypoxia). Temperature-dependent sex determination has been extensively studied in reptiles, where exposure to elevated temperature results in female development in some species (Bull and Vogt, 1979). These temperature-dependent effects appear to be mediated in part by influencing aromatase activity and estradiol synthesis in female lizards, turtles and crocodilians, and by steroid receptors in both sexes (Crews and Bergeron, 1994; Wibbels and Crews, 1994; Crews, 1996; Lance, 2009). Intriguingly, blocking estrogen synthesis with aromatase inhibitors in embryos of a unisexual, parthenogenetic all-female whiptail lizard, Cnemdophorus uniparens, resulted in male offspring (Wibbels and Crews, 1994; Wennstrom and Crews, 1995). The sex of several fish species/populations has been shown to be under the control of many genes of minor effect [e.g. poecilid fish (Volff and Schartl, 2001), sea bass (Vandeputte et al., 2007), and tilapia (Baroiller et al., 2008)] and these are particularly good examples of the evolutionary lability of genetic sex determination systems (see also chapter 14.2.1). Asexuality has arisen via hybridization at least 90 times in vertebrates, and all of them are fish, amphibians or reptiles (Dawley and Bogart, 1989; Avise, 2008). I do not insinuate that environmental sex determination systems and parthenogenesis have the same evolutionary basis, but it is striking that both occur in the same or related species, the lability of genetic sex determination possibly being a common denominator.

Clonally reproducing all-female lineages of plants and animals are often more frequent at higher latitudes and altitudes, on islands, arid environments and in habitats described as transient, ecotonal, disturbed or marginal (Glesener and Tilman, 1978; Bierzychudek, 1985; Beaton and Hebert, 1988; Cuellar, 1994; Peck et al., 1998; Eckert, 2002; Kearney, 2003; Hörandl, 2006). That it is parthenogenesis, and not outcrossed sexuality, that prevails in harsh, uncertain, disturbed and novel conditions has been taken as evidence against the idea that sex is creating preadaptation to an uncertain future, either permitting species to adapt more quickly or enabling individual females to produce a few unexpectedly fit offspring (Bell, 1985). Attempts to explain this pattern, known as geographical parthenogenesis, generally treat the parthenogens as fugitive species that occupy marginal environments to escape competition with their sexual relatives (Vrijenhoek and Parker, 2009). Parthenogenetic reptiles are found in arid environments (Wright and Lowe, 1968; Darevsky et al., 1985; Adams et al., 2003; Kearney et al., 2003), another general pattern under the umbrella of geographical parthenogenesis. The aridity of the environments occupied by parthenogenetic lineages of Heteronotia has been quantified relative to the sexual races (Kearney et al., 2003). While the two sexual progenitor lineages of parthenogenetic Heteronotia also occur within the arid zone of Australia, the parthenogenetic forms inhabit the driest regions and their distributions are very tightly associated with rainfall contours (Kearney et al., 2003; Strasburg et al., 2007). Moreover, the most recently discovered case of natural parthenogenesis in reptiles, the scincid lizard Menetia greyii, also occurs in this region (Adams et al. 2003), along with a diverse array of other parthenogenetic taxa (Kearney, 2003; Kearney et al., 2006).

Small populations frequently occur in marginal environments or near the edges of geographic distributions. These situations are associated with exposure to unfavorable conditions (Brown, 1984; Hoffmann and Blows, 1994). Small populations are expected to suffer particularly from Muller’s ratchet (Muller, 1964; Felsenstein, 1974; Gabriel et al., 1993; Charlesworth and Charlesworth, 1997). Hence, within the conceptual framework of Muller’s ratchet theory of sexual reproduction (see chapter 18.1), it should be highly counter-intuitive that asexual reproduction occurs preferentially in small marginal populations.

There are two general explanations for this pattern, the General Purpose Genotype (GPG) and Frozen Niche Variation (FNV). In essence, the GPG and FNV models view clonal lineages as generalists versus specialists, respectively. According to the GPG model (Parker et al., 1977; Lynch, 1984), an asexual species consists of clones that can all survive and reproduce in all the different niches. The GPG postulates that asexuals do not need to adapt at all to changing environments, for example to fluctuating climates, if they have a genotype which allows them to survive in a wide range of ecological conditions. If this genotype produces a phenotype with wide ecological tolerance, then the GPG hypothesis predicts that the resulting phenotype will have (i) a broad tolerance against a wide range of environmental factors; and (ii) a very low variance in the tolerances of phenotypes derived from different populations. Both predictions were confirmed in populations of Darwinula stevensoni by Van Doninck et al. (2002): the tested specimens (with similar genotypes; Schön et al., 1998 and Van Doninck et al., 2002) showed a wide tolerance for a mixture of temperature and salinity treatments, while a logit linear model analysis of the “survival” data showed that responses between animals from several freshwater lakes (Ireland, France) were indistinguishable. Only the responses of animals from a (slightly) saline Belgian lake deviated to some extent, which might indicate that there was a maternal effect. Responses from other darwinulid species varied: Penthesilenula brasiliensis (a species with intercontinental distribution) has even wider tolerances than D. stevensoni for some variables, while endemic darwinulids such as Vestalenula molopoensis had much narrower tolerance ranges and P. aotearoa is found in-between (Van Doninck et al., 2003b; Schön et al., 2009).

One could argue that a GPG can only persist in fully asexual lineages as recombination will almost certainly break-up the allele-combinations required for a GPG. Therefore, a GPG can also only originate and persist in fully asexual lineages. However, the chance that a sexual lineage would have a GPG at the time clones originate from this sexual root (and can thus freeze the GPG in the clonal lineage) is very small indeed (Van Doninck et al., 2003b), which might explain why few taxa seem to have evolved a real GPG. This represents something of a paradox, as this would imply that a GPG can only evolve through adaptation in asexual lineages, which by definition have impeded evolvability.One of the key generalizations of the GPG theory is that thelytokous races are distributed in areas of lower biotic diversity and more abiotic, mainly temperature, seasonality (Hoy Jensen et al., 2002). From this, Levin (1975) and Glesener and Tilman (1978) proposed the ‘‘biotic uncertainty hypothesis’’ for the maintenance of genetic recombination as an adaptation for frequency-dependent selection response to coevolutionary changes in predators, parasites and/or competitors (Jaenike, 1978; Hamilton, 1980; Hoy Jensen et al., 2002). Under this model, asexual derivatives can only establish themselves in areas of lower (or non-coevolved) biotic diversity, mainly because of their colonizing ability, putatively high reproductive rates and independence from males in environments subject to density-independent mortality factors (Baker, 1965; Gerritsen, 1980; Ladle et al., 1993). Declines in within-population genetic diversity and/or increases in among-population differentiation towards range margins of sexual populations of plants and animals using nuclear molecular genetic markers were detected (Eckert et al., 2008). Small, marginal or colonizing populations go through periods of inbreeding. Inbreeding has adverse effects primarily on male fertility in many animals, including insects and mammals (Saccheri et al., 2005; Asa et al., 2007; Fitzpatrick and Evans, 2009; Zajitschek et al., 2009; Malo et al., 2010; Okada et al., 2011). Moreover, inbreeding generally increases the sensitivity of a population to stress, thereby increasing the amount of inbreeding depression (Miller, 1994; Pedersen et al., 2011; Bijlsma and Loeschcke, 2012). The joint effect of inbreeding and stress is more pronounced for male than for female reproductive performance (Enders and Nunney, 2010; Pedersen et al., 2011).

If marginal populations of sexual ancestral species are subject to genetic drift and inbreeding relative to central populations (da Cunha et al., 1950), then the hypothesized advantages of sex (variation and recombination) may be reduced or lost, even in the presence of coevolving enemies. Moreover, a theoretical model suggests that natural selection for maintenance of adaptation to habitats that contribute little to the population's reproduction, is weak. This can result in loss of fitness in such marginal habitats and involves accumulation of mutations that are deleterious in the marginal habitat but neutral or nearly so in the main habitat (Kawecki et al., 1997), eroding the advantage of sexual reproduction in marginal habitats. With a few exceptions, experimental studies on parthenogenetic animals (Gade and Parker, 1997; Robinson et al., 2002; Vorburger et al., 2003; Vrijenhoek and Parker, 2009; Loxdale et al., 2011) and apomictic plants (de Kovel and Jong, 1999) do not support the GPG hypothesis. In addition to Darwinula stevensoni (see above), the invasive allopolyploid apomictic grass weed, Pennisetum setaceum Forsk. Chiov. (fountain grass) may represent a GPG. Numerous molecular markers and extreme low quantitative trait variance in this grass weed indicate complete monoclonality. A single global genotype and widespread invasiveness under numerous environmental conditions even suggests a super-genotype. The super-genotype likely evolved high levels of plasticity in response to fluctuating environmental conditions during the Early to Mid Holocene (Le Roux et al., 2007).

Alternatively, the FNV model (Vrijenhoek, 1979; 1984) postulates that clonal spin-offs from sexual populations will “freeze” the ecological niche of these sexual populations (e.g. with regard to tolerances related to temperature, salinity, oxygen, etc.). Because most sexual populations are adapted to current environmental conditions, their asexual spin-offs will generally inherit these limited tolerance ranges. However, since a species with mixed reproduction can have a large number of different clones spinning off from sexual ancestors [amongst other origins of asexual lineages; see for ostracod examples Schön et al. (2000) and Rossi et al. (1998)], the total ecological tolerance of a set of clones might still cover a wide range of environmental conditions. The phytophagous mite Brevipalpus phoenicis (Geijskes) is a species that is found in a wide range of environments. In a cross-transplantation experiment of mites from three populations from three different host plant species (citrus, hibiscus and acerola), fitness was seriously reduced when mites were transplanted to the alternative host plant species, except when the alternative host was acerola. Thus, B. phoenicis clones are specialized to different niches and thus the FNV best described the broad ecological niche of this species but that there was also some evidence for host plant generalization (Groot et al., 2005).

Various gynogenetic fish species are thought to be FNV model species. In gynogenetic systems, unreduced eggs are produced by an all-female species. However, gynogens require sperm of closely related sexual species to trigger embryogenesis, but normally the sperm does not contribute any genetic material to the offspring (Dawley, 1989; Vrijenhoek, 1994; Schlupp, 2005). Occasional leakage of genes from a paternal host into sperm-dependent clones may however provide a source of adaptive variation to circumvent the disadvantages of asexuality (Lamatsch and Stöck, 2009; Lampert and Schartl, 2010). Expression of paternal genes may provide a local adaptive advantage in physiological or phenotypic sexual mimicry traits (Beukeboom and Vrijenhoek, 1998). It has also been argued that paternal leakage leading to the expression of paternal genes plays a pivotal role to stop Muller’s ratchet (Schartl et al., 1995a; Schlupp, 2005; Loewe and Lamatsch, 2008). A number of recent studies have shown that parthenogens can have cryptic sex (e.g., D’Souza et al., 2006, Omilian et al., 2006; D’Souza and Michiels, 2009) and suggest that rare sexual processes may be more common than previously thought (Beukeboom, 2007; Lampert and Schartl, 2010).

Due to their sperm dependence, gynogens have to coexist with a closely related sexual species that is likely to be similar ecologically (Beukeboom and Vrijenhoek, 1998; Niemeitz et al., 2002; Choleva et al., 2008). Due to this exploitation of the host, gynogenesis has also been called “sperm parasitism” (Hubbs, 1964). Unless the unisexual form is constrained by a unique carrying capacity or by mating behaviors that limit its reproductive potential, the all-female form should eliminate its sexual host and thereby ensure its own demise (Clanton, 1934; Moore, 1976; Kawecki, 1988; Heubel et al., 2009). Coexistence would be greatly facilitated by resource partitioning that diminishes direct competition between the sexual parasite and its host (Stenseth et al., 1985; Schley et al., 2004) or may be stabilized by parasites (Hakoyama et al., 2001; Hakoyama and Iwasa, 2004). Therefore, it is not surprising that essentially all of the sperm-dependant parthenogens exhibit some degree of niche separation from their sexual hosts. It has been documented that asexual lineages may rarely use sperm from a non-parental species or even switch a host. This pattern most probably results from the expansion of gynogenetic lineages into new areas. Such expansion was independent of the original parental species, suggesting that sperm-dependence is not as restrictive to geographical expansion as previously thought (Choleva et al., 2008).

Sexual and asexual forms of Poeciliopsis live in the desert streams of Sonora, Mexico, and are exposed to environmental extremes, ranging from flash floods to hot, desiccating, residual pools. Poecilia formosa presumably originated through a single natural hybridization event (between P. latipinna and P. mexicana) about 120,000 generations ago (Schartl et al., 1995b; Lampert et al., 2005; Lampert and Schartl, 2008; Stöck et al., 2010). Asexual and sexual topminnows of the genus Poeciliopsis have roughly similar fecundity when in mixed populations, but in monocultures the sexual Poeciliopsis monacha has a higher fecundity (Weeks, 1995; Schlupp et al., 2010). The reproductive output of both sexuals and asexuals was strongly affected by density and was higher in lower densities. These findings are in agreement with studies by Hubbs (1964) comparing the offspring numbers of P. formosa and P. latipinna in dissections and Balsano et al. (1985) comparing the reproductive output of females of P. mexicana and P. formosa. Neonate survival has been found to be lower in Amazon molly, P. formosa, than in Sailfin molly under food stress (Tobler and Schlupp, 2010). Under benign conditions no differences were found (Hubbs and Schlupp, 2008). Two coexisting clones of the triploid gynogenetic fish P. 2 monacha-lucida differed dramatically with respect to survival during stress and swimming endurance in an artificial flume: clone MML/II had the best survival during heat and cold stress and the worst survival during hypoxic stress, whereas clone MML/I had the best survival during hypoxic stress and the worst during heat stress. Poeciliopsis monacha, the sexual species with which these clones coexist, had intermediate survival during heat and hypoxic stress and very poor swimming endurance in the flume (Vrijenhoek and Pfeiler, 1997). The physiological differences were considered consistent with the FNV model and provide some insights into environmental factors that affect the distribution and abundance of these fishes.

That gynogens are not resource-limited in their habitat is suggested by their lower food stress resistance compared to the sexual mollies that appear to live at their habitat carrying capacity. Considering their distinct niches, not surprisingly, food competition plays a minor role in mediating coexistence between closely related asexual and sexual mollies (Scharnweber et al., 2011). There is evidence for male discrimination ability in this unisexual–bisexual species complex (Schlupp, 2005). Males do discriminate at several levels: they have mating preferences for conspecific females (Ryan et al., 1996; Gabor and Ryan, 2001; Schlupp and Plath, 2005), produce more sperm in the presence of such females (Aspbury and Gabor, 2004), and transfer less sperm in matings with heterospecific females (Schlupp and Plath, 2005; Riesch et al., 2008). It has been shown that is primarily subordinate males that mate with unisexuals (Mckay, 1971) that risk being driven away by dominant males and have less time to identify the female as conspecific (Kawecki, 1988). In asexuals, selection among clones operates to overcome the mating preference for conspecifics in the sexual species by more aggressive mating behavior or by the evolution of sex-mimicry (Beukeboom and Vrijenhoek, 1998). Altogether, the gynogens appear to be sperm-limited (Mckay, 1971; Moore and Mckay, 1971; Moore, 1976; Moore, 1984; Schlupp and Plath, 2005; Riesch et al., 2008; Mee and Otto, 2010; Schlupp, 2010) so that they cannot fully exploit their habitats’ resources. Evidence suggests that in these stressful environments males from sexual species live at the edge of their reproductive capacity exhibiting reduced sexual activity (Plath, 2008).

The Phoxinus eos/Phoxinus neogaeus/hybrid gynogen complex of cyprinid fishes is widely distributed in north-eastern America and occupies very heterogeneous habitats (Angers and Schlosser, 2007). It originated by multiple hybridizations between males of the northern redbelly dace (P. eos) and females of the finescal dace (P. neogaeus) (Dawley et al., 1987; Goddard et al., 1998; Angers and Schlosser, 2007). Although these diploid hybrids reproduce by sperm-dependent parthenogenesis (Goddard et al., 1998), the exclusion mechanism,which normally clears the egg from the sperm, often fails in this hybrid complex, leading to an unusually high level of sperm incorporation. As a consequence, five different hybrid biotypes are found in the complex (Goddard and Dawley, 1990; Lamatsch and Stöck, 2009). Intriguingly, there was an extreme lack of clonal diversity in these gynogens across a range of habitat types (Elder and Schlosser, 1995). Despite its genetic uniformity, however, the P. eos-neogaeus clone is no less variable than its sexual progenitors, suggesting that this single genotype may actually respond to environmental variation with as much phenotypic variation as a genetically variable sexual population (Doeringsfeld et al., 2004). Data from a large number of additional sites indicated that the proportion of polyploid hybrids within an environment was negatively related to hybrid relative frequency. The incorporation and expression of a third genome in triploid and diploid-triploid mosaic biotypes derived from the gynogenetic clone significantly expanded phenotypic variation of the clone (Doeringsfeld et al., 2004).

Phoxinus eos was more abundant in active beaver ponds, while Phoxinus eos-neogaeus gynogens were at higher frequencies in shallow, collapsed pond and stream environments (Elder and Schlosser, 1995; Schlosser et al., 1998). The increased frequencies of gynogens in pelagic and benthic zones, along with their greater survival times under oxygen stress, indicate that the gynogenetic clone is more general in its use of marginally suitable habitats and is physiologically more tolerant to anoxic conditions than its sexual progenitors. Oxygen availability, has been shown to alter species composition of fish communities in northern temperate environments (Tonn and Magnuson, 1982) and oxygen level is spatially and temporally quite variable among habitat types within these drainages. In the majority of ponds sampled, parasite loads were higher on asexual than on sexual Phoxinus fishes (Mee and Rowe, 2006).

Other types of non mutually-exclusive hypotheseshave been proposed to explain geographical parthenogenesis (Hörandl, 2006; Jose and Dufresne, 2010). Asexuals can establish populations from a single individual and hence can colonize remote and disturbed areas more rapidly than sexuals [reproductive assurance hypothesis, (Baker, 1955; Cuellar, 1994)]. They do not suffer from genetic bottlenecks at low population densities and hence may outcompete sexuals at the edge of a geographical range (Peck et al., 1998; Parker andNiklasson, 2000; Haag and Ebert, 2004; Ben-Ami and Heller, 2007; Golovatch and Kime, 2009; Guzmán et al., 2012). Under an ecological scenario, asexuals are better competitors or have fitness advantages over sexuals under certain ecological conditions (Glesener and Tilman, 1978), owing largely to their frequent hybrid origins. Historical explanations refer to the association between parthenogenesis and environments that were strongly affected by the Pleistocene glacial cycles (Stebbins, 1984; Kearney, 2005). Therepeated advances and retreats of glaciers have resulted in the creation of refugial races which were free to colonize new environments as glaciers retreated. Due to their better colonizing abilities, asexuals may have been able to follow glacial retreats faster than sexuals, hence the pattern of geographical parthenogenesis.

Many studies argue that plasticity enhances ecological niche breadth because plastic responses allow organisms to express advantageous phenotypes in a broader range of environments (Bradshaw, 1965; Van Valen, 1965; Whitlock, 1996; Sultan et al., 1998a; b; Donohue et al., 2001; Sultan, 2001; Richards et al., 2005; 2006). In invasion biology there are two primary scenarios which describe how a different reaction norm might contribute to invasion success: (i) a jack-of-all-trades situation, where through the plasticity of morphological or physiological traits, the invader is better able to maintain fitness in a variety of environments, a characteristic clearly related to the concepts of a general purpose genotype (Baker, 1965); (ii) a master-of-some situation, in which the plasticity of morphological or physiological traits allows the invader to take advantage of favorable environments; in addition, an invader might be (iii) a jack-and-master that combines some level of both of these abilities (Richards et al., 2006; Scheiner and Holt, 2012). Theories of the evolution of niche breadth have traditionally depended on the assumption that a ‘‘jack-of-all-trades is a master of none’’ (Levins, 1968; MacArthur, 1972a; Futuyma and Moreno, 1988; Rausher, 1988; Fry, 2003; Richards et al., 2006; Palaima, 2007). According to this view, no genotype can have maximal fitness in each of a set of environments (e.g., habitats or hosts), so that a population’s improvement in fitness in one environment comes at the expense of its fitness in others. Evolution of generalization necessarily entails a cost in terms of fitness loss elsewhere along an environmental gradient that leads to a genetic fitness trade-off between a generalist and a specialist (Palaima, 2007). However, empirical evidence that generalists perform less well than specialists on shared resources is scarce (Strickler, 1979; Dykhuizen and Davies, 1980; Bernays, 2001; Straub et al., 2011) and the issue has remained contentious (Huey and Hertz, 1984; Bernays and Graham, 1988; Futuyma and Moreno, 1988; Berenbaum, 1996; Fry, 1996; Palaima and Spitze, 2004).

Computer simulations support the idea that a single basic process may account for much of what is known about geographic parthenogenesis (Peck et al., 1998). Chapter 14.1 presents comprehensive evidence that, compared to female reproduction, male reproduction is more sensitive to environmental disruption. This single factor fits perfectly well with the ecological and experimental data of geographic parthenogenesis. Importantly, the geographic pattern due to various stress factors is superimposed by other factors that determine the sex-asex distribution like resource availability in relation to reproductive resource investment, other strategies that increase phenotypic plasticity like polyploidy, and lability of asex/sex expression e.g. in modular organisms (see chapter 15.2). In mammals and birds, embryonic development at the time of sex determination occurs under controlled temperature conditions. However, fish are poikilothermic, and embryonic development proceeds in full exposure to the external physical environment where relatively large temperature alterations can occur. Sex determination can be influenced by external physical variables such as temperature in most fish families examined. In some cases, the species utilize these influences as a strategy to improve reproductive success, whereas in others, the effects on sex determination may not occur naturally, and may arise from disruptions of normal sex-determination processes under extreme environmental conditions (Devlin and Nagahama, 2002). In sexual Poecilia sphenops from Oaxaca higher temperatures appear to result in a female-biased sex ratio (Barón et al., 2002). Poikilotherm fishes display a male lower-temperature comfort zone. Male guppies prefer a significantly lower temperature (24.5°C) than females (28.2°C) or juveniles (28.1°C). Treatment of juveniles and females with testosterone lowers their preferred temperature to that of males (Johansen and Cross, 1980). This gender-differential temperature preference, at least during the reproductive period, has been observed in other fish species as well (Hagen, 1964; Baker et al., 1970; Swain and Morgan, 2001; Hernández-Rodríguez et al., 2002; Podrabsky et al., 2008). In rainbow trout, the maximal heat shock response of male germ cells, that are located in the same body compartment like the other organs, occurs at a significantly lower temperature (22°C) than for somatic cells (28°C) (Le Goff and Michel, 1999). Reproduction in fishes is especially sensitive to disturbance by hypoxia (Poon et al., 2001; Wu et al., 2003; Thomaset al., 2006). Asexual Poeciliopsis are living in an ecological niche that, due to its environmental stress, is inaccessible to sexual Poeciliopsis. Obviously in summer, water temperatures and/or water hypoxia rise that, in the gynogens’ niche, may be deleterious to male gametogenesis (Wu et al., 2003; Landry et al., 2007; Thomas et al., 2007; Thomas and Rahman, 2010). Due to the higher oxidative stress exerted during male gametogenesis (resulting in male-driven mutagenesis), hyperthermia and hypoxia impair male fish gametogenesis more than the femaleone (Thomas et al., 2007; Podrabsky et al., 2008). In comparisons between sexuals and asexuals, males of the millipede Nemasoma varicorne (Enghoff, 1976; Hoy Jensen et al., 2002) and the cockroach Pycnoscelus surinamensis (Niklasson and Parker, 1994; Parker and Niklasson, 1995; Gade and Parker, 1997) were found to be less tolerant to environmental stressors.

16. Germ granules and transgenerational epigenetic information transfer


Summary

Rather than the continuity of the germ cells, it is the continuity of substances passed down from the parent’s germ cells to the germ cells of the progeny that ensures the heredity of the species, hence the collective term ‘‘germ plasm’’ to represent these substances. Electron microscope techniques led to the description of germ cell-specific structures with granular or fibrous shape, no confining membrane,located in the perinuclear cytoplasm, and usually associated with clusters of mitochondria. Germ granules are thought to be a signature feature of germ cells in animals and the site ofnon-coding RNA biogenesis that play a key role in epigenetic regulation. What makes epigenetic processes fundamentally different from genetic processes is that in some cases environmentally induced epigenetic changes may be inherited by future generations. Epigenetics is closely linked to environmental conditions and mitochondrial bioenergetics. Thus, the epigenome provides the interface between the environment and the regulation of nuclearDNA gene expression. Conditions of stress seem to be particularly important as inducers of heritable epigenetic variation and lead to changes in germline epigenetic and genetic organization. A multitude of mechanisms may convert reversible epigenetic changes into stable epigenetic and genetic transgenerational effects. Cumulative evidence suggests that cytoplasmic ribonucleoproteins that are stabilized in germ granules are the carriers of transgenerational information. The following scenario is proposed: (i) in a close cross-talk with germline cells, somatic gonadal cells determine the ncRNA profile in the germline cells; (ii) after intercellular transfer secondary piRNA biogenesis (ping-pong) amplifies the message in the germline. (iii) the ncRNA message is stored in the germ granules until the information is retrieved for embryonal development (preformation); (iv) in species with extended parent-embryo/fetus developmental support (e.g. in mammals) reprogramming by transgenerational message transfer is not confined to gametogenesis (preformation) but occurs also during embryonal development (epigenesis).

16.1 Germ granules are germ cell markers

Rather than the continuity of the germ cells, it is the continuity of substances passed down from the parent’s germ cells to the germ cells of the progeny that ensures the heredity of the species, hence the collective term ‘‘germ plasm’’ to represent these substances (Weismann, 1892; Gao and Arkov, 2012). The first evidence supporting a germ plasm model came from the successful tracing of granules from the posterior pole cytoplasm of insect oocytes in one generation to the germ cells of the next generation, with the substances referred to as ‘‘germ cell determinants’’ (Hegner, 1914). Also, the term ‘‘chromatoid body’’ was introduced by early investigators to describe a germ cell organelle in mammalian spermatocytes and spermatids, based on the fact that this exhibits structure similar to chromosomes and nucleoli when examined under a light microscope (Benda, 1891; Hermann, 1889; Yokota, 2008). Today, germ plasm can be recognized by various distinctive features (Eddy, 1975), including electron-dense granules composed of ribonucleoprotein complexes, variable association with dense concentrations of mitochondria and nuclear pores and all or part of a conserved set of mRNAs and proteins (notably Piwi, Nanos, Vasa, PL10, Pumilio, Boule/Dazl and Bruno) involved in transposon silencing and mRNA regulation (Ewen-Campen et al., 2010; Juliano et al., 2010; Voronina et al., 2011). The employment of electron microscope techniques from the 1950s to 1970s led to a more detailed morphological description of the germ cell-specific structures: high electron density, granular or fibrous in shape, no confining membrane, frequently surrounded by small vesicles, usually associated with clusters of mitochondria, and located in the perinuclear cytoplasm (Mahowald, 1962; Brokelmann, 1963; Fawcett et al., 1970; al-Mukhtar and Webb, 1971; Eddy, 1974; Russell and Frank, 1978; Gao and Arkov, 2012). Germ cells in more than 80 animals from at least eight phyla contain this characteristic morphological feature variously called germ plasm or germinal plasm, germ granules, sponge bodies, dense bodies, chromatoid bodies, Balbiani bodies, mitochondrial clouds, polar plasm, pole plasm, oosome, nuage. The dynamics of organelle movement during the assembly of these aggregates also shows striking similarity between different animals (Heasman et al., 1984; Holland and Holland, 1992; Carré et al., 2002; Kloc et al., 2004a). Close association of the germ granules with nuclear pores has been reported in ultrastructural studies of C. elegans (Strome and Wood, 1982; Pitt et al., 2000; Sheth et al., 2010), Xenopus (Czolowska, 1969), zebrafish (Knaut et al., 2000; Kloc et al., 2004a), and mouse (Chuma et al., 2009). Thus, germ granules are optimally positioned for cytoplasmatic-nuclear information transfer. Germ granules are thought to be a signature feature of germ cells in animals (Mahowald, 1968; al-Mukhtar and Webb, 1971; Eddy, 1974; 1975; Wilsch-Brauninger et al., 1997; Houston and King, 2000; Extavour and Akam, 2003; Kloc et al., 2004a; Snee and Macdonald, 2004; Updike and Strome, 2010). Germ granules are found in germ cells in many stages of development, ranging from primordial germ cells (PGCs) in embryos to gametes in adult gonads. In many animals nuage has been shown to contain a combination of RNAs, proteins, endoplasmic reticulum and mitochondria, and may sometimes contain other organelles (such as microtubules) as well. There are two key routes to the embryonal initiation of the germ cell lineage (Extavour and Akam, 2003; Seydoux and Braun, 2006; Extavour, 2007; Rosner et al., 2009). One is through the inheritance of preformed germ cell determinants or germ plasm as observed in Drosophila melanogaster, Caenorhabditis elegans and Xenopus (Eddy, 1975; Saffman and Lasko, 1999; Wylie, 1999). The other route occurs in mammals, where a group of pluripotent cells are first established with seemingly equivalent potential from which both germ cells and somatic cells are derived (McLaren, 1999; 2000; Saitou et al., 2002; Surani et al., 2004; 2008; Saitou, 2009). These two modes of germ cell specification are referred to as preformation and epigenesis, respectively (Extavour and Akam, 2003). The two mechanisms are not necessarily mutually exclusive, but rather are better viewed as two extremes of the continuum along which development of germ cells can be mapped, since at some stage of germ cell development, both types of mechanism are inevitably used (Extavour, 2007).

16.2 Transgenerational epigenetic information transfer

Accumulating evidence indicates that both genetic and non-genetic inheritance, and the interactions between them, have important effects on evolutionary outcomes. There is increasing awareness that epigenetic information can also be inherited across generations (Grishok et al., 2000; Richards, 2006; Jirtle and Skinner, 2007; Alcazar et al., 2008; Youngson and Whitelaw, 2008; Cairns, 2009; Hammoud et al., 2009; Nadeau, 2009; Slatkin, 2009; Cuzin and Rassoulzadegan, 2010; Daxinger and Whitelaw, 2010; 2012; de Boer et al., 2010; Nelson and Nadeau, 2010; Burton et al., 2011; Danchin et al., 2011a; Day and Bonduriansky, 2011; Guerrero-Bosagna and Skinner, 2012; Nelson et al., 2012; Lim and Brunet, 2013).

What makes epigenetic processes fundamentally different from genetic processes is that in some cases environmentally induced epigenetic changes may be inherited by future generations (Richards, 2006; Whitelaw and Whitelaw, 2006; Jirtle and Skinner, 2007). For instance, Fieldes and Amyot (1999) experimentally altered DNA methylation in flax and showed that this significantly affected the phenotypes of at least four generations of progeny. In mice, environmental toxins (Anway et al., 2005; Crews et al., 2007) and dietary supplements (Cropley et al., 2006) induce changes in DNA methylation that are inherited over several generations (Bossdorf et al., 2008). In Drosophila, experimental reduction of the heat shock protein Hsp90 (which also occurs in response to environmental stress) (Jarosz and Lindquist, 2010; Jarosz et al., 2010) causes stable phenotypic changes which appear to be due to the release of hidden epigenetic variation (Sollars et al., 2003). Moreover, Hsp90 can regulate the length of trinucleotide repeats to fine-tune gene function and can regulate the mobility of transposable elements to enable larger functional changes (Fonville et al., 2011). Non-genetic inherited information can arise through several interacting mechanisms, including epigenetics, parental effects and ecological and cultural inheritance (Hercus and Hoffmann, 2000; West-Eberhard, 2003; Jablonka and Lamb, 2005; Bonduriansky and Day, 2009; Helanterä and Uller, 2010).

An extrinsic transgenerational phenotype requires a continued multigenerational exposure to the factor (often at only a specific period of development) triggering an epigenetic change. For example, good maternal behavior towards offspring (e.g., early postnatal pup licking in rodents) can program the same good maternal behavior in the grown-up female adults that then pass this on to their offspring in a similar manner (Champagne, 2008; Mcgowan et al., 2008; Szyf et al., 2008). Rats that are nurtured by stressed mothers are more likely to be stressed (Francis and Meaney, 1999). This phenotype involves the setting of a ‘stressed’ state by the hypothalamic–pituitary–adrenal axis (HPA axis) in the pup (Weaver et al., 2004). However, without the continued generational maternal behavior and epigenetic programming of the brain and behavior in the female offspring, the transgenerational phenotype would be lost (Mcgowan et al., 2008; Szyf et al., 2008). The exposure also has the potential to promote intrinsic transgenerational phenomena, which will promote a transgenerational phenotype independent of continued environmental exposures (Anway et al., 2005; Skinner and Guerrero-Bosagna, 2009). Exposures of mother rats to particular endocrine disruptors can induce epigenetic changes in the male germline that are associated with changes in male fertility and reproductive behavior up to four generations later (Anway and Skinner, 2006; 2008). Several studies in mammals support the hypothesis that transgenerationally inherited epigenetic alterations affect the health and longevity of future generations (Morgan et al., 1999; Lane et al., 2003; Rakyan et al., 2003; Pembrey et al., 2006; Chandler, 2007; Youngson and Whitelaw, 2008; Franklin and Mansuy, 2010; Bonduriansky et al., 2012). Transgenerational epigenetic inheritance has also been demonstrated in other eukaryotic organisms, for example, plants, yeast, nematodes, rotifers, insects and fishes (Grewal and Klar, 1996; Hercus and Hoffmann, 2000; Chandler and Stam, 2004; Bashey, 2006; Chandler, 2007; Galloway and Etterson, 2007; Youngson and Whitelaw, 2008; Burns and Mery, 2010; Burton et al., 2011; Kaneko et al., 2011; Bonduriansky et al., 2012).

Epigenetics is closely linked to environmental conditions andmitochondrial bioenergetics (Naviaux, 2008; Smiraglia et al., 2008; Wallace and Fan, 2010; Minocherhomji et al., 2012; see chapter 10.3). The epigenome provides the interface between the environment and the regulation of nuclearDNA gene expression (Feinberg, 2007; 2008). Adverse environmental conditions play a key role in transgenerational inheritance. Conditions of stress seem to be particularly important as inducers of heritable epigenetic variation, and lead to changes in epigenetic and genetic organization that are targeted to germline specific genomic sequences (Badyaev, 2005a; Jablonka and Lamb, 2005; Jirtle and Skinner, 2007; Rando and Verstrepen, 2007; Jablonka and Raz, 2009; Boyko et al., 2010; Curley and Mashoodh, 2010; Franklin and Mansuy, 2010; Nätt, 2011; Seong et al., 2011). Stress-induced changes of DNA methylation are common and are mostly heritable (Chinnusamy and Zhu, 2009; Verhoeven et al., 2010a; Richards, 2011; Verhoeven and van Gurp, 2012). In B. subtilis, the physiological state of the cell's ancestor (more than two generations removed) does affect the outcome of cellular differentiation and bacterial aging by epigenetic inheritance (Veening et al., 2008b). In plants and aquatic invertebrates, intra- and interspecific competitive interactions and stressors leave their transgenerational signature on life history traits (Galloway and Etterson, 2007; Allen RM et al., 2008; Boyko et al., 2010). Intriguingly, the longevity-extending effect of dietary restriction is transgenerationally inherited by rotifer offspring (Kaneko et al., 2011). In placental animals, maternal conditions such as nutritional stress affect growth rate, immune capacity, survival and breeding performance of offspring (Festa-Bianchet and Jorgenson, 1998; Lummaa and Clutton-Brock, 2002; Lummaa, 2003; Jones OR et al., 2005). A mother’s diet during pregnancy may have an enduring influence on succeeding generations, independent of later changes in diet (Waterland and Jirtle, 2003; Lillycrop et al., 2005; Cropley et al., 2006; Langley-Evans, 2009; Chmurzynska, 2010). The inducibility and transmissibility of epigenetic variants depend on developmental conditions.

Germline and fetal programming of obesity, cardiovascular disease and insulin resistance has been investigated in a wide range of epidemiological and animal studies; these investigations elucidated transgenerational adaptations that effect epigenetic modification of genes involved in a number of key regulatory pathways with long-term sequelae for morbidity and mortality (Armitage et al., 2005; Gluckman et al., 2007; Symonds et al., 2009; Burns and Mery, 2010; Carone et al., 2010; Chmurzynska, 2010; Alfaradhi and Ozanne, 2011; Ferguson-Smith and Patti, 2011; Nätt, 2011; Ozanne et al., 2011; Rakyan et al., 2011; Skilton et al., 2011; Guilmatre and Sharp, 2012; Herring et al., 2012). Female rats that were exposed to a high-carbohydrate diet as neonates spontaneously transmitted the obesity phenotype to their offspring, thus establishing a vicious transgenerational effect (Srinivasan et al., 2003; Patel et al., 2009; Patel and Srinivasan, 2011). Similarly, F1 sons of female mice that are 50% dietary restricted during late gestation but fed ad libitum throughout their own life develop metabolic syndrome and their own F2 offspring also exhibit impaired glucose tolerance (Jimenez-Chillaron et al., 2009). In another study, when the fertilized eggs of adult females born to dietary restricted dams were embryo-transferred to control dams, the inheritance of metabolic syndrome was still observed suggesting that this transmission may occur via alterations in the germline of both parents (Thamotharan et al., 2007). In humans, a link between grandparental and parental periods of low or high food availability and offspring mortality and morbidity was found (Bygren et al., 2001; Kaati et al., 2002; 2007; Pembrey et al., 2006). Thus, evidence suggests that the quality of diet and the caloric regimen can influence the epigenetic state of the genome and its transgenerational modulation. Certain metabolic pathways are also epigenetically controlled revealing a tight crosstalk between metabolism and epigenomes. Particularly, methylation-based epigenetic reactions are controlled by different metabolic pathways and reciprocally can influence metabolism (Chiacchiera et al., 2013). Oxidative stress as the final common effector of stress signaling pathways (Lindquist, 1986; Sanchez et al., 1992; Finkel and Holbrook, 2000; Heininger, 2001; Mittler, 2002; Mikkelsen and Wardman, 2003; Sørensen et al., 2003; Apel and Hirt, 2004; Ardanaz and Pagano, 2006; Rollo, 2007; Miller et al., 2008; Slos and Stoks, 2008; Jaspers and Kangasjärvi, 2010; Steinberg, 2012; Choudhury et al., 2013) appears to mediate the epigenetic reprogramming (Franco Mdo et al., 2002; Luo ZC et al., 2006; Cyr and Domann, 2011). This makes perfect evolutionary sense since stress understood as ecological condition that erodes Darwinian fitness calls for (epi)genetic innovation. Intriguingly, stress exposure in intrauterine life (Entringer et al., 2010; 2011) and during early childhood (Drury et al., 2011) is associated with shorter leukocyte telomere lengths in young adulthood which is a predictor for earlier onset of age-related disease and mortality (Blackburn, 2000; Epel et al., 2004; Serrano and Andrés, 2004; Epel, 2009; Mather et al., 2011).

For intrauterine exposure, maternal-derived, embryonal stress hormones appear to be the mediators, arguing for a common neuroendocrinological mediation via stress-sensing pathways (Dufty et al., 2002; Drake et al., 2005; Robert and Bronikowski, 2010). In fishes and birds, content of stress hormones in egg yolk and albumen has adverse effects on offspring development and longevity (Eriksen et al., 2003; 2007; Hayward and Wingfield, 2004; Love et al., 2005; Rubolini et al., 2005; Saino et al., 2005; Love and Williams, 2008; Gagliano and McCormick, 2009; McCormick and Gagliano, 2009). Importantly, these effects are adaptive under environmental adversity (Meylan and Clobert, 2005; Love and Williams, 2008; Love et al., 2009). A central question in considering evolutionary change in response to environmental change is whether the inherited epigenetic markers could facilitate genomic change (Johnson and Tricker, 2010; Bateson, 2012). There are a multitude of mechanisms that may convert reversible epigenetic changes into stable epigenetic and genetic transgenerational effects (Pembrey, 1996; Young, 2001; Beaudet and Jiang, 2002; Jablonka, 2004; Cullis, 2005; Gallou-Kabani and Junien, 2005; Mittelman and Wilson, 2010). The epigenetic, selectable variation might enable a lineage to adapt and “hold” the adaptation until genetic changes take over; thus, the heritable epigenetic variations in protein architecture pave the way for genetic adaptation (True et al., 2004; Sangster et al., 2004; Jablonka and Raz, 2009). Thus, demographic responses can, over time, evolve into new, genetically mediated traits (Reznick et al., 1990; Kokko and López-Sepulcre, 2007; Jones et al., 2008; Mittelman and Wilson, 2010).

16.3 Are germ granules the vehicles of transgenerational epigenetic information transfer?

Unicellular organisms are directly exposed to the vicissitudes of the environment. This direct exposure elicits the stress responses that result in changes to their genetic make-up in a bet-hedging response. Multicellular organisms have evolved an internal milieu that is increasingly stable, protected from the environment by outer barriers. This improved barrier, however, has the disadvantage that the germline cells are increasingly shielded from the environment. Germline cells are not equipped to sense the outside environment and are dependent on messages from the neuroendocrine system about changes in their future biosphere. The gonads are integrated into a complex neuroendocrine network that can relay signals. We are only beginning to understand how these signals are transmitted to the germline cells and what (if any) (epi)genetic changes they may induce. Assuming that there is a transgenerational information transfer about environmental conditions, the question is which signals carry this information, how this information is transferred and what type of information can be transferred to the gametes and the embryo. Obviously, most relevant for the future struggle for survival is information about environmental resource availability and stressors. Importantly, throughout their development germline cells are in intimate contact with cells of the somatic gonads. For instance, germ cells in mammals develop in a microenvironment of supporting stromal cells of somatic origin that interact with the former through autocrine/paracrine mechanisms as well as direct cell-to-cell interactions (Matzuk et al., 2002). Deprived of this support, isolated germ cells in culture fail to survive and maintain their characteristics. Signals from these somatic gonad cells are vital for germ cell maturation. However, to act transgenerationally, this information must persist as molecularsignals even after the physical contact between gamete/zygote and somatic gonads has been interrupted.

The search for molecules that determine germ cell fate began when Illmensee and Mahowald (1974) showed that cytoplasm taken from the posterior of a Drosophila embryo (germ or pole plasm) was sufficient to induce ectopic germ cells, when injected into a host embryo. A major function of the large numbers of noncoding RNAs is to direct chromatin-modifying complexes to their sites of action (Mattick et al., 2009; Koziol and Rinn, 2010). Following the definitions introduced by Youngson and Whitelaw (2008), this may well be called an example of transgenerational epigenetic inheritance. The last decades have revealed that epigenetic processes operate in the transgenerational nongenetic determination of phenotype. This has been termed soft inheritance (Mayr, 1982), non-Mendelian inheritance, parental effects, and fetal programming, among others.

RNA interference (RNAi) is an evolutionarily conserved system mediated by, and targeted against, RNA. The RNAi pathway is triggered by long double-stranded (ds)RNA that is cleaved by an endonuclease named Dicer to produce short interfering (si)RNAs. A relatively complex RNAi machinery was already present in the last common ancestor of eukaryotes and consisted, at a minimum, of one Argonaute-like polypeptide, one Piwi-like protein, one Dicer, and one RNA-dependent RNA polymerase (Cerutti and Casas-Mollano, 2006). From a mechanistic standpoint, the ancestral RNAi machinery may have been capable of both siRNA-guided transcript degradation as well as siRNA-guided transcriptional repression of homologous sequences (Cerutti and Casas-Mollano, 2006). An intact Dicer/Argonaute RNA processing machinery both in somatic and germline gonadal cells is essential for gametogenesis and fertility (Deng and Lin, 2002; Kuramochi-Miyagawa et al., 2004; Murchison et al., 2007; Tang et al., 2007; Hayashi et al., 2008; Hong et al., 2008; Kalidas et al., 2008; Maatouk et al., 2008; Nagaraja et al., 2008; Otsuka et al., 2008; Gonzalez and Behringer, 2009; Ma et al., 2009; Papaioannou et al., 2009; 2011; Kim GJ et al., 2010; Lei et al., 2010; Korhonen et al., 2011; Romero et al., 2011; Liu D et al., 2012; Wu et al., 2012; Ortogero et al., 2013). In this context it is alarming that a multitude of environmental mutagens form stable complexes with Dicer, that are more stable than those formed by Dicer with its natural substrate, i.e. double strand short RNAs (Ligorio et al., 2011). This may be a critical factor in the human spermatozoon, already a cell in crisis operating near its threshold of error catastrophe, and a possible additional factor in a multifactorial process resulting in deteriorating human semen quality (see chapter 14.1).

It was demonstrated first in plants (Jones et al., 1999; Mette et al., 2000) and then in human cells (Kawasaki and Taira, 2004; Morris et al., 2004), that siRNAs can direct methylation of homologous DNA. When this DNA methylation affects a promoter sequence, it can lead to transcriptional silencing of the downstream gene. In plants and animals, gene silencing by RNA-directed DNA methylation (Aufsatz et al., 2002; Kanno et al., 2004; Kawasaki and Taira, 2004; Mathieu and Bender, 2004; Matzke et al., 2004; Morris et al., 2004; Matzke and Birchler, 2005; Cerutti and Casas-Mollano, 2006; Huettel et al., 2007; Aravin et al., 2008; Aravin and Bourc'his, 2008; Morris, 2008; Zhai et al., 2008; Law and Jacobsen, 2010; Bender, 2012; Lorkovic et al., 2012; Castel and Martienssen, 2013; Dalakouras and Wassenegger, 2013; Sabin et al., 2013) and histone modification (Volpe et al., 2002; Schramke and Allshire, 2003; Noma et al., 2004; Pal-Bhadra et al., 2004; Cerutti and Casas-Mollano, 2006; Weinberg et al., 2006; Gu et al., 2012; Huang et al., 2013; Sabin et al., 2013) has now been widely demonstrated. Importantly, DNA methylation (Rakyan et al., 2002; Waterland and Jirtle, 2003; Blewitt et al., 2006; Peaston and Whitelaw, 2006; Ekram et al., 2012) and histone modifications (Dolinoy et al., 2010) within metastable epialleles has a stochastic element due to probabilistic reprogramming of epigenetic marks during gametogenesis and embryogenesis. For instance, stress-induced methylation changes may be targeted specifically to stress-related genes. Alternatively, methylation changes may generate nonspecific (random) differences between individuals, which may have adaptive significance during times of stress (Rapp and Wendel, 2005), because they increase the range of variation that natural selection can act upon. Part of the variation may be due to off-target effects since siRNAs may cross-react with targets of limited sequence similarity (Jackson et al., 2003; Scacheri et al., 2004; Lin et al., 2005; Birmingham et al., 2006; Ma et al., 2006; Alemán et al., 2007) with possible adverse effects on cell viability (Fedorov et al., 2006).

Flies have five Argonaute proteins — molecular ‘scissors’ that use the small RNA guide to bind to and cut a second RNA molecule, the ‘target’: Ago1, which uses microRNAs to regulate gene expression; Ago2, which uses siRNAs to fight viral infection; and three closely related Piwi proteins — Piwi, Aubergine and Argonaute3 (Ago3) — which use piRNAs to safeguard the animal germline genome against deleterious retroelements (Lau et al., 2006; Saito et al., 2006; Vagin et al., 2006; Aravin et al., 2007a; Brennecke et al., 2007; Gunawardane et al., 2007; Houwing et al., 2007; Li C et al., 2009; Malone et al., 2009). These data have led to the hypothesis that Piwi/piRNA complexes might serve as sequence-specific guides that direct the de novo DNA methylation machinery to transposable elements (Aravin et al., 2007a; 2008; Kuramochi-Miyagawa et al., 2008; Zamudio and Bourc’his, 2010). In the Drosophila female germline, two distinct groups of piRNAs are involved in repressing the retroelements in different cell types: one in germline cells and the other in the ovarian somatic cells (Lau et al., 2009; Li C et al., 2009; Malone et al., 2009; Khurana and Theurkauf, 2010; Olivieri et al., 2010). Many transposons are expressed in germ cells, where movement can lead to heritable expansions in their number. Examples in Drosophila include TAHRE, TART, HetA, copia, and the I element (Vagin et al., 2004; Brennecke et al., 2008; Chambeyron et al., 2008; Shpiz et al., 2007; 2009). Some transposons are exclusively or additionally expressed in somatic cells of the ovary, with gypsy, ZAM, and idefix occupying this category in the Drosophila ovary (Prud’homme et al., 1995; Desset et al., 2003; 2008; Sarot et al., 2004; Mével-Ninio et al., 2007; Pélisson et al., 2007). In the D. melanogaster female germline TE silencing is dependent on the expression of its own copies and the presence of small RNAs in nurse cells surrounding the oocyte (Brennecke et al., 2008; Chambeyron et al., 2008). Many factors involved in the production of piRNAs localize to germ granules, suggesting that they may function as a site for processing of germline piRNAs (Harris and Macdonald, 2001;Vagin et al., 2006; Brennecke et al., 2007; Gunawardane et al., 2007; Lim and Kai, 2007; Pane et al., 2007; Klattenhoff and Theurkauf, 2008; Aravin et al., 2009; Li C et al., 2009; Malone and Hannon, 2009; Patil and Kai, 2010). The piRNA pathway is active in the follicle cells, nurse cells, and the egg cell of the Drosophila egg chamber (Lin and Yin, 2008; Li C et al., 2009; Malone et al., 2009). Nurse cells are connected to the growing egg cell by cytoplasmic bridges and, throughout development, the nurse cells deposit their cytoplasmic contents into the egg cell. Active TE transcripts in the nurse and egg cells are processed into piRNAs through the activity of Piwi, Aubergine (Aub), and Drosophila melanogaster Argonaute 3 (dmAgo3). Upon fertilization, these TE-derived piRNAs are maternally inherited in the embryo (McCue and Slotkin, 2012).

In mice, two Piwi family members, MILI and MIWI2, also specify de novo DNA methylation of transposon sequences in embryonic germ cells, leading to their transcriptional repression (Aravin et al., 2007b; Carmell et al., 2007; Kuramochi-Miyagawa et al., 2008). In MIWI-null male mice the nucleolar signal from miRNA and piRNA probes in Sertoli cells is largely diminished (Marcon et al., 2008), implicating Sertoli cell-specific miRNA and piRNA biogenesis as essential for spermatogenesis (Papaioannou et al., 2009; 2011; Kim GJ et al., 2010; Wu et al., 2012; Ortogero et al., 2013). Careful analysis of piRNAs that associate with different PIWI proteins has revealed at least two distinguishable classes of piRNAs in mammalian testis which are named by their expression time as pre-pachytene and pachytene piRNAs (Meikar et al., 2013). Pre-pachytene piRNAs are a strikingly uniform subclass of piRNAs that originate from repeat sequences related to transposable elements and heterochromatic regions. They are now the most studied and understood piRNAs, although they represent only a tiny fraction of all mammalian piRNAs. MILI and MIWI2 together with pre-pachytene piRNAs participate in silencing of transposable elements both at epigenetic and posttranscriptional level in fetal and neonatal germ cells (Aravin et al., 2007b; 2008; Carmell et al., 2007; Kuramochi-Miyagawa et al., 2008). In knock-out mice of either of these proteins the transposons are uncontrollably expressed, causing damage in the genome integrity of the cell, which eventually leads to meiotic arrest and sterility (Kuramochi-Miyagawa et al., 2004; Carmell et al., 2007). Pachytene piRNAs arise in spermatocytes during meiosis, peak in haploid round spermatids and disappear during later steps of spermiogenesis, overlapping the expression of their binding PIWI partners, MILI and MIWI (Meikar et al., 2013). The amount of pachytene piRNAs per each pachytene spermatocyte or round spermatid is remarkable, around a million molecules, but the individual copy number is low, which means that this piRNA population is very heterogeneous consisting of hundreds of thousands of different piRNAs (Aravin et al., 2006; 2007a). Compared to their pre-pachytene counterparts, the biogenesis and function of pachytene piRNAs is unknown. Pachytene piRNAs are devoid of sequences relative to active transposons (Meikar et al., 2013). Instead they map into large sparse clusters, from tens to hundreds of kilobases along the genome with most of the clusters being derived from one of the two genomic strands (Aravin et al., 2007a; b). Interestingly, pachytene piRNAs in round spermatid are concentrated in chromatoid bodies – male germ cell-specific germ granules of remarkable size and peculiar features (Kotaja and Sassone-Corsi, 2007; Meikar et al., 2010; 2011). As the chromatoid body also concentrates MIWI, RNA binding proteins and helicases in addition to longer polyadenylated RNAs (Kotaja et al., 2006; Meikar et al., 2010), it has been speculated that it serves as a processing centre for pachytene piRNAs and mRNAs. Besides the well-known function of transposon silencing, piRNAs function at various levels in regulating germline stem cell differentiation, mitotic chromosome dynamics, and gene expression (Pek et al., 2012b; Mani and Juliano, 2013). Most piRNAs in mouse spermatocytes do not match transposable element sequences. There is very little conservation of individual piRNA sequences between different mammals, but surprisingly there is a significant conservation of the genomic locations of mammalian piRNA clusters (Betel et al., 2007). It is therefore plausible that many piRNAs in the mouse act through entirely different mechanisms or regulate different biological functions altogether (Stefani and Slack, 2008). It has been proposed (Mani and Juliano, 2013) that the PIWI/piRNA pathway controls genome stability in several ways: suppression of transposons, direct regulation of chromatin architecture and regulation of genes that control important biological processes related to genome stability. An intriguing possibility is that the PIWI/piRNA pathway is using transposon sequences to coordinate the expression of large groups of genes to regulate cellular function (Mani and Juliano, 2013).

RNAs are intercellular signaling molecules (Dinger et al., 2008; Mittelbrunn and Sánchez-Madrid, 2012). When eukaryotic cells encounter double-stranded RNA (dsRNA), genes carrying a matching sequence are silenced through RNAi. RNAs are known to enter cultivated mammalian cells by natural processes, including direct cell-to-cell contact, membrane receptors, and channels (Winston et al., 2002; 2007; Feinberg and Hunter, 2003; Valiunas et al., 2005; Shih and Hunter, 2011; McEwan et al., 2012). mRNA required for Drosophila oocyte specification and growth is transcribed in neighboring nurse cells and trafficked through connecting ring canals to the oocyte in the egg chamber (Steinhauer and Kalderon, 2006). Similarly, RNA within chromatid bodies is transferred between germ cells via cytoplasmic bridges during mouse spermatogenesis (Ventela et al., 2003). In plants, mobile sRNAs promote epigenetic modifications in the genome of recipient cells (Molnar et al., 2010). When the recipient cells are seed or pollen, mobile sRNAs induce transgenerational epigenetic changes to enhance adaptation of progeny to future stresses (Slotkin et al., 2009). Soma-to-germline transfer of information mediated by small RNAs and, more generally, systemic RNAi is gaining experimental support in plants (Chitwood and Timmermans 2010) and protozoans (Gao and Liu, 2012; Schoeberl et al., 2012). There is also emerging evidence for trans-generational transfer of epigenetic information in other systems, including Metazoa (Ashe et al., 2012; Gu et al., 2012; Lee et al., 2012; Shirayama et al., 2012).

In C. elegans, gene silencing triggered by injected, ingested, or locally expressed dsRNA can move throughout the organism to reduce endogenous gene expression in all non-neuronal cells, including the germline, thus transmitting gene silencing to the next generation (Fire et al., 1998; Tabara et al., 1998; Timmons and Fire, 1998; Winston et al., 2002; Jose and Hunter, 2007; Whangbo and Hunter, 2008; Jose et al., 2011). In fact, the multigenerational inheritance of RNAi-mediated phenotypes delivered to somatic cells in Caenorhabditis elegans demonstrated that soma and germline do communicate (Burton et al., 2011). Key components required for systemic RNAi in C. elegans are the broadly expressed transmembrane proteins, systemic RNA interference defective protein (SID) (Winston et al., 2002; 2007; Feinberg and Hunter, 2003; Shih and Hunter, 2011; McEwan et al., 2012). SID homologs are present in a variety of other invertebrates and in all sequenced vertebrate genomes, indicating an ancient origin and likely conserved function (Duxbury et al., 2005; Li H et al., 2006; Wolfrum et al., 2007; Dinger et al., 2008; Tomoyasu et al., 2008; Xu and Han, 2008; Huvenne and Smagghe, 2010; Ren et al., 2011). One exciting finding to emerge from genetic screens of C. elegans was the discovery of three genes, RNAi spreading defective (rsd ), rsd-2, rsd-3, and rsd-6, required for the transfer of dsRNA into germline cells (Tijsterman et al., 2004). The protein encoded by rsd-3, which has both mouse and human orthologs (Wasiak et al., 2002), contains an ENTH (epsin NH2-terminal homology) domain that is commonly found in proteins involved in vesicle trafficking (Legendre-Guillemin et al., 2004). However, rsd mutants do not exhibit generalized defects in endocytosis. These observations suggest that RSD-3 may control vesicle trafficking in a pathway that is specific for the import of the silencing signal into the germline (Tijsterman et al., 2004; Jose and Hunter, 2007) providing a means by which environmental factors may lead to heritable changes. Vastenhouw and colleagues (2006) found that feeding C. elegans with bacteria expressing double-stranded RNA that targets specific nematode genes led to morphological and physiological variations that were transmitted for at least 10 generations (Vastenhouw et al., 2006; Jablonka and Lamb, 2008a). Clear evidence of transgenerational epigenetic inheritance in mice involving RNA has recently been provided (Rassoulzadegan et al., 2006; Ashe and Whitelaw, 2007; Li and Lu, 2009). Thus RNA may supplement endocrine and paracrine signaling by small molecules and proteins, and act as an efficient and evolutionarily flexible source of sequence-specific information transfer between cells, both locally and systemically andmay play a key role in transgenerational epigenetic information transmission (Dinger et al., 2008).

Another factor worth considering is the half-life of a potential transgenerational factor. Most of the components of the RNA silencing pathways, such as siRNAs and miRNAs, have a rapid turnover with half-lifes in the range of a couple of hours (Chatterjee and Großhans, 2009; Krol et al., 2010; Rüegger and Großhans, 2012). However, an important feature of RNAs in stress granules is that their translation is suppressed and RNA stability is increased (Anderson and Kedersha, 2002a; b; Kedersha and Anderson, 2002; Stöhr et al., 2006; Barckmann and Simonelig, 2013) that should protect transcripts from decay. The RNA-binding protein HuR (also known as ELAV1) binds to the 3’-untranslated region of mRNAs and regulates transcript stability and translation. Misregulation of HuR, due to expression of a HuR transgene, prevents the production of fully functional gametes in mice, providing evidence for the importance of mRNA stability in gametogenesis (Levadoux-Martin et al., 2003). 2'-O-methylation on the 3' terminal ribose is a major mechanism that increases the stability of small RNAs (Ji and Chen, 2012). The small RNA methyltransferase HUA ENHANCER1 (HEN1) and its homologs methylate microRNAs and small interfering RNAs (siRNAs) in plants, Piwi-interacting RNAs (piRNAs) in animals, and siRNAs in Drosophila. piRNAs and Drosophila Ago2-associated siRNAs are 2'-O-methylated at their 3' endby HEN1 homologs in the animal kingdom (Horwich et al., 2007; Kirino and Mourelatos, 2007; Saito et al., 2007; Kamminga et al., 2010). HEN1 mutants have a variety of gametogenic defects in various plants and animals (Ji and Chen, 2012). Most animal Hen1 proteins have a similar accumulation pattern as the Piwi proteins (Kirino and Mourelatos, 2007; Saito et al., 2007; Kamminga et al., 2010) and localize to nuage (Kamminga et al., 2010; 2012; Billi et al., 2012; Scott and Norbury, 2013).

As mentioned earlier, the germline can form (i) early in embryogenesis from an inheritance of maternal factors (maternally derived, also referred to as preformation) as used in flies and nematodes, (ii) by cell-cell interactions early in embryogenesis (inductive, also referred to as epigenetic) as seen in mice, and (iii) any time in the animal’s life, even in adulthood, from a multipotent stem cell precursor (persistent multipotent cell-derived germ cells), such as in planaria and hydra (a mode of germ cell specification pathway that also occurs in plants) (Gustafson and Wessel, 2010). In their review on this topic Extavour and Akam (2003) differentiated only between two modes of germline specification: preformation and epigenesis. Mode (iii) in the scheme of Gustafson and Wessel (2010) was grouped with the epigenesis mode (Extavour and Akam, 2003). Mode (iii), occurring in basal Metazoa such as Porifera (sponges) and Cnidaria (corals, jellyfish, hydra), planarians, ascidians, and plants is thought to be the phylogenetically oldest mode of germline specification (Extavour, 2007). In these phyla, germ cells obviously are not preformed but arise from a somatic stem cell population (e.g. archaeocytes, interstitial cells, meristem cells) that also generates other cell types (Müller, 2006; Watanabe et al., 2009). Importantly, these stem cells also play a key role in the almost unlimited regenerative potential of these taxa. In hydrozoan Cnidarians, pluripotent interstitial cells (i-cells) contain electron-dense cytoplasmic bodies similar to those associated with germ cells in all phyla (Eddy, 1975). In the ctenophore Pleurobrachia, expression of Piwi, Vasa and PL10 genes occurs in both the germ line and in a variety of non-germ line stem cells (Alié et al., 2011). Likewise, in the hydrozoan cnidarian Clytia hemisphaerica the multipotent i-cells in larvae and adult medusae, from which germ cells derive, express a set of conserved germ cell markers: Vasa, Nanos1, Piwi and PL10 (Leclère et al., 2012). In situ hybridization analyses revealed maternal mRNAs for all these genes highly concentrated in a germ plasm-like region at the egg animal pole and inherited by the i-cell lineage, strongly suggesting i-cell fate determination by inheritance of animal-localized factors. On the other hand, experimental tests showed that i-cells can form by epigenetic mechanisms in Clytia, since larvae derived from both animal and vegetal blastomeres separated during cleavage stages developed equivalent i-cell populations (Leclère et al., 2012). Thus Clytia embryos appear to have maternal germ plasm inherited by i-cells but also the potential to form these cells by zygotic induction. Reassessment of available data indicates thatmaternally localized germ plasm molecular components were plausibly present in the common cnidarian/bilaterian ancestor, but that their role may not have been strictly deterministic (Leclère et al., 2012). These bodies become more numerous in i-cells that develop into germ cells, and decrease in number in i-cells that differentiate into nematocytes (Extavour and Akam, 2003).

If germ granules are the vehicles of transgenerational information transfer the differential time course of their formation may tell us something about the mode of information transfer. The situation appears clear-cut for preformation. In animals where the germline is preformed, germ granules are present continuously in germ cells (with the exception of mature sperm) and are inherited maternally with the germ plasm (Seydoux and Braun, 2006). Oocytes carry transgenerational messages in the cytoplasmatic ribonucleoproteins of the germ granules. The soma-germline information transfer and, hence, the formation of the germ granules is completed with the release of the gametes/offspring into the environment. Epigenesis implicates that germ granules are not observed before, or immediately following, fertilization but are induced by cell-cell interactions during early embryogenesis. However, epigenesis is anumbrella term for various modes of formation. In basal Metazoa transgenerational information transfer appears to occur via maternal germ plasm that is only detectable by in situ hybridization analysesand that is distributed among the somatic stem cells (Alié et al., 2011; Leclère et al., 2012). In vertebrates, epigenesis occurs in urodele amphibians and mammals. In both, experimental manipulations can induce re-specification of cells from various embryonic regions to PGC fates, and there is no detectable mRNA or protein localization for the germ plasm “markers”, or localization of electron-dense granules during early development (Extavour and Akam, 2003).

The regeneration of body tissues and organs is a widespread phenomenon among urodele amphibians (Tanaka, 2003). The Mexican axolotl (Ambystoma mexicanum) has an unparalleled regenerative capacity among vertebrates (McCusker and Gardiner, 2010). Among various vertebrate tissue/organ regeneration models in axolotls, limb regeneration is the most intensively studied (Brockes and Kumar, 2005). Cells at the site of amputation form a blastema, a mass of undifferentiated cells. Both lineage-committed progenitor stem cells and lineage-uncommitted pluripotent stem cells appear to be present throughout the individual as reserve populations of stem cells. During tissue replacement these quiescent stem cells become activated, proliferate, and differentiate into the missing tissues (Young, 2004). Recent evidence suggests that the regenerating blastema may acquire a germline-like state (Zhu et al., 2012a; b). The DAZ-like gene is essential for meiosis or gametogenesis in several systems and is a molecular marker for germ cells (Johnson et al., 2001). Importantly, maternal dazl RNA is inherited in the animal cap and Axdazl RNA is found quite equally distributed in all regions of the axolotl embryo with similar concentrations in the ventral and dorsal marginal zone (Johnson et al., 2001). However, the products of germ cell-specific genes, such as Dazl and vasa, are not localised in the oocytes and are not zygotically transcribed in PGCs until they approach the gonadal ridges (Johnson et al., 2001; Johnson et al., 2003). Thus, epigenetic information transfer in axolotls is reminiscent of the mode of information transfer in basal metazoan that, like axolotl, have a high regenerative potential.

Mammals are unique in that there is an intimate mother-embryo/fetuscommunication during gestation. The epigenetic formation of germ cells has been well characterized in mice and appears to be common to all placental mammals. The placenta has an important role in embryonal/fetal programming (Jansson and Powell, 2007; Gheorghe et al., 2010). Since this epigenetic programming may persist for several generations (Stewart et al., 1975; Aerts et al., 1990; Oh et al., 1991; Aerts and Van Assche, 1992; Zambrano et al., 2005; Burdge et al., 2007; Pinheiro et al., 2008; Dunn and Bale, 2011; see also chapter 16.2), this programming can be expected to include also the germline cells. Between fertilization and implantation, both the maternal and paternal genomes lose their DNA methylation (although methylation of imprinted germline differentially methylated regions is faithfully maintained after fertilisation as a lifelong memory of parental origin of the allele in the new generation) (Smallwood and Kelsey, 2012). Concomitant with blastocyst implantation and cell-lineage determination, new methylation landscapes become established. PGCs emerge in mouse embryos at E7.5 and, concomitant with their proliferation and migration towards the genital ridge, DNA methylation is globally erased. Following sex-determination, new DNA-methylation landscapes are established in germ-cell precursors in an asymmetrical fashion in male and female embryos. In the male embryo, de novo methylation takes place before meiosis in mitotically arrested cells and is completed before birth. In the female embryo, DNA methylation is established after birth during the follicular/oocyte growth phase and is completed at puberty (Smallwood and Kelsey, 2012). Importantly, nuage formation in mouse PGC parallels the loss of DNA methylation during migration towards the genital ridge (Eddy, 1974; 1975). Overall, the dynamics of germ granule formation and zygote/gamete reprogramming suggest a role of germ granules in transgenerational information transfer. Hence, the different time course and regulation of germline specification reflects the various reproductive systems ranging from recruitment of somatic stem cells in basal Metazoa to mammal fetal programming as parental start-up support in a capricious environment. The following scenario is proposed: (i) in a close cross-talk with germline cells, somatic gonadal cells determine the ncRNA profile in the germline cells; (ii) after intercellular transfer secondary piRNA biogenesis (ping-pong) amplifies the message in the germline; (iii) the ncRNA message is stored in the germ granules until the information is retrieved for embryonal development (preformation); (iv) in species with extended parent-embryo/fetus developmental support (e.g. in mammals), reprogramming by transgenerational message transfer is not confined to gametogenesis. In mammals, reprogramming occurs also during embryonal development (epigenesis) and may even extend to maturity as is suggested by the completion of oocyte DNA methylation pattern.

Although posttranscriptional mechanisms likely contribute to silencing, CpG methylation is critical for transposon repression in mammals (Yoder et al., 1997; Liang et al., 2002; Lippman et al., 2003; Bestor and Bourc'his, 2004; Bourc'his and Bestor, 2004; Gaudet et al., 2004; Aravin et al., 2007a). Loss of CpG methylation can be expected to weaken the control of TE mobilization. During reprogramming, covalent modification such as DNA methylation is replaced by non-covalent RNA-RNA and RNA-DNA base pairing interactions. Mostly, however, non-covalent interactions are considerably weaker (by one to three orders of magnitude) than covalent bonds (Hobza et al., 2006). Thus, in germ cells, pre- and postimplantation embryos, the host surveillance of TEs becomes less than optimal (Dupressoir and Heidmann, 1996; Loebel et al., 2004; Peaston et al., 2004; Svoboda et al., 2004; Taruscio and Mantovani, 2004; Maksakova et al., 2008; Leung and Lorincz, 2012). During both male and female gametogenesis, temporary relaxation of the epigenetic control of TEs during early germline development opens a risky window that can allow TEs to escape from host constraints and to propagate abundantly in the host genome (van der Heijen and Bortvin, 2009; Bao and Yan, 2012). Under the prevailing paradigm that it is the primary mission of the germline to faithfully transmit the genome to the next generation, the question has not been asked whether it could be of evolutionary significance that germline cells endanger their mission by unleashing the destructive potential of TE, stripping their DNA methylation during epigenetic reprogramming. Taking into account the huge evolutionary significance of TE mobility (see chapters 10.6 and 12.4) this evidence should be another opportunity to reconsider the paradigm of the immutable germline.

17. Stochasticity and selection, the organizing principles of evolution


Neither a purely reductionist approach nor a merely holistic perspective is sufficient to encompass the intrinsic nature of the system’s behavior.

Pahl-Wostl, 1993

Summary

There is an important difference between breeding (Darwin’s role model of evolution) and evolution itself: while in breeding the final goal is preset and constant, adaptation to varying biotic and abiotic environmental conditions is a moving target and selection can be highly fluctuating. Evolution is a cybernetic process that can be understood as a learning automaton with input and output channels. The output/outcome has been defined: selection. The input has been less rigorously defined. In learning automata, output variables allow to make inferences about the informational input. Variation is pervasive at every level of biological organization. Variation is the bet-hedging strategy that tries to cover all bases in an often unpredictable environment when it does not make sense to “put all your eggs into one basket”. Importantly, most revealing is that variation is created condition-dependently, when variation is most needed– in organisms under stress. Bet-hedging is the risk-spreading response to environmental uncertainty. Bet-hedging is the output variable that allows to infer that environmental uncertainty/stochasticity is the informational input in evolution. Sexual reproduction with its regulated genetic-variant generation is the proof of concept that (epi)genetic variation is no happenstance outcome but a highly regulated process and environmental stochasticity is its evolutionary “mastermind”. A dual system, characterized by stochasticity and selection is the organizing principle of evolution. Both are interdependent. Sexual reproduction is the ultimate bet-hedging enterprise and its evolutionary success is the selective signature of stochastic environments.

The Modern Synthesis built on Darwin's two major realizations: (i) that all living organisms are related to one another by common descent; (ii) that a primary explanation for the pattern of diversity of life—and especially for the obvious "fit" of organisms to their environments—is the process that he called natural selection. Importantly, he recognized the importance of variation for the action of selection (1859, chapters I, II and V). However, he had no idea how this variation arose: “our ignorance of the laws of variation is profound” (1859). In Darwin`s tradition, the Modern Synthesis understood selection as the only driving force in evolution. Genetic variation was the result of accidental mutations. Thus, variation was understood as the accidental outcome of error-prone replication. Darwin`s concept of evolution was molded by his role model of artificial selection. But there is an important difference between artificial selection and evolution. In breeding, artificial selection has the goal to improve a certain predefined trait, e.g. oil content in maize (Laurie et al., 2004), milk quantity and quality in dairy cattle breeding, a certain morphological trait in pigeons, or, in the laboratory, flight speed in Drosophila (Weber, 1996). The target is pre-defined by the breeder. Importantly, breeding is an iterative process (Hill and Caballero, 1992; Williams and Lenton, 2007) in which the ultimate goal, e.g. increased oil content in maize (Laurie et al., 2004), is reached after many generations, but the setting of the ultimate goaland thus the direction of selection remain constant (figure 2B). Here, variation is an often unwanted noise, at least when it does not serve the ultimate target of selection. The breeder has at least two functions: he determines the goal of the breeding operation and selects the individual for the next round of breeding. In evolution, the direction and selective regime are established by the environment (figure 2C). However, the target, adaptation to varying biotic and abiotic environmental conditions, is a moving target and selection can be highly fluctuating (Bell, 2010). In this chapter, I argue that our understanding of the “laws of variation” has a key role in our understanding of evolution.

17.1 The cybernetics of evolution

Cybernetics is the study of control systems (Wiener, 1948, Ashby, 1956). Cybernetic systems are systems with feedback. They are a special class of cause-and-effect (input-output) systems. Patten and Odum (1981) offered a minimalist definition that distinguished cybernetic systems from non-cybernetic systems by the presence of feedback control; in cybernetic systems, ‘‘input is determined, at least in part, by output’’. Evolution is a cybernetic process (Ashby, 1954; 1956; Schmalhausen, 1960; Corning, 2005). Cybernetic systems are characterized by input and output variables and it is essential to distinguish the one from the other (Ashby, 1956) (figure 1A). Darwin was vague in the meaning of his new concept of “Natural Selection,” using it interchangeably as one of the causes for evolutionary change and as the final outcome (= evolutionary change). (Since evolution is a continuous iterative process where the resultant population of the previous contest becomes the input population to the next contest, Darwin’s ambiguity was not completely unfounded.) But his clearest definition of natural selection (Darwin, 1859 p. 61: “I have called this principle, by which each slight variation, if useful, is preserved, by the term of Natural Selection, in order to mark its relation to man’s power of selection.”) is an outcome definition, not that of a cause (Bock, 2003).First, natural selection is a metaphor, an umbrella term that serves to label and characterize a vast array of specific factors with survival consequences. The generally accepted modern definition of natural selection is that it is an outcome (Fisher 1930; Endler, 1986; Bock, 2003; 2010; Reese, 2005), corresponding to an output in a cybernetic system. Arguably, the most fundamental impacton this outcome comes from the environment. Surprisingly, the input into the process of evolution has been less rigorously defined. That the environment informs the evolutionary agent on its state is anstatementthat is rather trivial. More meaningful for the interpretation of evolutionary processes is the recognition that environments have a high degree of uncertainty/stochasticity (Yoshimura and Clark, 1991; Lenormand et al., 2009) or capriciousness (Lewontin, 1966).To deal with this uncertainty organisms had to acquire the capacity to learn from past environments to generalize to new environments (Kirschner and Gerhart, 2005; Gerhart and Kirschner, 2007; Parter et al., 2008). However, evolution does not "know" in advance which evolutionary path will lead to the increase of fitness or how fluctuating, often unpredictable, environments will change. Therefore, prospectively, the best "strategy" to increase fitness is to take every possible path at every next step. As a result, no configurations should be missed (Fu, 2007). Which configuration is a “fit” one, is finally decided by the survival and reproductive success of the individual. Both an unpredictable, fluctuating abiotic environment and constantly coevolving web of life (Thompson, 2005; 2009) contribute to the stochasticity. There are winners and losers in this game. Evolution is a stochastic process (Lenormand et al., 2009; Kupiec et al. 2012). But what makes evolution a stochastic process? Is it the inherent error susceptibility of evolutionary processes or is it the bet-hedging in response to some unpredictable, stochastic event?

To address this question let us consider evolution as a cybernetic, learning box. Learning automata are adaptive decision-making devices operating on unknown random environments (Narendra and Thathachar, 1974; 1989). The automaton updates its action probabilities in accordance with the inputs received from the environment so as to improve its performance in some specified sense (Narendra and Thathachar, 1974; 1989). The basic operation carried out by a learning automaton is the updating of the action probabilities on the basis of the responses of the environment. The learning automaton has a finite set of actions and each action has a certain probability (unknown to the automaton) of getting rewarded by the controlled system, which is considered as environment of the automaton. The aim is to learn to choose the optimal action (i.e. the action with the highest probability of being rewarded) through repeated interaction on the system. If the learning algorithm is chosen properly, then the iterative process of interacting on the system can be made to result in the selection of the optimal action (Zeng et al., 2000). The learning model that is closest to the evolutionary approach is ‘‘reinforcement learning’’ based on the insight that successful strategies will be reinforced and used more frequently. Reinforcement learning has been successfully applied for solving problems involving decision making under uncertainty (Narendra and Thathachar, 1989; Barto et al., 1983; Zikidis and Vasilakos, 1996; Zeng et al., 2000). (When speaking of ‘decisions’, use of the term is in an evolutionary sense, not implying any conscious rationalization on the part of individual organisms.) In general, a reinforcement learning algorithm conducts a stochastic search of the output space, using only an approximative indication of the “correctness” (reward) of the output value it produced in every iteration. Based on this indication, a reinforcement learning algorithm generates, in each iteration, an error signal giving the difference between the actual and correct response and the adaptive element uses this error signal to update its parameters (Zeng et al., 2000).

By looking at its output we should be able to make inferences about the quality of the informational input to this learning automaton. Clearly, the fact that the black box generates winners and loserssuggests that some type of lottery unfolds, that stochastic processes play a role within the learning box. The outcome of any evolutionary process is not a single result; it is at best a probability distribution of possible outcomes (Proulx and Adler, 2010). Hence, evolution can be described by a lottery model (Chesson and Warner, 1981; Proulx and Day, 2001; Svardal et al., 2011). On the other hand, the descents of the winners of this lottery (that were fit enough to reproduce) again are araffle ticket in another round of this iterative evolutionary game. During the iterative process the direction of selection can fluctuate, often unpredictable (see figure 2).

17.2 Bet-hedging as response to stochasticity

In a review of published studies on variation in recruitment, Hairston et al. (1996) found that reproductive success of long-lived adults varied from year to year by factors up to 333 in forest perennial plants, 4 in desert perennial plants, 591 in marine invertebrates, 706 in freshwater fish, 38 in terrestrial vertebrates, and 2200 in birds. These figures represent the variation among years when some reproduction occurred; many of the studies also report years in which reproduction failed completely. Similarly, the recruitment success of diapausing seeds or eggs varied by factors of up to 1150 in chalk grassland annual and biennial plants, 614 in chapparal perennials, 1150 in freshwater zooplankton, and 31,600 in insects (Ellner, 1997). In response to uncertainty as to which phenotype will have highest fitness in the future, biological systems exert risk minimization by risk-spreading. Bet-hedging is a response to environmental uncertainty (Einum and Fleming, 2004; Marshall et al., 2008; Beaumont et al., 2009; Crean and Marshall, 2009; Gourbière and Menu, 2009; Olofsson et al., 2009; Rajon et al., 2009; Simons, 2009; Nevoux et al., 2010; Simons, 2011; Morrongiello et al., 2012) and risk minimization strategy on all levels of biological organization (Cohen, 1966; Gillespie, 1974a; Slatkin, 1974; Tonegawa, 1983; Hairston and Munns, 1984; Seger and Brockmann, 1987; Philippi and Seger, 1989; Frank and Slatkin, 1990; Moxon et al., 1994; Sasaki and Ellner, 1995; Ellner, 1997; Danforth, 1999; Hopper, 1999; Menu et al., 2000; Lips, 2001; Meyers and Bull, 2002; Fox and Rauter, 2003; Friedenberg, 2003; Balaban et al., 2004; Einum and Fleming, 2004King and Masel, 2007; Venable, 2007; Acar et al., 2008; Ackermann et al., 2008; Gourbière and Menu, 2009; Olofsson et al., 2009; Simons, 2009; 2011; Childs et al., 2010;Monro et al., 2010; Gremer et al., 2012; Morrongiello et al., 2012; Starrfelt and Kokko, 2012).Bet-hedging may involve the production of fewer and larger offspring (conservative) or of variable-sized offspring (diversified). As the environment becomes more stable, the bet-hedging strategies have a lower fitness advantage. Metazoan bet-hedging usually involves phenotypic diversification among an individual's offspring, such as differences in seed dormancy. Almost all known microbial bet-hedging strategies in contrast, rely on low-probability stochastic switching of a heritable phenotype by individual cells in a clonal group (Stumpf et al., 2002; Thattai and van Oudenaarden, 2004; Kussell et al., 2005; Kussell and Leibler, 2005; Wolf et al., 2005; Avery, 2006; Veenning et al., 2008a; b; Beaumont et al., 2009; Fraser and Kaern, 2009; Ratcliff and Denison, 2010; Rainey et al., 2011; Levy et al., 2012).

The key point in Evolutionary Game Theory (EGT) models is that the success of a strategy is determined by how good the strategy is in the presence of other alternative strategies, and of the frequency that other strategies are employed within a competing population. To create a sufficient amount of winners under all realistic assumptions an evolutionary stable strategy (ESS) must ‘cover all bases’. Both theoretical and experimental approaches demonstrated that in the face of variable and unpredictable environments, bet-hedging is the ESS (Haccou and Iwasa, 1995; Sasaki and Ellner, 1995; Beaumont et al., 2009; Olofsson et al., 2009; Rees et al., 2010; Ripa et al., 2010; Starrfelt and Kokko, 2012). In constant environments natural selection leads to each individual organism maximizing its expected number of descendants left far in the future. If there are no environmental fluctuations, fitness can be measured by the arithmetic mean number of surviving descendants. In fluctuating environments it may be optimal for different individuals of the same genotype to take different actions to spread the risk and ensure the genotype is represented in future generations. The fitness of the genotype is determined by the, perhaps complementary, actions of all individuals of the genotype, and the best action for one individual depends on the actions of others (McNamara et al., 1995). The standard criterion for evaluating the fitness of genotypes in stochastic environments is the geometric mean of the growth rates (geometric mean fitness) (Dempster, 1955; Cohen, 1966; Lewontin and Cohen, 1969; Frank and Slatkin, 1990; Yoshimura and Clark, 1991; Yoshimura and Jansen, 1996; Hopper, 1999; Simons, 2009; Yoshimura et al., 2009). In unstable environments, the geometric mean is always lower than the arithmetic mean. A rigorous definition of bet-hedging includes lower expected arithmetic mean fitness, as well as greater expected geometric mean fitness (Seger and Brockmann, 1987; Simons, 2009). In fluctuating environments, when geometric mean fitness is maximized, individual optimization fails (Cohen, 1966; Ellner, 1986; McNamara 1995; 1998; McNamara et al., 1995; Yoshimura and Jansen, 1996). As egg phenotype is linked to offspring phenotype, increased within-brood variation in egg phenotype can have a selective advantage in unpredictable environments by increasing maternal geometric fitness (Marshall et al., 2008; Crean and Marshall, 2009; Crean et al., 2012). The best action of an individual depends on the states and actions of other population members (McNamara et al., 1995; McNamara, 1998; Török et al., 2004; Simons, 2009).Under the geometric mean criterion, behavior appears to be determined largely by a worst case scenario; behavior may appear suboptimal if observed only under normal or average conditions (Yoshimura and Clark, 1991; Yoshimura and Jansen, 1996). For example, except under extreme environmental conditions, mammalian litters (Murie and Dobson, 1987; Risch et al., 1995) and avian clutches (Perrins, 1965; Klomp, 1970; Murray, 1979; Lessells, 1986; Murphy and Haukioja, 1986; Boyce and Perrins, 1987; Vander Werf, 1992) larger than those that are observed in nature might result in increased fecundity, with little if any cost of reproduction in terms of parental survival. However, in unusually bad years such large clutches might be disastrous, in terms of parental survival (Yoshimura and Clark, 1991;Yoshimura and Shields, 1992). An illustrative example was given by Philippi and Seger (1989): “Suppose that years are ‘good’ or ‘bad’ with equal probability, and that the wild type produces, on average, 9 offspring in good years and 1 offspring in bad years, for an average of 5. Now introduce a mutant that produces 5 offspring in good years and 3 offspring in bad years, for an average of only 4. Despite its lower mean fitness, the mutant quickly goes to fixation because its geometric mean fitness (3.87) is much higher than that of the wild type (3.0) and its variance lower. The mutant’s best performance is much worse than the wild type’s best, but its worst is better, and this is the key to its success.” Variance reduction can be accomplished in two ways, risk avoidance and risk spreading (den Boer, 1968). Risk-avoidance strategies reduce variance by producing relatively constant outputs under different environmental conditions. Risk-spreading strategies, on the other hand, reduce population-level variance by some process of averaging over independent events.

According to Lynch (2007a; b) out of the four major forces in evolution, natural selection, mutation, recombination and drift, the latter three are stochastic in nature (however, in chapter 12, I discussed that at least mutation and recombination are non-random). In addition to these stochastic genetic mechanisms,there is growing evidence of evolutionary selection for stochastic diversity-generating mechanisms in unicellular and multicellular organismson a variety of epigenetic, developmental, physiological and behavioral levels (True and Lindquist, 2000; Elowitz et al., 2001; Fraser et al., 2004; Raser and O’Shea, 2004; 2005; Kærn et al., 2005; Kussell and Leibler, 2005; Avery, 2006; Peaston and Whitelaw, 2006; Smits et al., 2006; Lim and van Oudenaarden, 2007; Maamar et al., 2007; Acar et al., 2008; Davidson and Surette, 2008; Freed et al., 2008; Losick and Desplan, 2008; Shahrezaei and Swain, 2008; Lenormand et al., 2009; Dercole et al., 2010; Lidstrom and Konopka, 2010; Huang, 2012).Comparing RNA sequences from human B cells of 27 individuals to the corresponding DNA sequences from the same individuals, Li et al. (2011) uncovered more than 10,000 exonic sites where the RNA sequences did not match that of the DNA, revealing infidelity of information transmission from DNA to RNA as an additional aspect of genome variation. The number of events varied among individuals by up to sixfold across 27 subjects (Li et al., 2011).

The question whether all these variation-generating processes are accidental or are evolutionarily intended amounts to the question whether bet-hedging is a haphazard process or an ESS. Bet-hedging can be represented by an evolutionary game (Olofsson et al., 2009). Variation is the bet-hedging strategy to cover all bases in an often unpredictable environment. It does not make sense to “put all your eggs into one basket”. In bet-hedging, spreading of the bets is a must to improve one’s chances to win. Examples of risk spreading include dispersal of progeny (spatial averaging: Levin et al., 1984), iteroparity (Murphy, 1968; Bulmer, 1985; Orzack and Tuljapurkar, 1989; Wilbur and Rudolf, 2006), delayed germination of seeds (temporal averaging: Ellner, 1985) and phenotypic polymorphism (Levins, 1968; Roughgarden, 1979, p. 272). Importantly, most revealing is that variation is created condition-dependently, when variation is most needed– in organisms under stress. Thus stress elicits increased mutagenesis, increased epimutagenesis, increased recombination, increased transposon mobility, increased repeat instability, increased phenotypic plasticity, and, in organisms that can reproduce both asexually and sexually, increased sexual reproduction. Several theoretical models confirm that sexual reproduction is selected for in variable environments (Hines and Moore, 1981; Weinshall, 1986; Roughgarden, 1991; Robson et al., 1999). Sexual reproduction is the ultimate bet-hedging enterprise and its evolutionary success the selective signature of stochastic environments.

Thus environmental stochasticity elicits bet-hedging as risk-spreading response resulting in (epi)genetic, developmental, phenotypic, physiological and behavioral variation on which selection can act (figure 2C).

17.3 Stochasticity and selection: dualism in evolution

This work has provided compelling evidence that a change of environmental conditions that reduce Darwinian fitness (see the ecological stress definition chapter 2) may elicit

  1. stress-induced mutagenesis
  2. stress-induced epimutagenesis
  3. increased recombination rate
  4. increase mutability of simple sequence repeats
  5. increased mobilization of transposable elements

all of which, when acting on the germline, increase heritable (epi)genetic variation. Sexual reproduction regulates these processes and, by changing the SMSC balance, is the process modulating the (epi)genetic variation-selection balance, e.g. by the action of hormones (see chapter 14.2). Theoretical models suggests that fluctuating selection may be an important factor in maintaining genetic polymorphism (Korol et al., 1996, Kirzhner et al., 1998; Bürger and Gimelfarb, 2002). Likewise, empirical studies of cyclical and fluctuating selection suggest an association between temporal environmental heterogeneity and the amount of genetic variation (Korol et al., 1996; Kondrashov and Yampolsky, 1996). These processes and their quasi-stochastic generation of variation appear to have an evolutionary rationale: fighting change with change (Meyers and Bull, 2002), risk-spreading in the face of environmental uncertainty, creating lottery tickets for the raffle of life. On the other hand, sexual reproduction as evolutionarily highly successful strategy, highlights an eminent characteristic of evolution: it pays off to diversify and be prepared for the unlikely event. And: generation of variation is no happenstance outcome but a highly regulated process and environmental stochasticity is its evolutionary “mastermind”.

According to Mayr (1980), selection is ‘‘the only direction-giving factor in evolution’’. On the other hand, Monod (1971) argued that chance is both the major creative force and a crucial problem for evolving biological systems. The paradigm of calculability, determinism and monocausality dominated the sciences until the beginning of the 20th century. Since the end of the 19th century, monocausal approaches in many different sciences started to collapse. Even in pure mathematics and logics, problems with the calculability of the universe arose (e.g. Russell´s paradox). Hilberts program failed with Kurt Gödel´s proof. At the level of physics, many different problems (ultraviolet catastrophe, wave-particle dualism, ...) lead to the development of new physics (Brunner and Klauninger, 2003). Like the wave paradigm could not explain a variety of physical properties of light, explaining evolution by natural selection as only organizing principle creates various implausibilities. As it stands, it is accepted that it makes sense to use stochastic models in population genetics. But why should a selection-only process be stochastic? It is agreed that natural selection has its limits (Barton and Partridge, 2000). But so far these limits have been explained by e.g. genetic architecture, historical contingency or developmental constraints.

Understanding evolution as cybernetic box and learning automaton identifies the informational input into the evolutionary self-corrective system: environmental stochasticity. Darwin already realized that variation is an essential commodity in evolution but he was unaware of its cause. The Modern Synthesis regarded variation as the result of accident, happenstance and imperfection. It is textbook knowledge that selection needs variation to work on. The central question, however, is whether variation is the result of accident and chance or whether it evolved as a means to create many lottery tickets in response to the unpredictability of life. The recognition that variation arises at all levels such as the genetic, epigenetic, developmental, physiological, behavioral and life-history level, that it is malleable in response to stress (when it is most needed) and that sexual reproduction evolved as master tool creating pre-selected variation, clearly points to the latter interpretation. And promoting stochasticity to the boardroom of evolution opens a new dimension of insight into a variety of evolutionary phenomena and enigmas (see chapter 19).

In the dualism of stochasticity and selection, variation is recognized as the result of a multitude of processes, resulting in a bet-hedging response to stochasticity. In essence, stochasticity and selection work against each other within the limits of total chaos and complete order, the two extremes where evolution is no longer feasible. On the other hand, stochasticity and selection are interdependent. None can prevail without depriving evolution of its very basis. Selection could not work without the stochastic phenomenon of variation; and stochasticity needs the ordering power of selection to create the complex structures of self-organized criticality (Bak et al., 1987; 1988). Intriguingly, part of the stochasticity is created by selection itself, e.g. through coevolutionary cycles, density- and frequency-dependent selection, or niche construction (Meyers and Bull, 2002). On the other hand, stochasticity drives variation and variation is the raw material for selection to work on. Although within wide boundaries, stochasticity and selection have to be balanced. Evolutionary biology already acknowledged mutation-selection equilibrium as evolutionary phenomenon; it is time to realize that there is a stochasticity-selection balance. Too much stochasticity would be detrimental for learning: if the cybernetic feedback concerning fitness effects would not behave with a certain stability and change too irregularly, learning would be impaired. Fortunately, with respect to living organisms, nature is capricious rather than completely random (Lewontin, 1961; 1966). There is a variable degree of ecological predictability: demographic cycles due to e.g. predator/prey interactions, seasons with their cyclicity of resource availability, circadian cycles, tides, etc. Bet-hedging only makes sense in a more or less stochastic environment. As the environment becomes more stable, bet-hedging strategies have a lower fitness advantage (Philippi and Seger, 1989). Stochasticity is ambiguous (e.g. beneficial, neutral and deleterious mutations) with regard to outcome while selection filters and directs the ambiguity. And learning attenuates the randomness. Selection is the stabilizing force that brings order into the chaos and provides the feedback for learning to occur. Both stochasticity and selection render evolution opportunistic.

There is compelling evidence, of which this paper only could sample a small amount, that whenever organisms and cells are in stress, i.e. are maladapted and in need of genetic novelty, they upregulate mutagenesis and hedge their bets in the face of environmental stochasticity. Modifying Dobzhansky’s notorious quote, Lynch (2007a) wrote: “Nothing in evolution makes sense except in light of population genetics”.However, in 1961 Lewontin did not consider population genetics an “adequate theory of evolutionary dynamics. On the contrary, the theory of population genetics, as complete as it may be in itself, fails to deal with many problems of primary importance for an understanding of evolution.” In this paper, Lewontin (1961) suggested that the modern theory of games (von Neumann and Morgenstern, 1944; 1953) may be useful in finding exact answers to problems of evolution not covered by the theory of population genetics. A first application of game theory to evolutionary issues was the work of Maynard Smith and Price (1973) on animal conflicts and their concept of an “evolutionarily stable strategy” (ESS).The vast body of theoretical work based on the concept of an ESS, often disregards environmental stochasticity; for example, Maynard Smith's often quoted book (Maynard Smith, 1982) contains no reference to stochasticity. An important feature of evolutionary game theory (EGT) models is repetition. If the games were not repeated, these EGT models would not be able to provide any insight into adaptive behaviors and strategies due to the dynamic nature of the mechanisms of evolution. Importantly, evolution “plays” both within-generation and trans-generation games. At each game repetition population make-up in turn is determined by the results of all of the previous contests before the present contest- it is a continuous iterative process where the resultant population of the previous contest becomes the input population to the next contest.Evolution is a stochastic process (Lenormand et al., 2009; Kupiec et al. 2012) that can be described by a lottery model (Chesson and Warner, 1981; Proulx and Day, 2001; Svardal et al., 2011). Bet-hedging is anESS in the face of environmental capriciousness (Lewontin, 1966) and sexual reproduction with its mutagenesis-selection cascades, is the evolutionary commander in chief, “master-minded” by evolutionary unpredictability.

The identification of stochasticity/uncertainty as input variable into the cybernetic machine of evolution (Ashby, 1956) has far-reaching implications for a multitude of evolutionary enigmas as discussed in chapter 19.

18. Earlier theories that attempted to explain why most organisms reproduce sexually


…that an opinion has been widely held is no evidence whatever that it is not utterly absurd….

Bertran Russell (1929)

Summary

What all current theories have in common is their focus on recombination as the only perceived mechanism in sexual reproduction. Thus they had to fail to decipher the evolutionary rationale of sexual reproduction. Deeply entrenched in population genetics thinking, the so-called genetic theories of sexual reproduction (Fisher-Muller-model, Muller’s ratchet and Kondrashov’s hatchet) completely missed the dynamic aspect of evolution. Apart from highly deleterious mutations that endanger viability, whether a mutation is beneficial, neutral or slightly deleterious is context-dependent and highly volatile. Evolution is opportunistic and to take every possible path at every next step, no viable configurations should be missed. Homologous recombination may have been used for DNA repair in microbes but as is often in evolution was later co-opted for innovation rather than conservation. That sexual reproduction evolved in hosts to counter the pressure of parasitic coevolution as the “Red Queen” hypothesis states, reflects a too narrow scope of ecological challenges in a constantly coevolving, highly dynamic web of life.

Up to now, the general perception with regard to sexual reproduction suffered from a too narrow perspective on the mechanism of recombination (e.g. Agrawal, 2006b). Consequently, theories of the evolution of sex were very much like searching for the lost key under the lamppost. Missing the most “ingenious” innovation of sexual reproduction, preselected (epi)mutagenesis, these theories had to fail.

18.1 Fisher-Muller-model, Muller’s ratchet and Kondrashov’s hatchet

How can such a rich theoretical structure as population genetics fail so completely to cope with the body of fact? Are we simply missing some critical revolutionary insight that in a flash will make it all come right, as the Principle of Relativity did for the contradictory evidence on the propagation of light?

Lewontin (1974)

Sex and recombination have long been seen as adaptations that facilitate the work of natural selection. The so-called genetic theories of sexual reproduction try to explain the evolutionary success of sexual reproduction within the framework of population genetics.

1) The positive mutation models or Fisher-Muller hypothesis propose that sex, allowing for recombinationand segregation, is advantageous because beneficial mutations that arise in different individuals can be united in the same genome and therefore offspring with above average fitness can be produced (Fisher, 1930; Muller, 1932; Crow and Kimura, 1965, 1970; Manning, 1976; Bell, 1982; Wolf et al., 1987; Findlay and Rowe, 1990; Peck, 1993; 1994; Dunbrack et al., 1995; Otto and Barton, 1997; Peck et al., 1997). In fact, a number of specific evolutionary scenarios lead to the general conclusion that higher levels of adaptation can be achieved when recombination is frequent than in its absence, because of the associated reduction in the amount of Hill-Robertson effect (a reduction in the efficacy of natural selection that occurs because finite populations accumulate associations of linked genes that interfere with selection) (Gordo and Charlesworth, 2001; Otto and Lenormand, 2002; Marais and Charlesworth, 2003). However, when the only evolutionary process acting is viability selection (no mutation, no departures from random mating, no drift, etc.), evolutionary theory predicts that populations should evolve lower and lower rates of sex and recombination. The underlying reason is that sex and recombination break apart the favorable gene combinations that have been built up by past selection (recombination load), offsetting the benefits of combining advantageous mutations (Maynard Smith 1968, 1978; Feldman et al., 1980; Altenberg and Feldman, 1987; Charlesworth and Barton, 1996; Otto, 2007). The balance of these two forces determines whether modifiers that increase recombination are favored. Recombination load has been shown in Chlamydomonas (Colegrave et al., 2002b; Kaltz and Bell, 2002), rotifers (Becks and Agrawal, 2011), and in yeast (Greig et al., 1998). In a constant environment these genetic associations are typically favorable, and theoretical analyses have demonstrated that decreased levels of recombination selection evolve under such circumstances (Feldman, 1972; Feldman e al., 1980; Altenberg and Feldman, 1987; Feldman et al., 1997). Moreover, the theory does not take into account that interactions between pairs of beneficial mutations may be antagonistic (Sanjuán et al., 2004b; Kryazhimskiy et al., 2009; 2011; Chou et al., 2011; Khan et al., 2011) leading to decompensatory epistasis with the double mutant being less fit than each single mutant (Wolf JB et al., 2000). Although homologous recombination increases genetic diversity by breaking haplotypes, it may also homogenize alleles through gene conversion (Gordo and Charlesworth, 2001; Chen JM et al., 2007; Mancera et al., 2008). Thus, recombination randomizes genotypes and homogenizes sequences between the participating alleles without regard to the fitness of the alleles being recombined (Ferreira, 2002; Hunter, 2006; Keeney, 2007; Otto, 2009).

2) Another genetic explanation for the predominance of sexual reproduction, called Muller’s ratchet, points to the purging of deleterious mutations from the genome by recombination (Muller, 1964; Felsenstein, 1974; Manning and Dickson, 1986; Kondrashov, 1988; 1993;Kondrashov and Turelli, 1992; Zeyl and Bell, 1997). There is support for the action of Muller’s ratchet in experimental laboratory populations (Chao, 1990; Duarte et al., 1992; Andersson and Hughes, 1996; Zeyl et al., 2001), indirectly from studies on the long-term effects of reduced population sizes on genetic diversity and fitness in the wild (Westemeier et al., 1998; Johnson JA et al., 2003; Rowe and Beebee, 2003) and in mitochondrial genomes of animals evolving in nature (Howe and Denver, 2008; Neiman et al., 2010). In the homothallic fungus Aspergillus nidulans,Bruggeman et al. (2003) were able to show that sex slows down the accumulation of deleterious mutations. The question is, however, whether Muller’s ratchet may be the cause for the evolution and maintenance of sexual reproduction. These arguments require genomic mutation rates that are higher than typically observed (Keightley and Eyre-Walker, 1999; Lynch et al., 1999; Bataillon, 2000; Fry, 2004; Haag-Liautard et al., 2007). Most well characterized asexual lineages fail to exhibit the high levels of allelic divergence expected in the absence of recombination (Normark, 1999; Schön and Martens, 2003; Omilian et al., 2006; Schaefer et al., 2006). Moreover, as outlined by Melzer and Koeslag (1991), due to the model’s constraints imposed on the modeled asexual isolates, namely immortality and constancy in isolate size, Muller’s ratchet operates only when mutations affect the outcome of intraspecific contests but not the organisms’ intrinsic ability to survive in the ecosystem. Barton (1995) outlined the contributions of short- and long-term effects to the evolution of sex in a general theoretical framework. Both the faster combination of favorable mutations (Fisher, 1930; Muller, 1932) and the avoidance of the accumulation of deleterious mutations (Muller, 1964) are long-term advantages since they operate very slowly in large populations and can be avoided with very little sex (Pamilo et al., 1987; Charlesworth et al., 1993; Green and Noakes, 1995; Hurst and Peck, 1996; Chasnov, 2000; Keightley and Eyre-Walker, 2000; Bengtsson, 2003; Haddrill et al., 2007; D'Souza and Michiels, 2010; Lampert and Schartl, 2010). Therefore, both processes have been deemed unlikely to explain the short-term maintenance of obligate sexuality. Moreover, even weak purifying selection may eventually lead to the complete cessation of the ratchet (Fontanari et al., 2003; Maia, 2009; Goyal et al., 2012; Parmakelis et al., 2013). Evidently, the long-term stability of an asexual population requires an influx of beneficial mutations that continuously compensates for the accumulation of the weakly deleterious ones. Mathematical model suggested that such a state can exist for any population size N and mutation rate U and that a surprisingly low fraction of beneficial mutations suffices to achieve stability, even in small populations in the face of high mutation rates and weak selection, maintaining a well-adapted population in spite of Muller's ratchet (Goyal et al., 2012).

Asexually reproducing all-female lineages of plants and animals are often more frequent at higher latitudes and altitudes, on islands, arid environments and in habitats described as transient, ecotonal, disturbed or marginal, a phenomenon called geographical parthenogenesis (Glesener and Tilman, 1978; Bierzychudek, 1985; Beaton and Hebert, 1988; Cuellar, 1994; Peck et al., 1998; Hörandl, 2006). These situations are associated with exposure to unfavorable conditions (Brown, 1984; Hoffmann and Blows, 1994). Small populations frequently occur in marginal environments or near the edges of geographic distributions. Small populations are expected to suffer particularly from Muller’s ratchet (Muller, 1964; Gabriel et al., 1993; Charlesworth and Charlesworth, 1997). Hence, within the conceptual framework of Muller’s ratchet theory of sexual reproduction, it should be highly counter-intuitive that asexual reproduction occurs preferentially in marginal populations. On the other hand, several studies have revealed higher levels of nonsynonymous amino acid substitutions in asexually reproducing organisms as compared to their sexually reproducing sister species (Moran, 1996; Normark and Moran, 2000; Schön et al., 2003; Paland and Lynch, 2006; Barraclough et al., 2007), which supports the idea of deleterious mutation accumulation, although the negative fitness effects of these mutations are generally only assumed. The same patterns (i.e. increased rates of nonsynonymous substitutions), however, are a major line of evidence for strong positive selection on genes in sexually reproducing organisms (e.g. Eyre-Walker, 2006), which emphasizes the importance of discerning the causes and population-level effects of selection at the molecular level (Schwander and Crespi, 2009).

3) The mutational deterministic hypothesis for the origin and maintenance of sexual reproduction (Kondrashov’s hatchet) posits that sex enhances the ability of natural selection to purge deleterious mutations after recombination brings them together into single genomes (Kondrashov, 1988; 1993; Kondrashov and Turelli, 1992). A situation, when the fitness effect of one allele state depends on the allele states at other loci, is called epistasis (Wolf et al., 2000; de Visser et al., 2011; Lehner, 2011;Macía et al., 2012). Kondrashov’s hatchet requires negative/synergistic epistasis, a type of genetic interaction where the combined detrimental effect of two unfavorable alleles is greater than the sum of the individual effects. Several authors have speculated that synergy emerges from competition for resources (King, 1967; Sved et al., 1967) and a mathematical model suggested that, if individuals usually compete in small groups, then competition can lead to synergistic epistasis (Peck and Waxman, 2000). However, theoretical modelling has questioned the importance of epistasis for the evolutionary fate of sexual reproduction (Howard and Lively, 1998; Keightley and Otto, 2006; Kouyos et al., 2007; MacCarthy and Bergman, 2007; Barton, 2009). Moreover, empirical studies on a variety of organisms have not supported the theory (Keightley and Eyre-Walker, 2000; de Visser and Elena, 2007; Kouyos et al., 2007; Barton, 2009), reporting every conceivable form of directional epistasis: synergistic (Mukai, 1969; de Visser et al., 1996; 1997a; Rivero et al., 2003; Salathé and Ebert, 2003), antagonistic (Bonhoeffer et al., 2004; Burch and Chao, 2004; Sanjuán et al., 2004b) and no significant directional epistasis (de Visser et al., 1997b; Elena and Lenski, 1997; Elena, 1999; Peters and Keightley, 2000; Whitlock and Bourguet, 2000; Wloch et al., 2001b; Rivero et al., 2003; Cooper et al., 2005; Martin et al., 2007). Positive epistasis is almost as prevalent as negative/synergistic epistasis (de Visser and Elena, 2007), and this variability in the form of epistasis tends to select more strongly against recombination than in its favor (Otto and Feldman, 1997). There is a common intuition that stress increases selection (Agrawal and Whitlock, 2010). Theoretical analyses of metabolic pathways and biological networks show that a genetic stress (the first mutation) can increase or decrease selection on a subsequent mutation (i.e. negative or positive epistasis) depending on the structure of the network and the relative positions of the two mutations within these pathways (Szathmáry, 1993; Keightley, 1996; Segrè et al., 2005; Sanjuán and Nebot, 2008). Depending on the position of the initial mutation within the pathway, average selection on subsequent mutations could be stronger or weaker than in the absence of the genetic stress (i.e. the initial mutation) (Agrawal and Whitlock, 2010). Sanjuán and Elena (2006) pointed out that epistasis correlates to genomic complexity associated with mutational robustness. Moreover, under constant mutation rates, which is one of the basic assumptions of the deterministic mutation hypothesis (Kondrashov, 1988), sexual and asexual populations of infinite size have equal fitnesses at equilibrium. When mutation rates are fitness-dependent, as is the case in real life, the fitness of a sexual population depends on the shape of the curve relating fitness to mutation rate while the fitness of an asexual population depends only on the mutation rate of the least loaded class (Agrawal, 2002).

Importantly, these theories are rooted in population genetics. According to Lewontin (1961): “the theory of population genetics, as complete as it may be in itself, fails to deal with many problems of primary importance for an understanding of evolution.” and Lewontin (1974, p. 267): “...population genetics is not an empirically sufficient theory …. Built into both deterministic and stochastic theory are parameters and combinations of parameters that are not measurable to the degree of accuracy required.”

Importantly, none of these theories were tested with more realistic assumptions introducing the additional costs of gamete overproduction (see chapter 8.1), and decreased benefit of sex due to canalization (the buffering of deleterious mutations, see chapter 10.4) and phenotypic plasticity. One can predict that with these assumptions, explaining the evolutionary rationale of sex within the framework of population genetics principles would have ended in a grandiose failure.

18.1.1 Flawed static concepts

I always go by official statistics because they are very carefully compounded and, even if they are false, we have no others ...

Jaroslav Hašek, 1911

There is a message hidden in Hašek’s aphorism (Baluška and Mancuso, 2007). “All those mathematical models, scientific theories and concepts, however appealing, harmonious and long-standing … but which do not correspond to reality …; inevitably will be ‘killed by ugly’ facts generated by scientific progress, and finally replaced by new models, theories, and concepts” (Baluška and Mancuso, 2007). A fundamental weakness of mathematical modeling is, as was conceded by Maynard Smith and Brookfield (1983): “A mathematical model is only as good as its assumptions.” (And the assumptions are often rather dictated by computational constraints than by evolutionary reality). However, the assumptions of the genetic theories of sexual reproduction are not based on evidence but deduced from assumed laws or premises set up by population genetic theory (Fisher, 1930). With the “right” assumptions it is even possible to model that under evolving mutation rates, asexuals are able to spread to fixation even when sexuals faced no cost of sex whatsoever (Sloan and Panjeti, 2009). Empirical observations that were made to support the genetic theories showed the long-term advantage of sexual reproduction but did not vindicate the genetic theories despite assertions to the contrary (Rice, 2002). And so, with regard to the genetic theories of sexual reproduction, in Hašek’s sense “even if they are false, we have no others ...”.

Fitness effects of mutations strongly depend on genotype (epistasis) and environment (Cuevas et al., 2002; Elena and de Visser, 2003; Kishony and Leibler, 2003). Regulatory plasticity generates a distribution of phenotypes of a given genotype in response to variation in the environment, or a norm of reaction (Schlichting and Pigliucci, 1998), a term originally coined by Woltereck (1909). All theories come down to the fact that a broad reaction norm smoothes the path of increasing fitness (Frank, 2011). Whether a mutation is adaptive or maladaptive is not an inherent property of the mutation itself but is determined by the environmental conditions (including the genetic background, Estes and Teotónio, 2009) in which the mutation manifests its fitness effects. Adaptation to a new environment will often have negative pleiotropic effects on fitness in the original environment, the so-called ‘cost of adaptation’ (Bell and Reboud, 1997; Agrawal, 2000a; Kassen, 2002; Strauss et al., 2002; Palaima, 2007; Hereford, 2009). There is extensive evidence for the occurrence of genotype x environment interactions (Falconer and Mackay, 1996). In this sense, any adaptive, “beneficial” trait in one environment may become maladaptive, “deleterious” in another ecological context (Elena and de Visser, 2003; Kishony and Leibler, 2003). A population can adapt to a sudden environmental change by using either new mutations or alleles from the standing genetic variation. In the first case the population must wait for the appearance of the desired allele, while in the second it can respond immediately. Species with larger population sizes may have more standing variation with which to respond to novel selection pressures, leading to a smaller fraction of adaptations from new mutations (Hermisson and Pennings, 2005; Leffler et al., 2012). If a population uses standing variation, the alleles selected may have been previously neutral or deleterious and are maintained in the ancestral population through a balance of recurrent mutation, selection and drift (Orr and Betancourt, 2001; Hermisson and Pennings, 2005; Przeworski et al., 2005; Barrett and Schluter, 2008; Pritchard et al., 2010). It has been shown theoretically (Falconer and Mackay, 1996; Kawecki, 1997; Orr and Betancourt, 2001; Lenski et al., 2003; Hermisson and Pennings, 2005; Cowperthwaite et al., 2006) and experimentally (Mills et al., 1967; Futuyma and Moreno, 1988; Fry, 1990; Bennett and Lenski, 1993; Cooper and Lenski, 2000; Elena and Lenski, 2003; Barrett and Schluter, 2008; Beaumont et al., 2009; Philippe et al., 2009; Seixas et al., 2012) that the properties of mutations, whether beneficial or deleterious, are not fixed but depend on a variety of ecological and genetic factors. Thus, a variety of environmental stresses’ influence on deleterious mutations is strongly biased towards the alleviation of mutation effects (Kishony and Leibler, 2003). Due to the costs of adaptation (Kassen, 2002; Hereford, 2009; Fankhauser, 2010), traits that are disadvantageous at one time and environmental condition may be advantageous at another and vice versa (Ellner et al., 1999; Grant and Grant, 2002; Kassen, 2002; Kishony and Leibler, 2003; Childs et al., 2004; Seamons et al., 2007; Sletvold and Grindeland, 2007; Robinson et al., 2008; Beaumont et al., 2009; Simons, 2009; 2011; Greene et al., 2010).

A couple of examples may illustrate the issue. Probably the earliest known demonstration of the existence of a cost of adaptation was an experiment by the Rev. Dallinger (1887), who grew bacteria at steadily increasing temperatures for 7 years. The optimum temperature for his founding population was between 15.5 and 18.3 °C. By the end of his experiment these bacteria were growing and reproducing normally at 70 °C, well beyond their normal thermal limit of 60 °C, but did no longer survive at the ancestral 15.5 °C optimum. Crow (1957) was the first to predict that resistance alleles should have fitness costs in the absence of the selective agent. It can be expected that resistance carries costs when it is not needed, lowering the fitness of defended individuals in the absence of enemies (Rhoades, 1979, Simms and Rausher, 1987, Clark and Harvell, 1992, Sheldon and Verhulst, 1996; Jokela et al., 2000). This prediction has been verified in plants for herbicides, pathogens, and herbivores (Simms and Rausher, 1987; Simms and Triplett, 1994; Bergelson and Purrington, 1996; Purrington and Bergelson, 1999; Burdon and Thrall, 2003; Vila-Aiub et al., 2009), in animals for pathogen resistance (Kraaijeveld and Godfray, 1997; Yan et al., 1997; Langand et al., 1998; Searle and Blackwell, 1999; Webster and Woolhouse, 1999; Woolhouse et al., 2002; Carton et al., 2005), in bacteria for antibiotic resistance (Andersson and Levin, 1999; Komp Lindgren et al., 2005; Andersson, 2006; Rozen et al., 2007; Andersson and Hughes, 2010), and in many pest species for pesticide resistance (Roush and McKenzie, 1987; Andreev et al., 1999; Coustau et al., 2000; Bourguet et al., 2004). Likewise, defences themselves are costly and therefore enhance fitness only when enemies are present (Lively, 1986; Moran, 1992). Individuals with chemical or behavioral defences against predators or herbivores may enjoy high fitness in risky environments compared with undefended individuals, whereas the undefended phenotype may do better in the absence of risk (De Meester et al., 1995; Van Buskirk et al., 1997; Agrawal, 2000b; van Hulten et al., 2006). Whereas studies of susceptibility and pathogenicity generally take a static view of the underlying population genetics, co-evolution is a dynamic process: if host and pathogen co-evolve, then a ‘good’ gene in one time and place may be a ‘bad’ gene in another time and place (Woolhouse et al., 2002). Having eyes, due to their energetic costs, is clearly “deleterious” when an animal is living in a cave (Heininger, 2012). “Thrifty” genotypes are “beneficial” during times of food shortage but are clearly “deleterious” in times of nutritional abundance, as is highlighted by the epidemics of non-insulin-dependent diabetes mellitus and obesity in indigenous populations exposed to Western diets (Neel, 1962; Crespi, 2010; Chakravarthy and Booth, 2004; Prentice et al., 2005; Gluckman and Hanson, 2006; Hancock et al., 2010; Carrera-Bastos et al., 2011). The decline of lactase activity after the suckling period is a physiological, beneficial process in all mammals (Sebastio et al., 1989; Buller et al., 1990; Lacey et al., 1994). However, in dairy economies adult lactose maldigestion may become deleterious and has been profoundly selected against (McCracken, 1971; Gudmand-Høyer, 1994; Vesa et al., 2000; Matthews et al., 2005; Hancock et al., 2010). Thus, lactase gene expression has evolved repeatedly to continue throughout life in dairy farming populations in Europe, East Africa, and the Middle East (Bersaglieri et al., 2004; Harris and Meyer, 2006; Tishkoff et al., 2007; Enattah et al., 2008; Ingram et al., 2009).

Owing to the dynamics of evolution, the beneficial or deleterious value of mutations may change repeatedly. For example, the selective value of a genotype is frequency dependent when its contribution to the following generation relative to alternative genotypes varies with the frequency of the genotype in the population (Ayala and Campbell, 1974). As a gene conferring disease resistance spreads through a population, the incidence of infection declines, reducing the fitness advantage of carrying the resistance gene. Thus genes conferring complete resistance cannot become fixed (i.e., universal) by selection in a host population, and diseases cannot be eliminated solely by natural selection for host resistance (Roy and Kirchner, 2000). Moreover, this frequency-dependent selection has its short- and long-term components. In the latter, density dependence becomes essential (Heino et al., 1998). Moreover, an adaptive mutation’s effect may depend on the presence of other, possibly nonadaptive, mutations (Zuckerkandl and Pauling, 1965; Lockless and Ranganathan, 1999; Reetz et al., 2005; Weinreich et al., 2006; Bloom and Arnold, 2009; Breen et al., 2012).Particular mutations appear to have different effects in high- versus low-fitness virus lines (Silander et al., 2007) and more complex organisms such as C. elegans (Estes and Lynch, 2003; Betancourt, 2007).

If one considers mutations at more than one locus, they may combine in several ways in their final effect on fitness. Two mutations that individually have no significant effect on a trait under selection can in combination be highly advantageous or deleterious. Well known examples for such epistatic interactions (Phillips, 2008) include resistance evolution in pathogens (Hall, 2002; Weinreich et al., 2006; Lozovsky et al., 2009) or metabolic changes in yeast (Segrè et al., 2005). Epistasis describes an important property of genotype-phenotype maps: the phenotypic effect of a genetic change at a single locus may depend on the values of other genetic loci, in other words, the nonlinear interaction of effects among alleles at different loci (Wolf JB et al., 2000). That such correlations should exist is not at all surprising. Given the many multi-scale physical processes involved in translating a genotype into a phenotype, it is rather the absence of epistasis that might be expected to be the exception to the rule (Schaper et al., 2011). Genetic constraint on fitness landscapes is due to fitness epistasis, where a mutation’s adaptive value depends on the genetic background in which it arises (Phillips, 2008). Sign epistasis occurs when mutations are beneficial within the context of some genetic backgrounds, but detrimental within others. To exhibit multiple fitness peaks, a biological system must contain reciprocal sign epistatic interactions, which are defined as genetic changes that are separately unfavorable but jointly advantageous. Reciprocal sign epistasis should be pervasive in nature as it is a logical consequence of specificity in molecular interactions and it creates rugged fitness landscapes (Weinreich et al., 2005; 2006; Poelwijk et al., 2007; 2011; Schoustra et al., 2007; Carneiro and Hartl, 2010; Kvitek and Sherlock, 2011). The presence of sign epistasis was established in several recent experimental studies, where all combinations of a selected set of individually beneficial or deleterious mutations were constructed and their fitness effects (or some proxy thereof) were measured (Weinreich et al., 2006; Lozovsky et al., 2009; de Visser et al., 2009; Carneiro and Hartl, 2010). In genetic networks that differ in their complexity and robustness against perturbations but that perform the same tasks, robustness increased with complexity and epistasis was found to be positive for small nonrobust networks but negative for large robust ones (Macía et al., 2012). Epistasis is pervasive in long-term protein evolution: about 90 per cent of all amino-acid substitutions have a neutral or beneficial impact only in the genetic backgrounds in which they occur, and must therefore be deleterious in a different background of other species (Breen et al., 2012): This raises the possibility that similar epistatic interactions may be prevalent in short-term evolution (Salverda et al., 2011; Woods et al., 2011) and that situations when a polymorphism is benign or beneficial to one individual but deleterious to another individual within the same population may be more common than is thought at present (Breen et al., 2012).An in silico evolution experiment (Lenski et al., 2003) demonstrated thata mutation that was highly deleterious when it appeared was highly beneficial in combination with a subsequent mutation.In some cases, mutations that were deleterious when they appeared served as stepping-stones in the evolution of complex features.

Under these circumstances, the beneficial-deleterious distinction becomes a moving target. Fisher (1930) was able to show that, unless the degree of inbreeding varies or the environment deteriorates, the mean fitness always increases (MacArthur, 1962). However, Fisher thought that environmental changes are so ubiquitous that, as he once said, Wright's peaks and valleys are more like the undulating wave crests and troughs of an ocean than a mountainous landscape. He believed that a population rarely, if ever, finds itself in a position where no allele frequency change could increase its fitness (Crow, 1987). This forces us to think of selection itself as a dynamical process, that is, to promote the static picture of fitness landscapes to dynamic fitness “seascapes”. Recently, the static concepts of fitness and fitness landscape were supplemented by the dynamic concepts of fitness seascape (Mustonen and Lässig, 2010) that takes the ever changing nature of environmental conditions into account. The dynamical approach leads to a quantitative measure of adaptation called fitness flux, which counts the excess of beneficial over deleterious genomic change (Mustonen and Lässig, 2009). The recently formulated fitness flux theorem generalizes Fisher’s fundamental theorem of natural selection to evolutionary processes including mutations, genetic drift, and time-dependent selection. It shows that a generic state of populations is adaptive evolution: there is a positive fitness flux resulting from a surplus of beneficial over deleterious changes. Fitness flux is a universal measure of adaptation in molecular evolution (Mustonen and Lässig, 2010). An increase of cumulative fitness flux is an almost universal evolutionary principle of biological systems. Positive contributions to the fitness flux arising from adaptive genotype changes accumulate over evolutionary periods of time. Negative contributions are limited to time intervals with a systematic loss of adaptation which cannot occur continuously in viable populations. These predictions are in accordance with experiments in bacteria and bacteriophages and with genomic data in Drosophila. In this sense, fitness flux is a more fundamental characteristic of evolution than fitness, for which no comparable growth law holds (Mustonen and Lässig, 2010).

The genetic theories of sexual reproduction completely ignore a variety of gene x environment interactions:

  • the fitness effects of genes only unfold within a certain ecological context
  • a multitude of genes has pleiotropic effects, that can be both synergistic or antagonistic, i.e. many genes may have both beneficial and deleterious effects (e.g. on soma and germline cells) that can be described by cost-benefit trade-offs (see Heininger, 2012)
  • alleles have both frequency- and density-dependent effects
  • epistatic effects are widespread and may be both synergistic and antagonistic and highly context-specific (sign epistasis)
  • the phenotypic effects of genes are subject to profound epigenetic regulation and canalization hiding deleterious mutations from purging by selection
  • the fitness landscape is in constant motion and should rather be described as fitness seascape (as already perceived by Fisher, see Crow, 1987)

Finally, Muller’s ratchet and Kondrashov’s hatchet face a fundamental dilemma. Evolution depends on genetic variation (Whitehead and Crawford, 2006). Classical theory holds that genetic variation is largely determined by neutral and slightly deleterious mutations (Barton and Turelli, 1989; Charlesworth and Hughes, 2000). Neutral mutations, since they, by definition, should not be seen by selection should not have any influence on the selective advantage of sexual reproduction. If Muller’s and Kondrashov’s theories would be correct, sexual reproduction would reduce genetic variation, slowing evolvability in variable environments.

18.1.2 Flawed teleological concepts

Evolution is not teleological in the sense that its processes or actions are for the sake of an end, i.e., the Greek “telo” or final cause. Clearly, once an organism has survived and/or reproduced one can point to its various attributes and say “yes, that attribute appears to have contributed to the organism’s survival/reproduction". However, that is no more evidence of “foresightedness” than a lottery winner saying “I chose these lottery numbers (or bought those particular scratch-off tickets) because I knew they would be winners". This is known as the “fallacy of affirming the consequent” (also called post hoc, ergo propter hoc argumentation) and is logically inadmissible in the natural sciences (MacNeill, 2009). `Backwards causation', by which some future state or event influences (`causes') an action in the present or past, is often characteristic of teleological arguments. The Modern Synthesis took pride in having discouraged such thinking (Mayr, 1992).

On the basis of genome-wide sequencing it has been estimated that, on average, each person carries approximately 250 to 300 loss-of-function variants in annotated genes and 50 to 100 variants previously implicated in inherited disorders (The 1000 Genomes Project Consortium, 2010). There is no reason to think that in other species the situation is fundamentally different. (Although, maybe the relaxed selection in modern human societies has aggravated the situation. But, on the other hand, this relaxed natural selection does not yet exist for too many generations and the SMSCs keep relaxed natural selection from doing too much harm.) Obviously, evolution and its bet-hedging strategy tolerate a lot of imperfection. Presumably, every organism is a mosaic of beneficial and deleterious mutations. Mind you, probably hundreds of both. Could recombination work assortative to combine the deleterious and beneficial mutations in separate organisms? On the basis of which information? Assuming that recombination, a quasi-stochastic, uncoded, process with e.g. an estimated 25,000–50,000 crossover hotspots in the human genome should be able to help natural selection to discriminate between beneficial or deleterious mutations, in essence implies that evolution is understood to be designed and foresighted. Isn’t it more probable that recombination just shuffles the mutations in varying composition irrespective of their fitness value? And often enough the deleterious mutation is not “seen” by evolution due to canalization. And who knows, maybe around the evolutionary corner the now deleterious mutation is the stepping stone of an innovation (Lenski et al., 2003)? In the end, what really counts is the compound effect of beneficial and deleterious mutations and most of all the variation.

The genetic theories use an outcome measure (the deleterious or beneficial fitness effect of a mutation) to explain the evolutionary rationale of sexual reproduction. The only evolutionary process that can differentiate between deleterious and beneficial mutations is natural selection that selects organisms for their reproductive success(Ellegren and Sheldon, 2008). The Modern Synthesis’ concept of natural selection is only plausible when one assumes a straightforward causation of phenotype by genotype (Huang, 2012). However, such simple 1:1 mapping must now give place to the concepts of gene regulatory networks and gene expression noise. Both can, even in the absence of genetic mutations, jointly generate a diversity of inheritable randomly occupied phenotypic states that could serve as a substrate for natural selection (Huang, 2012). Sexual reproduction uses several genetic and epigenetic tools such as mutations, epimutations, recombination, transposon mobility and mutability of simple sequence repeats that may yield functionally equivalent fitness results.This degeneracy of the molecular toolsin addition to the stochasticity of the genotype-phenotype mapping are so ambiguous that no certain molecular process can be assigned to a phenotypic effect.Thus, the myriad of molecular processes that determine an individual’s phenotype have no built-in sensor and allow no feedback that informs about the selective value of an action. These processes are totally blind with regard to their fitness-relevance.

According to Darwin (1859) evolutionary trajectories are determined not by the individuals in the trailing tail of the fitness distribution but by the more competitive, fitter individuals. It can hardly be advocated that the evolutionary success of sexual reproduction is based on purging the population from less fit individuals. Accordingly, the model proposed by Kondrashov has been criticized for modeling assumptions far from real-world observations (MacCarthy and Bergman, 2007). In fact, Wallace advocated the elimination of the unfit (Smith CH, 2012a; b) but this elimination is brought about by competition with the fitter individuals. Beneficial mutations that do occur are much more likely to achieve long-term evolutionary success and the dynamics of populations than deleterious mutations (Haldane, 1927; André and Godelle, 2006; Patwa and Wahl, 2008). Several mathematical models have demonstrated that deleterious mutations do not contribute significantly to overall evolutionary dynamics in larger populations when at least some fraction of mutations is beneficial (Rouzine et al., 2003; 2008; Desai and Fisher, 2007; 2011; Desai et al., 2007; Zeyl, 2007; Fogle et al., 2008; Park SC et al., 2010). Accordingly, as has been shown by Hallatschek (2011), evolutionary dynamics of an asexual population are determined by traveling waves of increasingly fitter genotypes (Fisher, 2011; Hallatschek, 2011). These theoretical findings were confirmed by experimental evolution dynamics: in the process of periodic selection, evolution is characterized by a series of clonal replacements, and diversity is reduced each time a beneficial mutant arises and sweeps through the population (Novick and Szilard, 1950; Atwood et al., 1951; Zambrano et al., 1993; Radman et al., 1999; Elena and Lenski, 2003; Maharjan et al., 2012). Thus, evolutionary dynamics are rather determined by fitness-dependent competition between members of the population than the accumulation of deleterious mutations (Maharjan et al., 2012). Since deleterious mutations do not affect adaptive evolution in asexual populations, the theoretical, so far even unproven, consideration that sexual reproduction may be able to purge the genome from deleterious mutations by truncation selection hardly can serve as rationale for the evolutionary origin of sexual reproduction. Moreover, selection reduces the effect of Muller’s ratchet over time (Itoh et al., 2002; Allen et al., 2009). The operation of Muller’s ratchet, resulting in mutational meltdown (Lynch et al., 1993; Vrijenhoek, 1998) may be a consequence of being not sexual, but it cannot serve as rationale for the origin of sexual reproduction which has to be searched at the other end of the fitness distribution, contingent upon the evolutionary dynamics of beneficial mutations. Thus, sexual selection does prevail because it is able, at least under most ecological conditions, to generate and select for the fitter individuals. Thus, when competitive fitness, as in K-selected habitats is relevant, sexual reproduction prevails.

18.2 DNA repair

The DNA repair hypothesis for the maintenance of sex invokes a proximate (or mechanistic) explanation positing that recombination is necessary for the repair of double-strand DNA damage eliminating new genetic variation (Bernstein et al., 1984; 1985a; Bernstein and Bernstein, 1991; Michod, 1993). Cells with DNA damage have been shown to undergo sex at higher rates in viruses (Bernstein, 1987), bacteria (Wojciechowski et al., 1989), and yeast (Bernstein and Johns, 1989). It was suggested that the evolutionary function of transformation may be to provide template strands for DNA repair (Felsenstein, 1974; Bernstein et al., 1985a; b; 1989; Michod et al., 1988; 2008; Long and Michod, 1995; Hörandl, 2009). As consequence of the maintenance of ploidy, elimination of chromosomal rearrangements and DNA repair by homologous recombination, sexual reproduction may reduce genetic variation (Shields, 1982; 1988; Gorelick and Heng, 2011). In fact, many proteins that function during somatic DNA-damage detection and repair are also active during homologous recombination. However, their meiotic functions may be altered from their somatic roles through localization, posttranslational modifications and/or interactions with meiosis-specific proteins (Marcon and Moens, 2005).Redfield (1993a; b) argued that if DNA repair is the primary function of transformation, one would expect competence to be regulated by DNA damage. However, this is not the case in B. subtilis and H. influenzae (Redfield, 1993a). Szathmáry and Kövér(1991) tested the repair hypothesis by comparing outcrossing sexuality with a hypothetical parthenogenic strategy (the Prudent Reparator) which destroys as little heterozygosity during repair as possible. The Prudent Reparator could solve the problem of repairing damage as well as that of invading an existing outcrossing population. They concluded, as we do not see this strategy widely adopted instead of sexuality, the repair hypothesis is likely to miss some essential feature of the evolution of sex.Burt (2000) asked: “Why should sexual taxa carry the costs of sex if they could engage in asexual DNA repair, e.g. by ameiotic recombination, at considerably lower costs?” Burt (2000) dismissed the idea that “meiotic crossing-over is just an incidental by-product of DNA repair processes” by arguing: “The most pressing problem with the idea is that (meiotic) crossing-over involves only one of the two sister chromatids of each chromosome, and so under the hypothesis only one of them should be damaged. However, if this is the case, why not just repair it with the sister chromatid? This is what usually happens in mitotic cells if damage occurs between DNA replication and cell division (Fabre et al., 1984; Kadyk and Hartwell, 1992). However, in meiotic cells, the preferred template for “repair” is the homolog, not the sister chromatid (Schwacha and Kleckner, 1994; 1997). This is as expected if the function of meiotic crossing-over is to produce recombinant chromosomes, but not if it is simply to repair broken chromosomes.”

Many asexual taxa are thought to be particularly efficient in DNA repair, which would allow them to reduce the accumulation of deleterious mutations (Castonguay and Angers, 2012). There is evidence for this in asexual taxa such as asexual weevils (Tomiuk and Loeschcke, 1992), aphids (Normark, 1999), darwinulid ostracods (Schön et al., 1998), Daphnia (Omilian et al., 2006), and oribatid mites (Schaefer et al., 2006). This led Schaefer et al. (2006) to ask “why not more taxa are ancient asexuals if the long-term disadvantages of parthenogenetic reproduction can be defeated.”

Stressing the conservative role of sexual reproduction, the proponents of the DNA repair hypothesis largely ignore the innovative role of sex (but see Hörandl, 2009). Whether recombination is DNA reparative, or mutagenic is condition-dependent. Thus, recombination as genetic variability- and innovation-creating mechanism (Bürger, 1999; Ochman et al., 2000; Narra and Ochman, 2006; Jeon et al., 2008) was advocated as the evolutionary rationale of bacterial transformation and hence the evolution of sex, while the role of DNA repair was contested (Mongold, 1992; Redfield, 1993a; Redfield et al., 1997). In microorganisms environmental stress can alter the balance between stability/repair and mutagenesis/exploration/bet-hedging (Foster, 2005). In an E. coli model, stationary-phase sigma factor RpoS (which encodes the stress response σS transcription factor)-controlled switch from high-fidelity to mutagenic double-strand break repair activates stress-induced mutagenesis during stationary phase/starvation (Ponder et al., 2005). B. subtilis has similar stationary phase/starvation-dependent mutagenic programs (Robleto et al., 2007). Lower levels of oxidative stress favor DNA repair while higher levels favor mutagenesis (Greenberg and Demple, 1988; DeRose and Claycamp, 1991; McBride et al., 1991; Escarceller et al., 1994; Kato T et al., 1994; Blanco et al., 1995; Hu et al., 1995; Touati et al., 1995; Urios et al., 1995; Wang and Humayun, 1996; MacPhee, 1999; Ruiz-Laguna et al., 2000; Bjedov et al., 2003; Cooke et al., 2003). Cellular stress also downregulates DNA repair and elicits mutagenesis in mammals (see chapter 10.1). Importantly, the repair/mutagenesis balance is modulated by interactions with heat shock proteins (Liu and Tessman 1990a; b; Donnelly and Walker, 1989; 1992; Petit et al., 1994) which, in dependence of the cellular energy level may provide the regulatory feedback with cellular metabolic/oxidative homeostasis and its derangement during stress (MacPhee, 1985; 1994; 1996; Seetharam and Seidman, 1992, Keszenman et al., 2000).

Different evolvability characteristics can be optimal under different circumstances as suggested by experimental studies in a variety of organisms (Drake et al., 1969; Nöthel, 1987; Sniegowski et al., 1997; Schapper, 1998), and computational evolutionary models (Bedau andPackard, 2003; Clune et al., 2008; Dees and Bahar, 2010). The fossil record indicates that evolutionary dynamics proceed by punctuational episodes rather than gradual change (Eldredge and Gould, 1972; Gould and Eldredge, 1977; 1993; Sneppen et al., 1995; Eldredge et al., 2005). In well adapted populations living in stable habitats, conservation may become more important than innovation. Well-documented examples of stasis range from Paleozoic brachiopods (Lieberman et al. 1995) to late Cenozoic bivalves (Stanley and Yang, 1987) and bryozoans (Jackson and Cheetham, 1999). But species-wide depletion of accessible beneficial mutations requires a degree of environmental constancy that is not typical of the earth’s history (Lambeck and Chappell, 2001; Zachos et al., 2001; Eldredge et al., 2005). Evolutionary models based on the asexual and sexual replication pathways in Saccharomyces cerevisiae suggested that sexual replication can eliminate genetic variation in a static environment, as well as lead to faster adaptation in a dynamic environment (Gorodetsky and Tannenbaum, 2008).

18.3 Host-parasite coevolution: the Queen abdicates in favor of the pluralistic, coevolving web of life

Now here, you see, it takes all the running you can do, to keep in the same place.

Lewis Carroll: Through the Looking Glass

The cited lines were the Red Queen’s explanation to a confused Alice as to why she could run as fast as she could in Wonderland but never get anywhere, a situation analogous to the constant evolutionary pressure exerted by a changing environment. Species evolve in response to their biotic environment and this can lead to coevolution between species that interact in either a mutualistic or an antagonistic fashion (Futuyma and Slatkin, 1983). Parasites are generally under selection to infect the most common genotypes in the local host population, assuming that (i) they contact hosts at random and (ii) that the different parasite genotypes are only able to infect a subset of the host-resistance genotypes. This parasite-mediated frequency-dependent selection can then lead to genetic polymorphism in both the host and parasite, as first recognized by Haldane (1949). Hence, if an asexual clone becomes the most common genotype in an otherwise sexual population, the parasites would be expected to quickly evolve to infect it. If the parasites evolve to disproportionately infect the clone, they may prevent, or help prevent, the clone from replacing the sexual population (e.g., Jokela et al., 2009). Thus, it is argued that a parasites have contributed to the short-term maintenance of sex, even if they do not drive the clone to extinction (Lively, 2010). Accordingly, one of the most widely accepted hypotheses posits that sexual reproduction evolved in hosts to counter the pressure of parasitic coevolution (Levin, 1975; Jaenike, 1978; Hamilton, 1980; Hamilton and Zuk, 1982, Hamilton et al., 1990, Howard and Lively, 1994; 1998; 2002; Møller and Saino, 1994; Able, 1996; John, 1997; Ooi and Yahara, 1999; Martins, 2000; Agrawal, 2006a; Salathé et al., 2008; Lively, 2010; Morran et al., 2011). The ‘‘Red Queen’’ hypothesis predicts an immediate advantage of being sexual, essentially because the highly diverse offspring resulting from sexual reproduction provide a moving target for parasites, whereas the genetically very uniform offspring of clonal females should be easily exploited by parasites. This type of “over-reactive” frequency-dependent selection has been shown to lead to cyclical dynamics, where parasites eventually track common host genotypes and, assuming there is a fitnes cost associated with infection, drive them to lower frequency (Hamilton, 1980; Neiman and Koskella, 2009). The occurrence and characteristics of the predicted oscillatory cycles have been the focus of numerous theoretical and empirical studies (e.g., Jaenike, 1978; Bell, 1982; Hamilton et al., 1990; Dybdahl and Lively, 1998; Koskella and Lively, 2007). Covariance between the relative frequency of sex/outcrossing and the frequency of parasitism is now established in a wide range of taxa, providing broad but indirect support for a role for the Red Queen in the maintenance of sex (Lively, 1987; 1989; 1999; Lively et al., 1990; Moritz et al., 1991; Schrag et al. 1994; Lively and Dybdahl, 2000; Lively and Jokela 2002; Busch et al., 2004; Kumpulainen et al., 2004; Lively et al., 2004; Fischer and Schmid-Hempel, 2005; Tobler and Schlupp, 2008; Jokela et al., 2009; King et al., 2009; King and Lively, 2009; Morran et al., 2011). A number of studies indicate that host genetic diversity plays an important role in buffering populations against widespread epidemics, and that parasites represent powerful selective agents in natural populations (Jokela and Lively, 1995; Gaffney and Bushak, 1996; Dwyer et al., 1997; Paterson et al., 1998; Coltman et al., 1999; Little and Ebert, 1999; 2001; Hedrick et al., 2001; Altizer et al., 2003; Paterson et al., 2010). Comparative studies have found particularly high rates of molecular evolution in genes associated with infection (Buckling and Rainey, 2002b; Brockhurst et al., 2003; Blanc et al., 2005; Mu et al., 2006; Barrett et al., 2009; Paterson et al., 2010) or resistance to infection (Hedrick, 1994; Hughes et al., 1994; Kuma et al., 1995; Obbard et al., 2006; Clark et al., 2007). It has been argued that escape from coevolving parasites allows desiccation-tolerant rotifers (Wilson and Sherman, 2010) and microscopic algae that alternate between virus-sensitive diploid and virus-resistant haploid phases (Frada et al., 2008) to keep up asexual reproduction, vindicating the Red Queen hypothesis (Leung et al., 2012). Likewise, desiccation-tolerant (Clegg, 2005) and diapausing eggs (Lass and Ebert, 2006) may represent an escape strategy from infectious agents (Ladle et al., 1993).

However, the role of host-parasite interactions for the maintenance of sexual reproduction is not uncontested. Empirical studies testing the predictions of the Red Queen hypothesis have also shown negative results (Brown et al., 1995; Hanley et al., 1995; Vernon et al., 1996; Weeks, 1996; Ben-Ami and Heller, 2005; 2008; Tobler and Schlupp, 2005; 2008; Killick et al., 2008; Elzinga et al., 2012). Data from cyclical parthenogenetic Daphnia even suggest that sexual reproduction may erode immunocompetence (see also Heininger, 2012) and could inhibit the evolution of parasite resistance (Duncan et al., 2006). Furthermore, the host-parasite coevolutionary dynamics that are required for operation of the Red Queen select for genetic diversity rather than sex per se (Lively and Howard, 1994). This means that, all else being equal, a genetically diverse array of asexual lineages is as well-equipped to deal with parasitism as a sexual population (Glesener and Tilman, 1978; Lively and Howard, 1994; Lythgoe, 2000). Accordingly, positive associations between parasitism and outcrossing are not a ubiquitous feature of mixed sexual/asexual animal (Ben-Ami and Heller, 2005; 2008) and plant populations (Parker, 1994). Stearns (1990) argued that parasites, and the diseases they cause, are only important agents of selection when they have significant consequences for the fitness of the host. Reduced virulence and prevalence of, and increased resistance to, parasites may work against the maintenance of sexual reproduction. The strong selection required to maintain sex in many models has also been a source of concern for many theoreticians because it implies that only very virulent and/or highly prevalent parasites will be able to maintain sex and outcrossing (May and Anderson, 1983; Howard and Lively, 1994; Ochoa and Jaffé, 1999; Otto and Nuismer, 2004; Peters and Lively, 2007; Salathé et al., 2008; Neiman and Koskella, 2009). In a broad-scale assessment of ecological and genetic interactions that could underpin the Red Queen process, Otto and Nuismer (2004) concluded that interactions between species are an unlikely explanation of the prevalence of sex (given the restricted definition of sex = recombination). If anything, species interactions select against higher levels of recombination. To derive general results, an assumption was made that the populations studied were at quasi-linkage equilibrium (QLE). This assumption follows from Barton’s (1995) genome-wide model for studying the evolution of recombination, which Otto and Nuismer extended to several interacting species. As QLE means that host and parasite populations are essentially able to track changes in the genotypic composition of one another’s populations, the negative effect of increased recombination predominates. There is therefore little advantage to modifiers that increase the rate of recombination. Even if the modifier becomes associated with rare favored combinations, this advantage is weak when recombination is frequent and is swamped by the cost of producing unfit combinations of alleles (Pomiankowski and Bridle, 2004).

Coevolutionary interactions in a dynamic ecosystem are highly pluralistic (Stenseth, 1985; Dieckmann and Law, 1996; Raffel et al., 2008; Thompson, 2009). This amounts to allowing feedbacks to occur between the evolutionary dynamics of a species and the dynamics of its environment (Lewontin, 1983). Abiotic and biotic environments are highly variable across space and time. Nature is relentlessly tooth-and-claw, continuingly reshaping the evolution of antagonistic interactions within and among species such as predator-prey interactions (Glesener, 1979; Yoshida et al., 2003). And, species inherently form intraspecific and interspecific partnerships that are often commensalistic or mutualistic and allow diversification into new adaptive zones. Populations continually evolve and interacting species continually coevolve, building a constantly coevolving web of life (Thompson, 2005; 2009). Thus it has a too narrow scope to only concede host-parasite interactions the status of a “queen-maker”. It is the incessantly changing biotic and abiotic environment acting in Red Queen and Court Jester dynamics (Benton, 2009; Ezard et al., 2011) whose adaptive stress on organisms may interact condition-dependently (Murray et al., 1998; Folt et al., 1999; Siva-Jothy and Thompson, 2002; Harley, 2003; Vinebrooke et al., 2004; Bolnick and Preisser, 2005; Mitchell et al., 2005; Kolluru et al., 2006; Beketov and Liess, 2007; Doroszuk et al., 2007; Rollo, 2007; Coors and De Meester, 2008; Ebert, 2008; Seppälä and Jokela, 2010; Violle et al., 2010; Seppälä et al., 2011) making sexual reproduction the superior reproductive strategy (Gessler and Xu, 2000; West et al., 2001; Hadany and Beker, 2003b; Hadany and Otto, 2007; 2009; Schoustra et al., 2010). Taken together, host-parasite coevolution is only one of a multitude of factors in a world of coevolutionary ecological webs (Thompson, 2005; 2009) that select for risk-spreading bet-hedging.

19. Evolutionary enigmas and controversies


Understanding sex and recombination has been perhaps the most puzzling issue in evolutionary biology—which suggests that its solution might give us a much better intuition about the evolutionary process in general.

NH Barton (2010)

Summary

Recognizing both stochasticity and selection as the organizing principles of evolution helps to resolve some of the most enigmatic characteristics and long-standing controversies of evolution, including the selection-genetic variation enigma, the levels of selection controversy, the selectionist-neutralist controversy, the adaptationism controversy and the mystery of low heritability. All of them are based on the perpetual combat of chaos and order, uncertainty and direction, between arithmetic and geometric mean fitness, making choices to optimize either individual- and population-level fitness. During the rise in atmospheric oxygen, sexual reproduction appears to have “learned” to use oxidative stress as evolutionary innovation that gave the coevolutionary germ-soma conflict an unprecedented dynamic, resulting in the Cambrian explosion.

The obvious environmental modulation of patterns of inheritance has renewed interest in Lamarckian-type interpretations of evolutionary processes. However, the concepts advocating a Lamarckian type of evolution did not take into account that the various processes, perfectly designed as they may appear, have inherently stochastic elements that allow a large sampling of the (epi)genetic search space. Selective sampling from this quasispecies cloud makes the process Darwinian. Darwinian principles of quasi-stochastically generated variation and selection also operate between sub-organismal entities, e.g. organelles and cells. If the process of germ cell competition generating pre-selected variation is overlooked, phenotypically the impression of a Lamarckian-type evolutionary process may be created.

Reproduction systems affect many population genetic processes, and thus genome evolution. Accordingly, the presented insight into sexual reproduction-related mutagenesis-selection cascades can shed light on a variety of evolutionary enigmas and controversies. Importantly, recognizing the dualism of stochasticity and selection in evolution helps to resolve some of these enigmas, e.g. the levels of selection enigma, the selection-genetic variation enigma, the adaptionism controversy, and the mystery of low heritability.

19.1 The selection-genetic variation enigma

A conundrum of evolutionary biology and population genetics is the coexistence of two basic observations (Walsh and Blows, 2009; Leffler et al., 2012): in natural populations genetic variation is found in almost all traits (Mousseau and Roff 1987; Houle 1991; 1992; 1998; Hill and Caballero 1992; Lynch and Walsh, 1998) in the presence of strong stabilizing natural and sexual selection (Haldane, 1949b; Clarke, 1979; Endler, 1986; Kingsolver et al., 2001; Hereford et al., 2004; Johnson and Barton, 2005). These two observations are in direct conflict as stabilizing selection should deplete genetic variation (Tomkins et al., 2004; Johnson and Barton, 2005; Walsh and Blows, 2009). Two main models of stabilizing selection have been developed. In one, a high rate of mutation at each locus is assumed and, therefore, very many alleles segregate: this assumption predicts high variance in the quantitative trait, but requires mutation rates per locus that are much higher than those observed (Kimura, 1965). In the other model, a low, realistic (at least according to classical theory) rate of mutation per locus with large effects of each mutant is assumed: but this predicts much less variation in the trait than is observed (Turelli, 1984). Lewontin (1974, p. 267) recognized the paradox of genetic variation and dismissed its explanation by the balance (Dobzhansky, 1970) and neutral (Kimura, 1983) theories: “On the one hand, there are strong reasons for rejecting a balance theory because it predicts tremendous inbreeding depressions that are not observed, because the rates of evolution of different molecules strongly suggest that the least functional evolve fastest, because heterozygosity does not seem to be sensitive to ecological stringency, and because selection has proved extremely difficult to find in operation. On the other hand, the theory that standing variation and most substitutions of amino acids have been neutral also strains our credulity to the limit, because it requires us to believe that population sizes for all species are effectively the same, because it requires adaptive mutation, to be several orders of magnitude less frequent than neutral changes, because there is too much variation from locus to locus in the amount of divergence between populations, because of striking similarities in allelic frequency distributions in closely related species, and because the majority of polymorphic substitutions do alter the functional properties of enzymes.” In fact, the increase in additive genetic variance due to mutation is ~10-3 of the standing variation, which makes a significant contribution after ~50 generations of selection (Hill, 1982b; Lynch and Walsh, 1998; Partridge and Barton, 2000). Lack of mutations does not limit straightforward selection response.Sustained responses to long-term artificial selection do not exhaust genetic variation (Frankham, 1980; Yoo, 1980b; Hill, 1982a; b; Dunnington and Siegel, 1996; Eitan and Soller, 2004; Laurie et al., 2004; Carlborg et al., 2006; Burke et al., 2010), arguing for a steady supply of new mutations.

A related conundrum that has been a constant challenge to evolutionary biologists is the problem of polymorphism maintenance in small populations (Wright, 1931; Dobzhansky et al., 1970; Carson, 1990; Grant, 1991). How can genetic diversity be maintained in small isolated populations, in spite of genetic drift, potential inbreeding, or sporadic bottleneck events? Long-term studies in natural populations (Nevo et al., 1974; 1994; 1997a; b; Nevo, 1989, 1991; 1999; Carson, 1990; Coates, 1992; Hartl and Hell, 1994) indicate that polymorphism could be preserved in rather small populations (less than 50-100 individuals), or even in those subjected to or recovered from narrow bottlenecks (Bryant and Meffert, 1986; Carson, 1990; Dinerstein and McCracken, 1990; Coates, 1992; Hartl and Hell, 1994).

The coexistence of selection and high genetic variation in natural populations is what can be expected in the light of the stochasticity-selection dualism of evolution. The genomic “fossil record” (Buss, 1987 p. 90) reflects in contemporary genomes which evolutionary forces shaped the genomic architecture during deep evolutionary time. Thus, the high genetic variation in populations witnesses the strongpremium that evolution paid in the evolutionary past to populations that took precautions for change. Sexual reproduction and its bet-hedging strategy is the evolutionary tool that integratedthe generation of genetic variation (the response to stochasticity) and pre-selection (ensuring viability even in small populations) into a single process.

19.2 Levels of selection controversy

Darwin consistently saw natural selection as choosing between individual organisms. Since Darwin (1859) wrote: “..natural selection works solely by and for the good of each being, all corporeal and mental endowments will tend to progress towards perfection.”, natural selection is often characterized as an optimizing process that results in maximizing of mean fitness via the differential reproduction of highly adapted phenotypes (Emlen, 1966; MacArthur and Pianka, 1966; Williams, 1966a; Dawkins, 1976; Maynard Smith, 1978b; Alexander, 1982; Grafen, 1999). The individual-as-maximizing-agent (Grafen, 1999), however, is a contradiction in itself. Evolutionary endpoint of such an organism would be a Darwinian demon (Heininger, 2012). A Darwinian demon can produce infinitely many offspring and live indefinitely. This “the winner takes it all” scenario is ruled out in a world of limited and unpredictably available resources. Accordingly, the proponents of the individual-as-maximizing-agent concept do have nothing in favor of their strong beliefs, nothing in addition to the orthodoxy of Darwinian thinking and population genetic dogmas. But: “A mathematical model is only as good as its assumptions” (Maynard Smith and Brookfield, 1983). Observational data reflect another evolutionary reality. Not surprisingly, there is a long-standing controversy about the units of selection. In the 1960ies and ‘70ies there was a strong opposition (Maynard Smith, 1964; 1976b; Williams 1966a) to the group-selective concept as advocated by Wynne-Edwards (1962). It was argued that group selection forces are weak and that only in special circumstances an advantage to the group can outweigh a disadvantage to the individual (Williams 1966a; Dawkins 1976; Maynard Smith, 1976b; Leigh, 2010b). In the last decades there has been a renaissance of group-selective concepts often formulated as hierarchical or multi-level selection theory (Lewontin, 1970; Frank, 1998; Sober and Wilson, 1998; Keller, 1999; Michod, 1999b; Gould, 2002; Hammerstein, 2003; Borrello, 2005; 2010; Okasha, 2006; Godfrey-Smith, 2009; Wilson, 2012). Regrettably, the level-of-selection debate was often loaded with ideological ballast. The good-for-the-individual or good-for-the-species arguments are anthropomorphic concepts that do not reflect evolutionary reality. Evolution does not act in terms of selfish (e.g. Dawkins, 1976) or selfless (eg. Wilson and Wilson, 2007) categories.Subliminally insinuating a goal-directed intention, these conceptsrevealteleological thinking.

The recognition of the stochasticity-selection dualism in evolution lays the conceptual groundwork for a deeper understanding of the levels of selection.When the fitness of a genotype varies over generations, the appropriate measure of its relative growth rate is its geometric mean fitness, rather than its arithmetic mean fitness. The geometric mean of n numbers is the nth root of their product. If the numbers vary, then the geometric mean is always less than the arithmetic mean; in general, the geometric mean becomes smaller as the numbers being averaged become more variable. Thus the geometric mean fitness of a genotype can be increased by reducing the variance of its fitness (over generations), even if the reduction of variance also entails a reduction of the arithmetic mean. The principle is similar to risk aversion in utility theory; the cost of a negative deviation from the mean is larger than the benefit of an equivalent positive deviation (Philippi and Seger, 1989).

When fitness fluctuates through time and the fluctuations are modest, the identity of the allele that predominates in a population depends on both the mean and the variance in fitness.Consequently, if two alleles have the same (arithmetic) mean fitness through time, the allele that ‘wins’ is the one with the smaller variance in fitness.Thus, it is advantageous for alleles to avoid large fluctuations in fitness.When there are no fluctuations in fitness through time (constant environments),the geometric mean fitnesscollapses to the arithmetic mean fitness (Orr, 2009).

In constant environments natural selection leads to each individual organism maximizing its expected number of descendants left far in the future. If there are no environmental fluctuations, population fitness ismaximized and measured by the arithmetic mean number of surviving descendants. In evolutionary computation, the Genetic Algorithm (GA) is based on the “survival of the fittest” principle and simulates natural evolution on computer systems to solve complex problems. Individuals are selected and reproduced according to a fitness performance criterion. The fitter the individual, the higher are its chances to produce offspring. Since the process is biased towards the regions of the solution space which enclose the fittest individuals, the evolving population gradually loses diversity and converges. After a population has converged, it is very difficult to readapt to a new optimum when the environment changes (Cobb and Grefenstette, 1993; Simões and Costa, 2002; Bui et al., 2005). Thus, premature convergence is a problem for the GA as it gradually loses its exploratory ability during the evolutionary process under an oversimplified “survival of the fittest” principle.

The standard criterion for evaluating the fitness of genotypes in stochastic environments, however, is the geometric mean of the growth rates (geometric mean fitness) (Dempster, 1955; Cohen, 1966; Lewontin and Cohen, 1969; Frank and Slatkin, 1990; Yoshimura and Clark, 1991; Yoshimura and Jansen, 1996; Hopper, 1999; Simons, 2009; Yoshimura et al., 2009). Geometric mean fitness is a concept widely used in ecology and evolutionary biology to understand persistence of populations in fluctuating environments (Lewontin and Cohen, 1969; Levins, 1969; Gillespie, 1974a; b; Kuno, 1981; Yoshimura and Jansen, 1996; Jansen and Yoshimura, 1998). In variable environments, the geometric mean fitness is always lower than the arithmetic mean. In fluctuating environments it may be optimal for different individuals of the same genotype to take different actions to spread the risk. Risk spreading polymorphism makes sense only for groups - by definition an individual cannot be polymorphic. The fitness of the genotype is determined by the, perhaps complementary, actions of all individuals of the genotype, and the best action of an individual depends on the states and actions of other population members (McNamara et al., 1995; McNamara, 1998; Török et al., 2004).Risk-sensitive reproductive strategies may reduce the average (arithmetic mean) of individual reproductive output, while yet maximizing the populational geometric mean; this trade-off in terms of average reproduction is 'bet hedging' (Gillespie, 1973; 1974a; Slatkin, 1974; Seger and Brockmann, 1987; Philippi and Seger, 1989). A rigorous definition of bet-hedging includes lower expected arithmetic mean fitness, as well as greater expected geometric mean fitness (Seger and Brockmann, 1987; Simons, 2009). Bet-hedging involves a trade-off between the mean and variance of fitness. If the environment varies temporally, phenotypes with low variances of fitness may be favored over alternatives with higher variances and higher mean fitnesses (Philippi and Seger, 1989). This reduction in among-generation variation in fitness (yielding a higher geometric mean) forms the basis of bet-hedging theory: bet-hedgers, reducing variance in fitness, don’t necessarily do best all the time, but they perform most consistently and are therefore favored by selection (Cohen, 1966; Roff, 1992). Thus, in stochastic environments, individual fitness maximization regimes are replaced by population-level fitness maximization strategies that yield suboptimal fitness results for individuals (Cohen, 1966; Ellner, 1986; McNamara 1995; 1998; McNamara et al., 1995; Yoshimura and Jansen, 1996).

Intriguingly, evidence for the evolutionary merit of reproductive restraint comes from multipletheoretical and experimental studies in various taxa (Gilpin, 1975; Nathanson, 1975; Wade, 1980; Walker, 1984; Rand et al., 1995; Savill and Hogeweg, 1998; Sober and Wilson, 1998; Boots and Sasaki, 2000; Haraguchi and Sasaki, 2000; Rauch et al., 2002; 2003; Werfel and Bar-Yam, 2004; Borrello, 2012). Reproductive prudence of cells arose as a necessary prerequisite of multicellularity (Buss, 1987; Maynard Smith and Szathmáry, 1995; Frank and Nowak, 2004). The overarching issue of this behavior is the prudent and sustainable use of resources. Resource availability is stochastic. And bet-hedging is a way to cope with this stochasticity.The tragedy of the commons (a situation where individual competition reduces the resource over which individuals compete, resulting in lower overall fitness for all members of a group or population) provides a useful analogy allowing to understand why shared resources tend to become overexploited (Hardin, 1968). The tragedy of the commons analogy has become increasingly used to explain why, in principle, selfish individuals in a multitude of animal and plant populations evolved means to avoid the overexploitation of limited collective resources (Frank, 1995; Gersani et al. 2001; Falster and Westoby, 2003; Foster, 2004; Wenseleers and Ratnieks, 2004; Rankin and López-Sepulcre, 2005; Kerr et al., 2006; Rankin and Kokko, 2006). Factors such as high relatedness in social groups (Wenseleers and Ratnieks, 2004), diminishing returns (Foster, 2004), policing and repression of competition (Frank, 1995; 1996b; Hartmann et al., 2003; Ratnieks and Wenseleers, 2005), pleiotropy (Foster et al., 2004) or control of population density (Suzuki and Akiyama, 2005; Hauert et al., 2006; Kokko and Rankin, 2006; Rankin, 2007; Frank, 2010) have been argued to constrain the evolution of overexploitative behavior, and thus reduce the potential for a tragedy of the commons to arise in such populations.

According to Fisher (1930), as noted in his treatment of sex ratio evolution, some evolutionary phenomena require the consideration of offspring counts over more than one generation. Thus, there are short-term and long-term aspects to fitness (Beatty and Finsen, 1989; Sober, 2001). This distinction is not trivial. In fact, short-term reproductive success may threaten the evolutionary success of a geno-/phenotype, by placing too great a demand on available resources (Beatty and Finsen, 1989). Long-term concepts of fitness have been forwarded by Thoday (1953; 1958) and Cooper (1984). Thoday suggested that fitness should be defined as the probability of leaving descendants in the very long run; he proposed 108 years as an appropriate time scale. Cooper argued that the “expected time to extinction” of a particular genotypic or phenotypic subpopulation may be the adequate measure of fitness.

In this context, a highly contentious issue revolved around the obvious reproductive restraint of predators. The classical model of predator-prey dynamics, the Lotka-Volterra equation, predicts that under most conditions predator populations, like prey populations, go through a series of oscillations between feast and famine, at each cycle approaching the brink of extinction (Holland, 1995; Mitteldorf et al., 2002; Mitteldorf, 2010). The “prudent predator” concept (Slobodkin, 1961; 1974; Goodnight et al., 2008) has revealed evolutionary outcomes of predator-prey interactions and provided evolutionary mechanisms to resolve the tragedy of the commons dilemma. Original studies by Wynne-Edwards (1962), suggesting that predators executed restraint in reproduction in order to avoid overexploitation of resources were dismissed as inadequate since any such restraint appears to require group selection (Maynard-Smith, 1964; Williams, 1966a). However, there is theoretical justification for a self-consistent limitation of reproduction by predators (Mitteldorf et al., 2002). The rapidly reproducing types modify their local environment, depleting resources in a way that is detrimental to their survival, but this environmental modification and its feedback to population growth may require many generations (Mitteldorf et al., 2002; Goodnight et al., 2008). When a predator evolves, the evolutionarily stable type is outcompeted in the short term by seemingly fitter mutants, which have the highest numbers of offspring for many generations but go extinct in the long term (e.g., after ~200 generations) (Rauch et al., 2002; 2003). The benefit of restraint is that better resource management may prolong the persistence of the group. Thus, the prudent predator attenuates the oscillations predicted by the Lotka-Volterra equation and stabilizes populations (Holland, 1995; Mitteldorf, 2006; 2010). To my knowledge, so far no formal mathematical treatment of the “prudent predator” concept with regard to the trade-off between mean and variance of fitness has been provided. Without going into the mathematical details, the graphical representation of the issue (e.g. Mitteldorf et al., 2002) makes it highly plausible that the “prudent predator” is yet another animal bet-hedging behavior (i.e. reducing variance in fitness) in response to environmental uncertainty.

Darwinian evolution favors genotypes with high replication rates, a process called ‘survival of the fittest’. However, knowing the replication rate of each individual genotype may not suffice to predict the eventual survivor (Wilke et al., 2001). According to quasi-species theory, selection favors the cloud of genotypes, interconnected by mutation, whose average replication rate is highest (Eigen, 1971; Eigen and Schuster, 1979; Schuster and Swetina, 1988; Eigen et al., 1989; Nowak, 1992).

A related issue concerns the population fitness of sexual and asexual populations. Agrawal (2002) pointed to a key difference in the equilibrium genetic load (the reduction in fitness of a population of interest relative to a population composed solely of the most-fit genotype) in sexual and asexual populations under constant and fitness-dependent mutation rates. If the mutation rate is constant (in constant environments, dependent on constant, accidental mutations), then sexual and asexual populations are expected to have the same genetic load at mutation-selection equilibrium. In contrast, if the mutation rate is fitness-dependent (as in fluctuating environments), the expected fitness in an asexual population at equilibrium will be that of the most-fit genotype, while in a sexual population, the equilibrium fitness will be less than that of the most-fit genotype, perhaps much less. And yet sexual reproduction, the commander in chief of bet-hedging strategy prevails in fluctuating environments.

It doesn’t make much sense to play a lottery with one lottery ticket and one individual. Raffles are played with many tickets and several players. If the goal is to have at least a couple of winners (i.e. individuals that are able to reproduce), it is worthwhile to buy as many lottery tickets as possible and to cover all bases (avoiding to put all eggs in one basket). The larger the size of the population, the better are its chances to survive the many iterations of the evolutionary game. In a variable world it is probably wise to take precautions against changes. At the next iteration of the game, the cards are being reshuffled and the winners of today may be the losers of tomorrow. Thus stochasticity and iteration of the game turn the individual-level selection pattern into a population-level selection regime.The evolutionary stochasticity-selection balance is dynamic.That means, a bet-hedging strategy is the ESS in stochastic environments. In relatively stable environments, the bet-hedging strategies have a lower fitness advantage. However, the evolutionary success of sexual reproduction implies that stable environments are rather the exception than the rule.

In the dualism of the evolutionary world the welfare of individuals and populations are inseparably connected.Clearly, the individual is the target of selection. But evolution, like statistics, cannot be done on n=1. Variation (the result of stochasticity) at the level of populations/groups is required for selection. And variation can only be feature of populations. Thus, populations/groups are the level of selection. This applies for groups of genes, cells, or organisms. When the individual would be the sole target of selection, selection would not go for variation but for individual perfection. In the face of pervasive environmental change and unpredictability not the fittest individual (which always means the fittest at the prevailing conditions) but variation of less fit individuals is the survival insurance of a population.

19.3 The selectionist-neutralist controversy

Neutral models characterize evolutionary or ecological patterns expected in the absence of specific causal processes, such as natural selection or ecological interactions. The discovery in mid-20thcentury of extensive variation in protein and DNA sequence, both within and between species, stimulated Kimura (1968) to propose that most of this variation has no effect on fitness, on the grounds that it could not all be maintained by selection (Barton and Partridge, 2000). Kimura (1968; 1983; 1991) argued that most mutagenic changes are selectively neutral or near-neutral and put forward the neutral theory of molecular evolution. “…the neutral theory claims that the overwhelming majority of evolutionary changes at the molecular level are not caused by Darwinian natural selection acting on advantageous mutants, but by random fixation of selectively neutral or very nearly neutral mutants through the cumulative effect of sampling drift (due to finite population number) under continued input of new mutations” (Kimura, 1991). At face value, confirmation of various predictions of the theory provided evidence for its correctness:

• in protein sequences, conservative changes—substitutions of amino acids that have similar biochemical properties and are therefore less likely to affect the function of a protein—occur much more frequently than radical changes.

• synonymous base substitutions of the third nucleotide in a triplet (i.e., those that do not cause amino acid changes) occur almost always at a much higher rate than nonsynonymous substitutions.

• noncoding sequences, such as introns, evolve at a high rate similar to that of synonymous sites.

• entirely untranslated pseudogenes, or dead genes, evolve at a high rate, and this rate is about the same in three-codon positions.

Since the formulation of the Neutral Theory of Molecular Evolution by Mooto Kimura (Kimura, 1968; 1983; King and Jukes, 1969), the neutralist–selectionist debate (Beatty, 1984; Gillespie, 1991; Millstein, 2002; Nei, 2005) has been around and is still on the agenda of evolutionary biologists.The debate was articulated around the relative importance of genetic drift and selection (Hey, 1999) with cases made both for selection (Gillespie, 2001; Hahn, 2008) and nonadaptive processes (Lynch, 2006; 2007a;b; Ellegren, 2009).However, the neutral theory is now a theory in retreat (Lewontin, 1974; Chamary et al., 2006; Begun et al., 2007; Hahn, 2008). First of all, the theory is hypothetico-deductive and rooted in population genetics that as Lewontin (1974, p. 267) put it: “… is not an empirically sufficient theory”. That the theory provides accurate predictions may be owed to correlation and not causation and does not vindicate the basic assumptions. Most importantly, the theory advances a static approach to selection implying that even in the face of pervasive ecological and genetic change the selectively neutral status of a mutation is unchanged. Positive and negative selection have been shown to be more extensive than predicted by the neutral theory (Fay et al., 2001; 2002; Smith and Eyre-Walker, 2002; Clark et al., 2003). Recent studies based on DNA sequence data from large numbers of genes have increasingly suggested the prevalence of adaptive evolution in coding (Sawyer SA et al., 2003; Bierne and Eyre-Walker, 2004; Bustamante et al., 2005; Macpherson et al., 2007; Shapiro et al., 2007; Kosiol et al., 2008; McVicker et al., 2009; Nielsen et al., 2009; Sella et al., 2009; Halligan et al., 2010) as well as noncoding (Kohn et al., 2004; Andolfatto, 2005; Haddrill et al., 2008; Sethupathy et al., 2008; McVicker et al., 2009; Sella et al., 2009) regions to an extent which is incompatible with the Neutral Theory of Molecular Evolution.

According to the Neutral Theory, mutation rate and substitution rate should be the same (Kimura, 1968). The fate of neutral mutations in a population is determined solely by genetic drift and, unless there is strong linkage to sites under selection, the fixation rate depends only on the rate at which the mutations are generated (Kimura, 1968). At a given population size, the fixation rates of beneficial and deleterious mutations are higher and lower, respectively, than the neutral fixation rate. However, pedigree studies suggest that mutation rates are much higher than substitution rates (Bendall et al., 1996; Howell et al., 1996; 2003; Mumm et al., 1997; Parsons et al., 1997; Sigurdardóttir et al., 2000; Ho et al., 2005; 2008; 2011; Santos et al., 2005). For instance, the mutation rate estimated from pedigrees of humans is a hundredfold higher than the substitution rate for the primate mitochondrial DNA control region (Sigurdardóttir et al., 2000). Highestimates of mtDNA mutation rates have since been obtained in pedigree studies of other organisms, including Caenorhabditis (Denver et al., 2000), Drosophila (Haag-Liautard et al., 2008) and Adélie penguins (Millar et al., 2008).

As has been emphasized by A. Wagner (2005c; 2008b), neutrality is not an essential feature of a mutation. That is, a once neutral mutation may cause phenotypic effects in a changed environment or genetic background. He argued that most, if not all, neutral mutations are of this sort, and that the essentialist notion of neutrality should be abandoned. Thus, two opposing views on the forces dominating organismal evolution, natural selection and random drift are reconciled: neutral mutations occur and are especially abundant in robust systems, but they do not remain neutral indefinitely, and eventually become visible to natural selection, where some of them lead to evolutionary innovations (Wagner, 2012).

Mayr denied that random genetic drift is an evolutionary mechanism. In 2001, Mayr wrote: “Molecular genetics has found that mutations frequently occur in which the new allele produces no change in the fitness of the phenotype. Kimura (1983) has called the occurrence of such mutations ‘neutral evolution’, and other authors have referred to it as non-Darwinian evolution. Both terms are misleading. Evolution involves the fitness of individuals and populations, not of genes. When a genotype, favored by selection, carries along as hitchhikers a few newly arisen and strictly neutral alleles, it has no influence on evolution. This may be called ‘evolutionary noise’, but it is not evolution”. However, Stebbins and Ayala (1981) argued that the “selectionist” and the “neutralist” views of molecular evolution are competing hypotheses within the framework of the synthetic theory of evolution.

The Neutral Theory was formulated under the impression ofextensive variation in protein and DNA sequence, questioning the pervasive action of selection.Importantly, it was only assumed (arguing that this level of variation could not be tolerated under the pressure of natural selection) but never shown that the extensive variation in protein and DNA sequence is in fact selectively neutral. Within the framework of the stochasticity-selection dualism, the degeneracy of molecular processes (see chapter 13), and the pre-selection of mutations in the SMSCs, extensive non-neutral variation is recognized as the genetic signature of the bet-hedging strategy in response to environmental unpredictability.

19.4 The adaptationism controversy

There is a longstanding controversy about adaptationism,the “programme based on the faith in the power of natural selection as an optimizing agent” (Gould and Lewontin, 1979; Godfrey-Smith, 2001; Lewens, 2009). Adaptationism seeks to identify adaptations and the specific selective forces that drove their evolution. But many aspects of genomic, cellular, and developmental evolution can only be understood by invoking a lower level of adaptive involvement (Kimura, 1983; Lynch, 2007a; b). Dobzhansky (1950), in a seminal statement on adaptation to diverse environments, wrote ‘Changeable environments put the highest premium on versatility rather than on perfection in adaptation’. In particular, Gould and Lewontin (1979) argued that adaptationists often use inappropriate evidentiary standards for identifying adaptations and their functions, analogous to Rudyard Kipling’s Just So Stories (outlandish explanations for questions such as how the elephant got its trunk). Of course, Gould and Lewontin freely admitted that adaptation by natural selection has been, and still is, a powerful force in shaping the phenotypic traits of organisms. The question they wished to raise was: what kind of evidence is necessary to support the hypothesis that a trait is an adaptation formed by natural selection and what evidence could point towards some other cause of the current state of the trait in question (Pigliucci and Kaplan, 2000)? The past 50 years have seen an increased recognition of sluggish evolution and failures to adapt (Futuyma, 2010). According to Lewin (1980), the existence of constraints meant that natural selection was involved at only one stage of the evolutionary process and thus was not the only essential factor in evolution. It has been speculated that these additional forces may been forces like drift, gene flow, epigenetic inheritance, pleiotropy, developmental, structural and phylogenetic constraints (Gould and Lewontin, 1979; Amundson, 1994; 2001; Futuyma, 2010). Natural selection apart, all evolutionary processes are random with respect to adaptation, and therefore tend to degrade it (Barton and Partridge, 2000).In the reformist counterparadigm, one invokes “chance”, “constraints”, and “history” to explain imperfections: some features don’t turn out perfectly, due to statistical noise, in-built limitations, and so on; some features, due to “historical contingency”, are side-effects or vestiges. Selection still governs evolution, as Darwin said, but there are “limits to selection” (Barton and Partridge, 2000).

With the recognition of the stochasticity-selection dualism of evolution these limits have a unified conceptual basis. Stochasticity contributes to maladaptation or limits adaptation (Travisano et al., 1995; Hereford, 2009; Lenormand et al., 2009). Bet-hedging is driving suboptimal adaptive strategies. Dempster (1955) introduced the model in which temporal fluctuations in reproductive success for competing genotypes favor the genotype with the highest geometric-mean reproductive success. The geometric mean criterion implies that evolution tends to minimize the variance (or, more precisely, to optimize the trade-off between mean and variance) in intergenerational per capita reproduction (Gillespie, 1977; Frank and Slatkin, 1990).Almost any individual life history or behavioral adaptation may be affected by environmental stochasticity. Environmental fluctuations preclude optimal adaptation to any single environment (Levins, 1968). Crow (1958) showed that the phenotypic variance of relative fitness places a limit on the evolution of fitness.

19.5 Paradox of viability

Dobshanzky (1937) noted the “paradox of viability”. On the one hand, each species needs to have variation present in order to adapt to environmental changes. On the other hand, the particular variations that might prove adaptive in the future are very likely maladaptive at present. “Evolutionary plasticity can be purchased only at the ruthlessly dear price of continuously sacrificing some individuals to death from unfavorable mutations.” (Dobzhansky, 1937, p. 127). Importantly, the viability costs of mutations are paid during the sexual selection cascades. By subjecting the mutated gametes to a rigorous quality selection, the SMSC succeed to reduce the viability costsof mutations and to keep the investment in mutagenic-selective “trial-and-error cycles” and its associated viability risk at a low economic rate.

19.6 Mystery of low heritability

Heritabilities are often less than 0.5. Of traits in Drosophila and other non-domestic animals classified as life-history, behavior, physiology and morphology, only morphological traits in animals, excluding Drosophila, average about 0.5; all other categories are less (Roff, 1997; Franklin and Frankham, 1998). This is particularly important for traits closely related to reproductive fitness where heritabilities are typically 10-20% (Roff, 1997). Moreover, field data demonstrate that heritability decreases under stressful conditions, at least for morphological traits (Charmantier and Garant, 2005). Genome-wide association studies (GWAS) have led to the identification of >1,200 loci harboring genetic variants associated with >165 common human diseases and traits, revealing previously unknown roles for scores of biological pathways (Manolio et al., 2008; Hirschhorn, 2009; Eichler et al., 2010; Lander, 2011; Zuk et al., 2012). However, early GWAS were puzzling because they appeared to explain only a small proportion of the “heritability” of the traits. With larger GWAS, the proportion of heritability apparently explained has grown (to 20–30% in some well-studied cases and >50% in a few), but, for most traits, the majority of the heritability remains unexplained (Lander, 2011).

Of course the expectation that heritabilities should be much higher is owed to the anticipation that natural selection and its effects on the (epi)genomeand thus heritability play a large role in evolution. Bet-hedging as response to environmental stochasticity and its manifold effects on (epi)genotypic variation and phenotypic plasticity dissipates the action of selection (Wilson et al., 2006; Simons and Johnston, 2006; Simons, 2011). Environmental variance reduces heritabilities, both through an increase in the environmental component of variance and a reduction in the additive genetic variance (Simons and Roff, 1994), and a recent meta-analysis shows that heritabilities are generally reduced under unfavourable conditions (Charmantier and Garant, 2005). Another evidence that stochasticity as second organizing principle in evolution counteracts selection.

19.7 The Cambrian explosion

For reasons that have remained unclear, almost all of the modern animal phyla arrive in the fossil record during a geologically short period of about twenty million years, starting about 540-545 million years ago (Valentine, 1995; Valentine et al., 1999; Baker, 2006; Marshall, 2006). The so-called Cambrian explosion (CE) with its dramatic increase in both disparity and diversity has been a unique and troubling anomaly in the history of life. Speculation about the rise of animals has run the gamut from purely intrinsic biological causes (Bengtson and Morris, 1992; Parker AR, 1998; Smith and Peterson, 2002; Peterson et al., 2005; Baker, 2006) to extrinsic triggers (Derry et al., 1994; Canfield and Teske, 1996; Hoffman et al., 1998; Brasier and Lindsay, 2001; Kirschvink and Raub, 2003; Horvath, 2003; Squire et al., 2006) or some combination of both (Peterson et al., 2005). Several factors that may be prerequisites for the explosion have been identified, necessary but not sufficient. By the start of the Cambrian, the large supercontinent Gondwana, comprising all land on Earth, was breaking up into smaller land masses. This increased the area of continental shelf, produced shallow seas, thereby also expanding the diversity of environmental niches in which animals could specialize and speciate (Veevers, 2004; Meert and Lieberman, 2008). Rapid change following a long period of quasi-stasis (due to “Snowball Earth”?) (Hoffman et al., 1998; Hyde et al., 2000) suggests the existence of a triggering or enabling event, either external to organisms or incorporated in organisms (Butterfield, 2007). The time of onset is constrained by the evolution of the environment, whereas its duration appears to be controlled primarily by rates of developmental innovation (Marshall, 2006). The real question is what drives morphological evolution (as opposed to what merely allows it, or might hold it back – the focus of most Ediacaran⁄Cambrian explosion models) (Butterfield, 2007). It has been noted that the oxygen content of the atmosphere was slowly rising (Thomas, 1997; Fike et al., 2006; Canfield et al., 2007). This was likely an enabler, and possibly a trigger (Nursall, 1959; Berkner and Marshall, 1965; Canfield and Teske, 1996; Ohno, 1997b; Canfield et al., 2007). Most animals require molecular oxygen in order to produce their energy, and this has led to the widespread presumption that a rise in atmospheric oxygen was the essential precursor to the evolution of animals (Runnegar, 1991; Catling et al., 2005; Shields-Zhou and Och, 2011). Multiple lines of evidence from evolutionary biology (Falkowski et al., 2005; Acquisti et al., 2007), geochemistry (Bekker et al., 2004), and systems biology (Raymond and Segre, 2006) build a compelling case for a central role of O2 in the evolution of complex multicellular life on earth (Catling et al., 2005; Falkowski, 2006; Raymond and Segre, 2006; Thannickal, 2009). Both a surge in seawater calcium and phosphate have been implied as capacitating environmental factors (Cook, 1992; Brennan et al., 2004; Porter, 2007; Fernàndez-Busquets et al., 2009).

So far, explanations for the acceleration of the evolutionary tempo during CE center on the invention of new trophic capacities, whether predation (Vermeij, 1990; Bengtson, 2002) or the related cropping (Stanley, 1973; 1976). Many of these theories emphasize the role of predation, specifically in an effort to explain the massive skeletonization event that characterizes the fossil record of the explosion (Vermeij, 1990). All of these have a common thread of coevolution, escalation, or arms races (Vermeij, 1987; 2004; Butterfield, 2007).

The animal phyla are divided into three groups. First are the sponges (Porifera) that do not have organized cell layers. Second are the diploblasts, which have two primary cell layers: an outer ectodermal layer and an inner endodermal layer. The Cnidaria (corals and jellyfish) and the jellyfish-like group, the Ctenophores, are the only diploblastic phyla. All of the remaining animal phyla are triploblasts, also collectively referred to as the bilateria. These grow from three primary cell layers: the outer ectoderm, the intermediate mesoderm (from which our skeleton and most of our muscles are derived), and the inner endoderm, which includes the gut. Over 99% of all living animals are triploblasts (Marshall, 2006). Both the geological fossil, ontogenetic and genetic records support the Ediacaran (c. 585–542 Myr ago) emergence of triploblasts (Aris-Brosou and Yang, 2002; 2003; Peterson et al., 2008; Rokas, 2008; Xiao and Laflamme, 2008; Erwin, 2009).A fact that has received relatively little attention in the attempts to explain the causes of the CE: it is a phenomenon restricted to triploblasts and thus its causes should be sought in some evolutionary innovation unique to these organisms. Hence, cellular differentiation that was proposed as a candidate new technology associated with CE (Phoenix, 2009), is an improbable cause of CE since it is a property of all multicellular organisms. The fact that the developmental toolkit was unequivocally established in the last common protostome-deuterostome ancestor (Valentine et al., 1996; Martindale et al., 2004, Martindale, 2005; Couso, 2009; Erwin, 2009) and that the last common protostome-deuterostome ancestor probably lived prior to 555 million years ago strongly suggest that developmental innovation may have been necessary but cannot be a sufficient cause of the main Cambrian radiation. The developmental innovations were a precondition to the later events and may explain the extraordinary breadth of the radiation, but not the triggering of the event itself (Erwin, 2005).

According to S. Ohno: “Assuming a spontaneous mutation rate to be a generous 10-9 per base pair per year and also assuming no negative interference by natural selection, it still takes 10 million years to undergo 1% change in DNA base sequences. It follows that 6-10 million years in the evolutionary time scale is but a blink of an eye. The Cambrian explosion denoting the almost simultaneous emergence of nearly all the extant phyla of the Kingdom Animalia within the time span of 6-10 million years can’t possibly be explained by mutational divergence of individual gene functions” (Ohno, 1996).

Let’s look at the ingredients for the Cambrian stew that we have assembled so far. We have an abiotic enabler, and possibly a trigger, the slowly rising oxygen content of the atmosphere. And we have the bilateria. What distinguishes the bilateria from the non-bilateria? Generally, in basal metazoa (Porifera, Ctenophora, Placozoa, Cnidaria) the adult body is itself the reproductive unit in which germ cells arise from a somatic stem cell population. Bilateria have a germline that is segregated from the soma (see chapter 6.2). The segregation of the germline underlies the germ-soma conflict (Heininger, 2012) and promoted the functional specialization in differentiated tissues (Simpson, 2012). But how can oxygen and the germ-soma conflict be brought together into a coherent process underlying the CE? Intriguingly, the missing link is sexual reproduction. Current evidence suggests that sex has a single evolutionary origin and was present in the last common ancestor of eukaryotes (Dacks and Roger, 1999). Sexual reproduction and eukaryotes probably evolved together (Cavalier-Smith, 2002), about 2.0 to 3.5 billion years ago (Miyamoto and Fitch, 1996; Gu, 1997). Despite their sexual reproduction, at least 1.5 billion years before the CE, the overarching pattern of pre-Ediacaran eukaryotes is one of minimal morphological diversity and profound evolutionary stasis (Butterfield, 2007). In pre-Ediacaran time, sex hardly was a dynamic evolutionary motor. But during the rise in atmospheric oxygen, sexual reproduction may have “learned” to use oxidative stress as a tool for evolutionary innovation that gave the coevolutionary germ-soma conflict an unprecedented dynamic. Coevolving sytems have significantly higher levels of heterozygosity and allelic diversity (Buckling and Rainy, 2002a; Duncan and Little, 2007; Duffy MA et al., 2008; Koskella and Lively, 2009; Bérénos et al., 2011) that correlates with their evolutionary potential (Fisher, 1930; Falk and Holsinger, 1991; Frankham, 1995; Falconer and Mackay, 1996; Franklin and Frankham, 1998; Reed and Frankham, 2003; Johnson et al., 2006; Leimu et al., 2006). Both the coevolutionary engine of the germ-soma conflict in bilateria and oxidative stress, the effector of mutagenesis and cellular selection and ultimate tool of the SMSCs (see chapter 10) gave animal evolution a kick-start during the Cambrian radiation. With these processes, sexual reproduction succeeded to accelerate evolution substantially, creating in the last 600 million years a wealth of taxa and species (Stanley, 1975). Thus, the CE was enabled by a multifactorial process. Abiotic and biotic factors that originated some 1.5 billion years apart were integrated into a highly efficient evolutionary vehicle.

19.8 The Darwinian-Lamarckian evolutioncontroversy

In my opinion, the greatest error which I have committed, has been not allowing sufficient weight to the direct action of the environment, for example, food and climate, independently of natural selection. When I wrote The Origin, and for some years afterwards, I could find little good evidence of the direct action of the environment; now there is a large body of evidence.

Charles Darwin (1876) in a letter to Moritz Wagner

Natural selection acts upon phenotypic variation of individuals that is determined by their (epi)genetic constitution, but also shaped by their specific environment (Jablonka and Lamb, 2005), developmental processes (Müller, 2007), and stochastic events (Huang, 2012). The process of evolution is thus a result of complex interactions between various intrinsic and extrinsic factors (Pigliucci, 2009).According to Bateson (2012), Lloyd Morgan (1896) “suggested that if a group of organisms respond adaptively to a change in environmental conditions, the modification will recur generation after generation in the changed conditions, but the modification will not be inherited. However, any variation in the ease of expression of the modified character which is due to genetic differences is liable to act in favor of those individuals that express the character most readily. As a consequence, an inherited disposition to express the modifications in question will tend to evolve. The longer the evolutionary process continues, the more marked will be such a disposition. Plastic modification within individuals might lead the process, and a change in genes that influence the character would follow; one paves the way for the other”.

In contrast to Darwin, Lamarck thought that evolution proceeds by the inheritance of acquired traits. Acquired traits have been gained by means of phenotypic plasticity during an organism’s lifetime in response to environmental perturbation. Thus, Lamarck advocated the transgenerational transmission of learned characters. Weismann (1891), based on his concept of the germ-plasm, rejected the inheritance of acquired characters on the grounds that changes to the soma cannot produce the kind of changes to the germ-plasm that would result in the altered character being transmitted to subsequent generations (Haig, 2007). Thus, in the tradition of the Modern Synthesis it is generally considered that direct and rapid feedback from the environment to the germline cannot happen.

During recent years a lively controversy concerning the relative significance of Darwinian and Lamarckian modes of evolution took place (Hoenigsberg, 2002;Jablonka and Lamb, 2005; 2008b; Haig, 2007; Shapiro and Sober, 2007; West-Eberhard, 2007; Koonin and Wolf, 2009; 2012; Feinberg and Irizarry, 2010; Merlin, 2010; Gissis and Jablonka, 2011; Dickins and Rahman, 2012; Koonin, 2012).The question is, formulated in molecular terms: can (epi)mutations in genes of somatic cells of an animal possibly be transmitted to the genes of germ cells and passed on to offspring of future generations?

Plants and benthic aquatic animals (e.g. Porifera, Cnidaria) have no sequestered germline. The germline can form at any time in the organism’s life from multipotent stem cells. Accordingly, somatically acquired mutationscan be inherited in these organisms by asexual and sexual reproduction and this has been discussed in chapter 15.2. In unicellular prokaryotic and eukaryotic microorganisms, the same cell is both soma and germline and, of course, mutations that have been acquired can be transmitted to the next generation. Thus, Weismann’s dogma only can apply to unitary organisms, mobile animals. However, particularly in discussions concerning the Darwinian-Lamarckian evolution controversy this distinction has been blurred and, particularly by proponents of Lamarckian-type inheritance, processes in bacteria and plants are taken to vindicate Lamarckian modes of evolution (Koonin and Wolf, 2009; 2012; Koonin, 2012).

Epigenetics was first suggested by Jablonka and Lamb (1995) to play a role in evolution through Lamarckian inheritance, that is, direct modification of the genome by the environment, which is then transmitted transgenerationally. As defined by Jablonka et al. (1998) “Acquired characters are the outcome of instructive processes, such as those seen in embryonic induction, transcriptional regulation, and learning, all of which involve highly specific and usually adaptive responses to factors external to the responding system.”Feinberg and Irizarry (2010) questioned this concept, proposing a non-Lamarckian theory for a role of epigenetics in evolution. As direct evidence for stochastic epigenetic variation, they referred to (i) the high variability of DNA-methylated regions in mouse and human associated with development and morphogenesis and (ii) a heritable genetic mechanism for variable methylation, namely the loss or gain of CpG dinucleotides over evolutionary time.

I think the discussions and insights presented in various chaptersof this work provided the insights on the basis of which we can revisit the Darwinian-Lamarckian controversy:

Chapter 8.1: Natural selection is only the proverbial tip of the iceberg of Darwinian selection processes. On all levels of development and sexual reproduction, cellular competition, fuelled by variation created by stochastic processes, results in cellular selection.

Chapter 18.1.2: The myriadof molecular processes that determine an individual’s phenotype have no inbuilt sensor that informs about the selective value of an action. These processes are totally blind with regard to their fitness-relevance. In addition, the degeneracy, although a source of robustness and evolvability, is so pervasive that no fitness-relevant result can be pinpointed to a certain process.

Chapters12 and 17: Evolution is a cybernetic process. Reinforcement learning enables organisms to learn from past “experiences” (in the sense that survival/reproduction is the evolutionary feedback loopenabling reinforcement learning),making “educated guesses” to face new, unpredictable challenges. Bet-hedging is the ESS in a capricious environment.

Chapter 16: Transgenerational epigenetic inheritance is an evolutionary reality. Germ granules are the carriers of transgenerational messages.

Chapter 17: Apart from selection, stochasticity is an organizing principle of evolution. Various processes, perfectly designed as they may appear have inherently stochastic elements.

Environmental stressors are the most relevant triggers of transgenerational information transfer. The metabolic and energetic stress is sensed by the mitochondria and can be relayed to the nucleus (see chapter 21) where it elicits epigenetic and genetic changes (see chapter 10.3). Close contact between the somatic gonads and germline cells are used to transmit these messages via autocrine, paracrine and endocrine messengers and ncRNAs. These messengers may induce both changes of the gametogenic mutagenesis-selection equilibrium and more directed sequence changesdue to transcription-associated mutagenesis and epigenetic changesaffecting gene expression. Most importantly, the several genetic and epigenetic tools have a constitutively stochastic character that allows a large sampling of the (epi)genetic search space and degeneracy that may yield functionally equivalent results. Selective sampling from this quasispecies cloud makes the process Darwinian. Missing this Darwinian aspect of the SMSC, created the impression of a Lamarckian-type evolutionary process. With few exceptions (Feinberg and Irizarry, 2010), this aspect has been consistently overlooked in the many discussions surrounding theDarwinian-Lamarckian issue.

The non-randomness of TE insertions (see chapter 12.4) has been taken as evidence for a (quasi)Lamarckian modality of evolution (Martienssen, 2008; Koonin and Wolf, 2009; Zhang Z et al., 2013). That TE insertion is largely irrespective of its adaptive nature can be inferred from circumstantial evidence. It has been estimated that 80% of the spontaneous mutations seen in Drosophila genetics result from TEs (Ashburner, 1992) that constitute 7–8% of the genome (Smith et al., 2007). Do mobile DNA insertions similarly create 80% of evolutionary changes in this species? Without question, they do not (Brookfield, 2004). Those mutations that survive over long periods of evolutionary time are expected to be a very small subsample of newly induced mutations (Kidwell and Lisch, 1997). The most revealing observation is the almost complete absence of fixed sites of mobile DNAs in D. melanogaster (Charlesworth and Langley, 1989). A mobile DNA insertion that created an advantageous phenotype would be expected to spread to fixation in the species by natural selection. This would create a site fixed for the element throughout the species. Such sites are very rare, although they have recently been detected for the S element family in heat shock protein genes (Maside et al., 2002).

The processes of adaptation in a multi-agent system consist of two complementary phases: 1) learning, occurring within each agent's individual lifetime, and 2) evolution, occurring over successive generations of the population. The key conceptual point in a couple of mathematical models was: mutations in genes of somatic cells of an animal can possibly be transmitted to the genes of germ cells and passed on to offspring of future generations. These models suggest that Lamarckian evolution would speed up the adaptation process and promote the early independence of the offspring from the parents by providing more explicit information about the environment in the genotype (Sendhoff and Kreutz, 1999; Arita and Suzuki, 2000). However, by evaluating the characteristics of two different mechanisms of genetic inheritance, i.e Darwinian and Lamarckian, Sasaki and Tokoro (1997) showed that while the Lamarckian mechanism is far more effective than the Darwinian one under static environments, it is found to be unstable and performs quite poorly under dynamic environments. In contrast, even under dynamic environments, a Darwinian population is not only more stable than a Lamarckian one, but also maintains its adaptability with respect to such dynamic environments and only Darwinian populations can adapt to the new world (Sasaki and Tokoro, 1997; 1999; Yamamoto et al., 1999). These models highlight the fundamental differences between the Lamarckian and Darwinian modes of evolution. In a world of uncertainty, organisms have to bet-hedge and cover all bases. Learning can limit the search space and increase the likelihood that some of the generated variations will be useful (Jablonka and Lamb, 2007) (see chapter 12). But unable to estimate the fitness value of (epi-)mutations in a given environment, the features of the Baldwin effect such as stress- and transcription-associated mutagenesis and phenotypic plasticity are nested within the Darwinian principles of quasi-stochastically generated variation and selection.

20. The resource-stress dimensions and ecological window of sexual/asexual reproduction


The key remaining questions of evolutionary biology are more ecological than genetic in nature.

Edward O. Wilson, 1987

The goal of science is to discover patterns of relations among recorded phenomena, so that a few principles can explain a large number of propositions concerning these phenomena.

Ayala (1968c)

Summary

Available resources build the framework for a variety of life history traits and strategies. Both reproduction and stress have energetic and metabolic costs. Threshold traits are phenotypically discrete but are causally related to a continuously distributed condition, called the liability. Individuals above the threshold display one morph and individuals below the threshold display the alternate morph. Sexual/asexual reproduction can be another threshold trait, e.g. in cyclical parthenogens where resource availability and environmental stress intensity are liabilities that often jointly, sometimes individually, determine reproduction strategies. In almost all organisms, reproductive mode is fixed as a result of long-standing selective pressures. Stress intensity (moderate or harsh) and resource availability (r- and K-selected environments) define the four quadrants of the ecological window of sexual/asexual reproduction. The ratio between resource availability and resource investment into mature individuals (that determines the relative costs of a reproductively competent individual, i.e. a lottery ticket in the raffle of life) and the habitat’s stress intensity are species-specific parameters. Importantly, in most animals with a larger-than-microscopic body size, ecological conditions favor sexual reproduction with its pre-selection of viable zygotes. Only by investing little into (epi)mutagenic gametes/zygotes and pre-selecting them before natural selection exerts the final “quality control”, a bet-hedging strategy in the face of unpredictable environments was a viable option in these animals.

20.1 The resource-stress dimensions

MacArthur (1972b) argued that the goal of community ecology (as of all science) is to find general rules. We have now reached the point where the wealth of information can be shaped into a general pattern. Available resources build the framework for a variety of life history traits and strategies. Both reproduction and stress (Parsons, 1994; 2005; 2007; Krebs and Loeschcke, 1994; Davis and Schreck, 1997; Sloman et al., 2000; Fisher, 2007; O’Connor et al., 2011; Johnstone et al., 2012) have energetic and metabolic costs. In a multitude of trade-offs, resources have to be partitioned between these energetic demands (Krebs and Loeschcke, 1994; Wang Y et al., 2004; Huang LH et al., 2007).

One attempt to construct a predictive framework for the relationships between habitat and species characteristics is the habitat templet theory (Southwood, 1977; 1988, Townsend and Hildrew 1994, Korfiatis and Stamou, 1999). This theory assumes that the habitat provides the templet on which evolution forges characteristic morphologies and life history strategies, being at the same time a ‘‘filter’’ resulting in the ecological sorting of the species able to occupy them (Ribera et al., 2001). The influence of habitat is epitomized in a small set of a priori defined axes that are hypothesized to summarize the environmental constraints acting on the populations. These axes are normally related to disturbance and adversity or stress (Southwood, 1977; 1988); habitat predictability and adversity (Greenslade, 1983); or to temporal and spatial heterogeneity, which can be taken as a measure of disturbance and availability of refugia, respectively (Townsend and Hildrew 1994, Korfiatis and Stamou, 1999). An equivalent hypothesis for plants is that of Grime (1977), with ecological strategies (with their respective morphological adaptations) accommodating to stress, disturbance, and biological competition (Grime, 1977; Grime et al., 1997).

Frequently in biology there are so-called threshold traits (Roff, 1996; 1998). These traits are phenotypically discrete but causally related to a continuously distributed condition, called the liability. Both are linked by a threshold of response: individuals above the threshold display one morph and individuals below the threshold display the alternate morph. For instance, reproductive mode such as semelparity/iteroparity, is a threshold trait (Roff, 1996; 1998; Lesica and Young, 2005; Heininger, 2012). I argue that sexual/asexual reproduction can be another threshold trait, e.g. in cyclical parthenogens (Rispe and Pierre, 1998; Serra and King, 1999). Both resource availability and environmental stress intensity are liabilities that often jointly, sometimes individually, are causally related to sexual or asexual reproduction strategies.

One possibility for the persistence of asex is that competition (with sexual species) would be lower, as there are simply fewer species occurring in such highly fluctuating and extreme habitats. Another reason could be that asexuals are able to survive in very low densities over many generations, since they do not need to find a mate to reproduce (Burt, 2000; Van Dijk, 2007, Hörandl, 2008). For all sexual populations, there is a density threshold (that is a proxy of resource availability and is mediated by size and mobility of animals, amongst other biological characteristics) below which the probability of finding a mate is too low to ensure sufficient reproduction for the population to remain viable. In marginal habitats, such as semi-terrestrial ones, conditions may vary widely and asexuals would have the advantage over sexuals.

20.2 The ecological window of asexual and sexual reproduction

…understanding the evolution of sex requires the synthesis of every important process in evolutionary biology.

Otto and Lenormand, 2002

The sexual-asexual reproduction balance is determined by the availability of resources that can be used to sample genotypic search space by trial-and-error reproductive entities. Stress intensity is the other liability that deliminates the framework for the distribution of reproduction strategies. In r-selected environments the organisms with the highest growth rates, i.e. asexually reproducing organisms, are favored. K-selected environments favor competitive organisms. Since Weismann, various authors advocated the idea that the generation of genetic variation underlies the evolutionary rationale of sexual reproduction (Weismann, 1889; 1904; Barton and Charlesworth, 1998; Burt, 2000). Traditionally, the benefits of sex by generating genetic diversity have been considered to become only visible in the long-term, but cannot offset the short-term, two-fold cost of sex compared to asexuality (Maynard Smith, 1978a). However, the SMSC can act at the individual level to create short-term, even immediate, effects and at the population or species level long-term effects. Without any doubt, the shuffling of genes by meiotic recombination and segregation increases genetic variance but the fitness-related consequences of genetic mixing may rather be detrimental than beneficial. Asexually reproducing organisms are able to generate a high amount of genetic variation (Loxdale and Lushai, 2003; Lushai et al., 2003) without having to pay the cost of recombination load (Fisher, 1930; Feldman et al., 1980; Charlesworth and Barton, 1996). In Chlamydomonas, for example, sex resulted in increased fitness variance in sexual relative to asexual lines, and despite an initial decrease in mean fitness in the generations immediately subsequent to sex (recombination load), long-term fitness increased at a faster rate when there were larger increases in variance (Kaltz and Bell, 2002). This work supports some Weismann–Fisher–Muller models of a short-term cost to sex accompanied by a long-term benefit due to increased variance (Colegrave et al., 2002b; Kaltz and Bell, 2002).

Figure 3 represents graphically the various reproduction domains. Moderate stress and K-selected environments define the sexual reproduction domain. Asexual reproduction is prevalent in high stress/K-selected environments (e.g. geographical parthenogenesis) and moderate stress/r-selected environments. Importantly, what is perceived as stressful by the various species is also dependent on their evolutionary history and various constraints. Likewise, whether the domain is r- or K-selected depends on the species-specific ratio between resource availability and investment into mature organisms. Thus environments that are resource-limited for species with a large body mass, may offer abundant resources allowing large population sizes for species with a small body mass. Cyclical parthenogens like Daphnia cycle between asexual and sexual domains, e.g. caused by population density/crowding and/or seasonal cycles.

21. Extension of Cricks central dogma of molecular biology


….it is quite wrong to think of the environment as just a selector of heritable variation. The environment has a dual role in evolution – it does not just select among heritable phenotypic variations, it also induces them.

Jablonka and Lamb, 1995

In essence, Francis Crick’s (1958; 1970) central dogma of molecular biology conveys that the flow of coded information is unidirectional, i.e. from DNA to protein and never in reverse (see Judson, 1979). Crick worded the Central Dogma thus: “This states that once ‘information’ has passed into protein it cannot get out again. In more detail, the transfer of information from nucleic acid to nucleic acid, or from nucleic acid to protein may be possible, but transfer from protein to protein, or from protein to nucleic acid is impossible. Information means here the precise determination of sequence, either of bases in the nucleic acid or of amino acid residues in the protein” (Crick, 1958). Over the years, the central dogma has come under criticism (Shapiro, 2009; 2011). The discovery of reverse transcriptase (Baltimore, 1970; Temin and Mizutani, 1970) and prions (Prusiner, 1991), allegedly questioned Crick’s central dogma (Hunter, 1999). However, as Keyes (1999) pointed out, these discoveries did not refute Crick’s original formulation but clearly violated James Watson’s later interpretation which claimed that the flow of genetic information was ‘unidirectional’ and that ‘RNA never acts as a template for DNA’ (Watson, 1965, p. 298). Genome sequencing has revealed abundant evidence of the importance of reverse transcription in genome evolution (Brosius, 1999; 2003; Betrán et al., 2002). Indeed, over one third of our own genomes comes from DNA copies of RNA (International Human Genome Consortium, 2001). The vision of the Central Dogma has resisted the many different challenges it has faced, including the role of chaperones in protein folding, and phenomena occurring in the transfer of information from DNA to proteins (epigenetic modifications of DNA, RNA interference, RNA splicing and RNA editing) (Morange, 2008). As Wilkins (2012), in response to Shapiro (2011) asserted, Crick’s specific conclusion is still valid: nucleic acid sequence information can be read into proteins or copied into each other (DNA → RNA or RNA → DNA) but protein sequences cannot be reverse-read into nucleic acid sequences. The degeneracy of the genetic code (there is more than one way to specify an amino acid) and of the protein code (very different amino acid sequences can form similar protein structures) clinches the argument.

It has been argued that the central dogma is of the profoundest import, for it illuminates the physiological necessity for the rejection of Lamarckism, the inheritance of acquired traits (Judson, 2004) (at least in organisms without a sequestered germline, see chapter 19.8). However, accepting the universal validity of Crick’s central dogma, we are left with a dilemma. Evidence is accumulating that the environment is able to shape the phenotype of organisms not only by the action of natural selection but also by transgenerational processes (Jablonka and Lamb, 1995; 2005; 2007; Caporale, 1999; 2003a; b; 2009; Radman et al., 1999; Shapiro, 2011). How is this compatible with the central dogma?

Extending an earlier model (Heininger, 2001), I suggest that the phenotype of an organism as result of the coded information transfer DNA/RNA → protein signals back to the (epi)genome via condition-dependent but uncoded signals of the bioenergy-redox-Ca2+ triangle (figure 4). This feedback loop is another manifestation of the dual stochasticity-selection balance.

Tight coordination between the nucleus and mitochondria is required for proper mitochondrial functioning and includes both anterograde (nucleus to mitochondria) and retrograde (mitochondria to nucleus) signals (Allen, 2003; Butow and Avadhani, 2004; Cannino et al., 2007; Pesaresi et al., 2007; Woodson and Chory, 2008). Growth and replication of the nucleus are limited by mitochondrial energy production and thus calorie availability. The regulation of nuclear replication, gene expression and mutagenesis is mediated by mitochondrial energetics. Mitochondria are the cellular sensors for substrate supply, oxygen, bioenergetics, and redox balance (Duchen, 1999; Carrasco et al., 2001; Dzeja et al., 2002; Lane, 2005; Wright et al., 2009). Mitochondria are the key components and conductors orchestrating the cellular stress responses (Biswas et al., 2005; Manoli et al., 2007). The metabolic condition of organisms determines the adaptive stress in a given environment. The vulnerability of organellar genomes to replicative and especially oxidative damage could provide quality-control functions for detecting, correcting or removing cells with long-term or cumulative damage to both genomes, which could otherwise go undetected and compromise the survival of an organism (Wright et al., 2009). In this capacity, mitochondria are long-term redox damage sensors (Wright et al., 2009). Such signalling is presumed to occur via tonic organelle-to-nucleus (retrograde) signaling systems (Butow and Avadhani, 2004; Janssen-Heininger et al., 2008; Woodson and Chory, 2008). These enable the nuclear compartment to respond by setting a new tonic level of nucleus-to-organelle (anterograde) signaling (Butow and Avadhani, 2004; Woodson and Chory, 2008).

Mitochondria take up calcium, and the high capacity mitochondrial calcium uptake pathway provides a mechanism that couples energy demand to increased ATP production through the calcium-dependent upregulation of mitochondrial enzyme activity (Duchen, 1999). Metabolic stress due to stress and maladaptation leads to dysregulation of the Ca2+-energy-redox triangle (Brookes et al., 2004; Camello-Almaraz et al., 2006; Feissner et al., 2009; Peng and Jou, 2010). Both Ca2+ and redox balance mediate the mutagenetic pressure exerted by energy homeostasis. In metazoans, cellular Ca2+ homeostasis is under control of a highly organized reticulum formed by mitochondria and endoplasmic reticulum (Rutter and Rizzuto 2000; Csordás and Hajnóczky, 2009). Mitochondrial Ca2+ signalling has a profound impact on nuclear Ca2+ homeostasis (Faulk et al., 1995). Thus Ca2+ changes are relayed to the nucleus to evoke specific changes in gene expression (Roche and Prentki, 1994; Chawla and Bading, 1998; Hardingham and Bading, 1998) and via regulation of DNA repair (Gafter et al., 1997; Korzets et al., 1999)also affectmutagenesis (Seetharam and Seidman, 1992).

ROS and RNS are the mediators of metabolic stress adaptations in all phyla. Nuclear redox systems are controlled independent of the cytoplasmic counterparts(Go and Jones, 2010). Superoxide anion, H2O2 and NO as relatively inert RONS have long-range signaling properties (Hancock, 1997; Rhee, 1999) and qualify as signal transducers and primary mitochondrial-nuclear messengers for gene expression (Burdon, 1995; Sen and Packer, 1996;Cimino et al., 1997; Lander, 1997; Dalton et al., 1999; Pfeilschifter et al., 2001; Turpaev, 2002; Shapiguzov et al., 2012). Proteins with oxidizable thiols are essential to many functions of cell nuclei, including transcription, chromatin stability, nuclear protein import and export, DNA replication and repair (Go and Jones, 2010). Redox regulation has been shown to play an important role in modulating the DNA binding activity of a number of transcription factors including AP-1, NFκB, HIF-1α, HLF, CREB, PAX, p53, Egr-1, and others (Abate et al., 1990; Xanthoudakis and Curran, 1992; Xanthoudakis et al., 1992; 1994; Huang and Adamson, 1993; Yao et al., 1994; Huang et al., 1996; Sun and Oberley, 1996; Akamatsu et al., 1997; Jayaraman et al., 1997; Ema et al., 1999; Gaiddon et al., 1999; Evans et al., 2000; Lando et al., 2000; Hanson et al., 2005). The thiol groups of the cysteines involved in the zinc-finger motif as part of the metal-binding, DNA-intercalating fingers (Kadonaga et al., 1987; Desjarlais and Berg, 1992) confer gene regulation by ROS (Cimino et al., 1997). These zinc finger structures are potentially very sensitive redox targets in DNA binding motifs of many DNA transcription factors and DNA repair enzymes (Berg, 1992; Rhodes and Klug, 1993; Wu et al., 1996). Through Fenton-type reactions transition metals, e.g. iron and copper (Pierre and Fontecave, 1999), may displace zinc ions in zinc-finger motifs located at the DNA target sites and convert RONS into more aggressive radicals, e.g. hydroxyl radical (Wink et al., 1994; Pecci et al., 1997) or peroxynitrite, which dose-dependently (Martin and Barrett, 2002) control gene expression of stress response systems and cell cycle events (O’Halloran, 1993;Cimino et al., 1997; Boldt, 1999), induce DNA breaks (Mello-Filho and Meneghini, 1984; Kazakov et al., 1988; Tachon, 1989; Bertoncini and Meneghini, 1995; Spear and Aust, 1995; Meneghini, 1997; Rodriguez et al., 1999), epimutagenesis (see chapter 10.3) and mutagenesis (see chapter 10.1) and mediate a variety of cell cycle events including differentiation (see chapter 6.1), apoptosis (see chapter 10.5) and carcinogenesis (Cerutti, 1985; Babbs, 1990; Lutz, 1990;Dreher and Junod, 1996; Tamir and Tannenbaum, 1996;Marnett, 2000; Klaunig and Kamendulis, 2004; Matés et al., 2008; Toyokuni, 2008; Klaunig et al., 2010; Liou and Storz, 2010; Sosa et al., 2013). Thus, the cellular energy flow, Ca2+ homeostasis and redox balance determine cellular-context and dose-dependent events like proliferation, differentiation, apoptosis, necrosis and oncosis.

This flow of information is not coded and specific as from gene to protein but code-free and stochastic. The randomness of the feedback process from environment to the genome process relies on the simple, codeless messenger agents ATP, Ca2+ and free radicals (Saran et al., 1998) both regulated by and regulating cellular and organismal homeostasis in a feedback triangle (Brookes et al., 2004; Camello-Almaraz et al., 2006; Yan Y et al., 2006; Feissner et al., 2009; Kowaltowski et al., 2009). Cellular oxidative stress-dependent responses, although undoubtedly programmed, are also highly variable (Heininger, 2012), at least in part based on the stochasticity of mitochondrial bioenergetic/oxidative events (Hüser et al., 1998; Genova et al., 2003; Passos et al., 2007; Wang W et al., 2008).In addition to cellular processes, these agents regulate organismal life history events like development and aging (Heininger, 2012) in response to environmental cues. The regulated stochastic nature of the effectors and the degeneracy of (epi)mutagenic tools that may act both as a source of robustness and evolvability (see chapter 13) allows multiple solutions for a given problem (Lenski and Travisano, 1994; Rosenzweig et al., 1994; Finkel and Kolter, 1999) and therefore has given rise to the huge diversity of evolution with an ever increasing complexity (Adami et al., 2000).

 

22. Conclusions


What is utterly baffling to me is why one cannot be a reductionist and a holist at the same time.

John Tyler Bonner, The Evolution of Complexity, 1988

Simplification may indeed be necessary for news articles, but it can distort the more complex and subtle realities of evolutionary patterns and mechanisms

Carroll (2005) Endless Forms Most Beautiful

Of all forms of mental activity, the most difficult … is the art of handling the same bundle of data as before, but placing them in a new system of relations with one another by giving them a different framework.

Thomas Kuhn, 1962

I attempted to decipher the evolutionary success of sexual reproduction by a Herschelian approach (Herschel, 1830). As Gildenhuys (2004) wrote: “For Herschel, scientific inquiry is a two-step process, beginning with the explanation of phenomena, through the exposure of their often hidden causes, and ending with the generalization of these causal processes to form laws of nature (1987, p. 144; first published 1830). Causal explanation involves decomposing a phenomenon into its component causes, a process Herschel calls ‘analysis’, comparing it to the decomposition of substances performed by chemists. During the analysis phase of inquiry, the researcher must discover the hidden operations that are responsible for outwardly sensible phenomena (ibid., p. 92).”

When embarking on the mission to elucidate the workings of a clock you can make two fundamental mistakes: the holist and reductionist mistake. The former was described by Feldman and Lewontin (1975) when they wrote: “There is a vast loss of information in going from a complex machine to a few descriptive parameters. Therefore, there is immense indeterminacy in trying to infer the structure of the machine from those few descriptive variables, themselves subject to error. It is rather like trying to infer the structure of a clock by listening to it tick and watching the hands.” Reductionists, on the other hand, may devote a whole scientific life to the study of a gearwheel in the clockwork, describe its size, the alloy it is made from, the number and form of its teeth, infer its function in the whole system from the effects of slowing-down and accelerating its movement and it is easily comprehensible that they finally may come to the conclusion that the gearwheel is the centerpiece and moves the clockwork. But the gearwheel may be only one amongst many and it is questionable whether the workings of the whole clockwork can be understood from its action. Yet, the reductionist approach is the only feasible method to understand the functioning of the clockwork as a whole. But there is no shortcut to study ALL gearwheels and their interactions. Conventional concepts of sexual reproduction considered a multitude of design features (most often from a population genetics approach) but they failed to apply an eminent principle: a systemic approach taking all nuts and bolts of the problem into account. A systemic approach is required to solve the Kuhnian puzzle. The insight that drives systems biology is that a full understanding of the role played by any one component in a biological process can be achieved only by considering it in its appropriate context in the whole system. In this sense, systems biology goes beyond a reductionist paradigm in which the properties of system components are considered in isolation (Bennett and Monk, 2008).

P. Kitcher (1981) advanced the idea that “..the natural sciences do not merely pile up unrelated items of knowledge of more or less practical significance, but that they increase our understanding of the world” (p. 508). And: “By using a few patterns of argument in the derivation of many beliefs we minimize the number of types of premises we must take as underived. That is, we reduce, in so far as possible, the number of types of facts we must accept as brute” (p. 529). Kitcher elaborated on earlier thoughts of Hempel: “What scientific explanation, especially theoretical explanation, aims at is not [an] intuitive and highly subjective kind of understanding, but an objective kind of insight that is achieved by a systematic unification, by exhibiting the phenomena as manifestations of common, underlying structures and processes that conform to specific, testable, basic principles” (Hempel 1966, p. 83; see also Hempel 1965, pp. 345, 444); and Feigl: “The aim of scientific explanation throughout the ages has been unification, i.e., the comprehending of a maximum of facts and regularities in terms of a minimum of theoretical concepts and assumptions” (Feigl 1970, p. 12).

In 1982, Graham Bell wrote: “Sex is the queen of problems in evolutionary biology”. Obviously, he coined this sentence for the issue of understanding the evolutionary rationale for the success of sexual reproduction. However, since genetic transmission by sexual reproduction is so pervasive in nature, the sentence has its connotation. A multitude of issues in evolutionary biology and population genetics have been virtually intractable without an in-depth understanding of the molecular mechanisms underlying sexual reproduction.

This and my previous work (Heininger, 2012) mark a fundamental change of paradigm. Both should not be seen independent of each other. In fact, they draw a huge coherent scenario within which reproduction and death unfold, both coselected and orchestrated by stochasticity and selection as the two general organizing principles of evolution.

These insights should have a major impact on the entire biomedical sciences, and particularly evolutionary theory.

23. Abbreviations


8 - oxoG: 8-oxoguanine

APE1: apurinic/apyrimidinic endonuclease 1

AQP: aquaporin

BER: base excision repair

CE: Cambrian explosion

CRH: corticotropin-releasing hormone

EGT: evolutionary game theory

eIF2: eukaryotic translation initiation factor 2

ESS: evolutionarily stable strategy

GA: Genetic Algorithm

GPG: General Purpose Genotype

HIF: hypoxia-inducible factor

HO: heme oxygenase

HR: homologous recombination

HPA: hypothalamic–pituitary–adrenal

HPG: hypothalamic–pituitary–gonadal

HSP: heat shock protein

MMR: mismatch repair

NER: nucleotide excision repair

NHEJ: non-homologous end joining

PAE: paternal age effect

PGCs: primordial germ cells

RONS: reactive oxygen nitrogen species

ROS: reactive oxygen species

SAM: S-adenosyl-L-methionine

SNPs: single nucleotide polymorphisms

SGs: stress granules

SMSC: sexual mutagenesis selection cascade

TAM: transcription-associated mutagenesis

TAR: transcription-associated recombination

TCA: tricarboxylic acid

TE: transposable element

Tet: ten eleven translocation

VEGF: vascular endothelial growth factor

24. References


Aasebo U, Gyltnes A, Bremnes RM, Aakvaag A, Slordal L (1993) Reversal of sexual impotence in male patients with chronic obstructive pulmonary disease and hypoxemia with long-term oxygen therapy. J Steroid Biochem Mol Biol 46: 799–803.

Abane R, Mezger V (2010) Roles of heat shock factors in gametogenesis and development. FEBS J 277: 4150-72.

Abasiyanik A, Dagdönderen L (2004) Beneficial effects of melatonin compared with allopurinol in experimental testicular torsion. J Pediatr Surg 39: 1238-41.

Abate C, Patel L, Rauscher FJI, Curran T (1990) Redox regulation of Fos and Jun DNA-binding activity in vitro. Science 249: 1157–61.

Abbott PJ (1985) Stimulation of recombination between homologous sequences on carcinogen-treated plasmid DNA and chromosomal DNA by induction of the SOS response in Escherichia coli K12. Mol Gen Genet 201: 129–32.

Abdalla H, Yoshizawa Y, Hochi S (2009) Active demethylation of paternal genome in mammalian zygotes.J Reprod Dev 55: 356-60.

Abdullah MFF, Borts RH (2001) Meiotic recombination frequencies are affected by nutritional states in Saccharomyces cerevisiae. Proc Natl Acad Sci USA 98: 14524–9.

Abele D, Heise K, Pörtner HO, Puntarulo S (2002) Temperature dependence of mitochondrial function and production of reactive oxygen species in the intertidal mud clam Mya arenaria. J Exp Biol 205: 1831–41.

Abele DK, Puntarulo S (2004) Formation of reactive species and induction of antioxidant defence systems in polar and temperate marine invertebrates and fish. Comp Biochem Physiol A 138: 405-15.

Abele D, Philipp E, Paula M. Gonzalez PM, Puntarulo S (2007) Marine invertebrate mitochondria and oxidative stress. Front Biosci 12: 933-46.

Able DJ (1996) The contagion indicator hypothesis for parasite-mediated sexual selection. Proc Natl Acad Sci USA 93: 2229-33.

Abrahamson S, Meyer HU, Himoe E, Daniel G (1966) Further evidence demonstrating germinal selection in early pre-meiotic germ cells of Drosophila males. Genetics 54: 687–96.

Abrams JM (2002) Competition and compensation: coupled to death in development and cancer. Cell 110: 403-6.

Abrams PA (2000) Character shifts of prey species that share predators. Am Nat 156 (suppl.): S46–S61.

Abrams PA (2006) Adaptive change in the resource-exploitation traits of a generalist consumer: the evolution and coexistence of generalists and specialists. Evolution 60: 427-39.

Abrams PA, Matsuda H (1997) Prey adaptation as a cause of predator–prey cycles. Evolution 51: 1742–50.

Acar M, Mettetal JT, van Oudenaarden A (2008) Stochastic switching as a survival strategy in fluctuating environments. Nat Genet 40: 471–5.

Achaz G, Rocha EPC, Netter P, Coissac E (2002) Origin and fate of repeats in bacteria. Nucleic Acids Res 30: 2987-94.

Achilli A, Matmati N, Casalone E, Morpurgo G, Lucaccioni A, et al. (2004) The exceptionally high rate of spontaneous mutations in the polymerase delta proofreading exonuclease-deficient Saccharomyces cerevisiae strain starved for adenine. BMC Genet 5: 34.

Ackerman JL, Bellwood DR, Brown JH (2004) The contribution of small individuals to density-body size relationships: examination of energetic equivalence in reef fishes. Oecologia 139: 568–71.

Ackerman S, Kermany AR, Hickey DA (2010) Finite populations, finite resources, and the evolutionary maintenance of genetic recombination. J Hered 101(Suppl 1): S135–S141.

Ackermann M, Chao L (2006) DNA sequences shaped by selection for stability. PLoS Genet 2: e22.

Ackermann M, Stecher B, Freed NE, Songhet P, Hardt WD, Doebeli M (2008) Self-destructive cooperation mediated by phenotypic noise. Nature 454: 987–90.

Ackley DH, Littman ML (1991) Interactions between learning and evolution. In: Langton CG, Farmer JD, Rasmussen S, Taylor CE, eds. Artificial Life II. Reading, MA: Addison-Wesley. pp 487-509.

Acquisti C, Kleffe J, Collins S (2007) Oxygen content of transmembrane proteins over macroevolutionary time scales. Nature 445: 47–52.

Adachi J, Cao Y, Hasegawa M (1993) Tempo and mode of mitochondrial DNA evolution in vertebrates at the amino acid sequence level: rapid evolution in warm-blooded vertebrates. J Mol Evol 36: 270–81.

Adachi-Yamada T, O'Connor MB (2004) Mechanisms for removal of developmentally abnormal cells: cell competition and morphogenetic apoptosis. J Biochem 136: 13-7.

Adami C, Ofria C, Collier TC (2000) Evolution of biological complexity. Proc Natl Acad Sci USA 97: 4463-8.

Adamo A, Woglar A, Silva N, Penkner A, Jantsch V, La Volpe A (2012) Transgene-mediated cosuppression and RNA interference enhance germ-line apoptosis in Caenorhabditis elegans. Proc Natl Acad Sci USA 109: 3440-5.

Adamo A, La Volpe A (2012) Try to disarm the intruder or kill him! Worm 1: 212–5.

Adams ES, Mesterton-Gibbons M (2003) Lanchester’s attrition models and fights among social animals. Behav Ecol 14: 719–23.

Adams KL, Wendel JF (2005) Novel patterns of gene expression in polyploid plants. Trends Genet 21: 539–43.

Adams M, Foster R, Hutchinson MN, Hutchinson RG, Donnellan SC (2003) The Australian scincid lizard Menetia greyii: a new instance of widespread vertebrate parthenogenesis. Evolution 57: 2619–27.

Adams RLP (1996) DNA methylation. In: Bittar EE, ed. Principles of Medical Biology, Vol. 5. New York, NY: JAI Press Inc. pp 33–66.

Adcock IM, Cosio B, Tsaprouni L, Barnes PJ, Ito K (2005) Redox regulation of histone deacetylases and glucocorticoid-mediated inhibition of the inflammatory response. Antioxid Redox Signal 7: 144-52.

Adelman R, Saul RL, Ames BN (1988) Oxidative damage to DNA: relation to species metabolic rate and life span. Proc Natl Acad Sci USA 85: 2706-8.

Adler GH, Levins R (1994) The island syndrome in rodent populations. Q Rev Biol 69: 473–90.

Adolfsson S, Michalakis Y, Paczesniak D, Bode SN, Butlin RK, et al. (2010) Evaluation of elevated ploidy and asexual reproduction as alternative explanations for geographic parthenogenesis in Eucypris virens ostracods. Evolution 64: 986-97.

Aerts L, Holemans K, Van Assche FA (1990) Maternal diabetes during pregnancy: consequences for the offspring. Diabetes Metab Rev 6: 147–67.

Aerts L, Van Assche FA (1992) Islet transplantation in diabetic pregnant rats normalizes glucose homeostasis in their offspring. J Dev Physiol 17: 283–7.

Aertsen A, Michiels CW (2005) Diversity or die: Generation of diversity in response to stress. Crit Rev Microbiol 31: 69-78.

Aeschlimann PB, Häberli MA, Reusch TB, Boehm T, Milinski M (2003) Female sticklebacks Gasterosteus aculeatus use self-reference to optimize MHC allele number during mate selection. Behav Ecol Sociobiol 54: 119-26.

Agani FH, Pichiule P, Chavez JC, LaManna JC (2000) The role of mitochondria in the regulation of hypoxia-inducible factor 1 expression during hypoxia. J Biol Chem 275:35863-7.

Agarwal A, Saleh RA, Bedaiwy MA (2003) Role of reactive oxygen species in the pathophysiology of human reproduction. Fertil Steril 79: 829-43.

Agarwal A, Allamaneni SS (2004) The effect of sperm DNA damage on assisted reproduction outcomes. A review. Minerva Ginecol 56: 235-45.

Agarwal A, Gupta S, Sharma RK (2005) Role of oxidative stress in female reproduction. Reprod Biol Endocrinol 3: 28.

Agarwal AK, Auchus RJ (2005) Minireview: cellular redox state regulates hydroxysteroid dehydrogenase activity and intracellular hormone potency. Endocrinology 146: 2531-8.

Agashe D, Falk JJ, Bolnick DI (2011) Effects of founding genetic variation on adaptation to a novel resource. Evolution 65: 2481–91.

Agbali M, Reichard M, Bryjová A, Bryja J, Smith C (2010) Mate choice for nonadditive genetic benefits correlate with MHC dissimilarity in the rose bitterling (Rhodeus ocellatus). Evolution 64: 1683-96.

Agoff SN, Hou J, Linzer DI, Wu B (1993) Regulation of the human hsp70 promoter by p53. Science 259: 84-7.

Agrafioti I, Swire J, Abbott J, Huntley D, Butcher S, Stumpf MPH (2005) Comparative analysis of the Saccharomyces cerevisiae and Caenorhabditis elegans protein interaction networks. BMC Evol Biol 5: 23.

Agrawal AA (2000a) Host-range evolution: adaptation and trade-offs in fitness of mites on alternative hosts. Ecology 81: 500–8.

Agrawal AA (2000b) Benefits and costs of induced plant defense for Lepidium virginicum (Brassicaceae). Ecology 81: 1804–13.

Agrawal AA (2001) Phenotypic plasticity in the interactions and evolution of species. Science 294: 321–6.

Agrawal AF (2001) Sexual selection and the maintenance of sexual reproduction. Nature 411: 692–5.

Agrawal AF (2002) Genetic loads under fitness-dependent mutation rates. J Evol Biol 15: 1004–10.

Agrawal AF (2006a) Similarity selection and the evolution of sex: revisiting the Red Queen. PLoS Biol 4: e265.

Agrawal AF (2006b) Evolution of sex: why do organisms shuffle their genotypes? Curr Biol 16: R696–R704.

Agrawal AF, Hadany L, Otto SP (2005) The evolution of plastic recombination. Genetics 171:803-12.

Agrawal AF, Wang AD(2008) Increased transmission of mutations by low-condition females: evidence for condition-dependent DNA repair. PLoS Biol 6: e30.

Agrawal AF, Whitlock MC (2010) Environmental duress and epistasis: how does stress affect the strength of selection on new mutations? Trends Ecol Evol 25:450-8.

Agre P, Kozono D (2003) Aquaporin water channels: molecular mechanisms for human diseases. FEBS Lett 555: 72–8.

Aguade M (1999) Positive selection drives the evolution of the Acp29AB accessory gland protein in Drosophila. Genetics 152: 543–51.

Aguade M, Miyashita N, Langley CH (1989) Reduced variation in the yellow-achaete-scute region in natural populations of Drosophila melanogaster. Genetics 122:607–15.

Aguilar-Mahecha A, Hales BF, Robaire B (2001) Expression of stress response genes in germ cells during spermatogenesis. Biol Reprod 65: 119–27.

Aguilera A (2002) The connection between transcription and genomic instability. EMBO J 21: 195–201.

Agulnik AI, Bishop CE, Lerner JL, Agulnik SI, Solovyev VV (1997) Analysis of mutation rates in the SMCY/SMCX genes shows that mammalian evolution is male driven. Mamm Genome 8: 134-8.

Aharoni A, Baran N, Manor H (1993) Characterization of a multi-subunit human protein which selectively binds single stranded d(GA)n and d(GT)n sequence repeats in DNA. Nucleic Acids Res 21: 5221-8.

Aharoni A, Gaidukov L, Khersonsky O, McQ Gould S, Roodveldt C, Tawfik DS (2005) The ‘evolvability’ of promiscuous protein functions. Nat Genet 37: 73–6.

Ahlbom E, Grandison L, Bonfoco E, Zhivotovsky B, Ceccatelli S (1999) Androgen treatment of neonatal rats decreases susceptibility of cerebellar granule neurons to oxidative stress in vitro. Eur J Neurosci 11: 1285-91.

Ahlbom E, Prins GS, Ceccatelli S (2001) Testosterone protects cerebellar granule cells from oxidative stress-induced cell death through a receptor mediated mechanism. Brain Res 892: 255–62.

Ahloowalla BS, Maluszynski M, Nichterlein K (2004) Global impact of mutation-derived varieties. Euphytica 135: 187-204.

Ahmadi A, Ng SC (1999) Fertilizing ability of DNA-damaged spermatozoa. J Exp Zool 284: 696-704.

Ahmed EA, van der Vaart A, Barten A, Kal HB, Chen J, et al. (2007) Differences in DNA double strand breaks repair in male germ cell types: lessons learned from a differential expression of Mdc1 and 53BP1. DNA Repair (Amst) 6: 1243-54.

Ahmed EA, Barten-van Rijbroek AD, Kal HB, Sadri-Ardekani H, Mizrak SC, et al. (2009) Proliferative activity in vitro and DNA repair indicate that adult mouse and human Sertoli cells are not terminally differentiated, quiescent cells. Biol Reprod 80: 1084-91.

Ahmed EA, Philippens ME, Kal HB, de Rooij DG, de Boer P (2010a) Genetic probing of homologous recombination and non-homologous end joining during meiotic prophase in irradiated mouse spermatocytes. Mutat Res 688: 12-8.

Ahmed EA, de Boer P, Philippens ME, Kal HB, de Rooij DG (2010b) Parp1-XRCC1 and the repair of DNA double strand breaks in mouse round spermatids. Mutat Res 683: 84-90.

Ahn SG, Thiele DJ (2003) Redox regulation of mammalianheat shock factor 1 is essential for Hsp gene activation and protection from stress. Genes Dev 17: 516–28.

Aisenberg A, Costa FG (2005) Females mated without sperm transfer maintain high sexual receptivity in the wolf spider Schizocosa malitiosa. Ethology 111: 545–58.

Aitken RJ (1994) A free-radical theory of male-infertility. Reprod Fertil Dev 6: 19–24.

Aitken RJ (1995) Free radicals, lipid peroxidation and sperm function. Reprod Fertil Dev 7: 659-68.

Aitken RJ (1999) The Amoroso Lecture. The human spermatozoon--a cell in crisis? J Reprod Fertil 115: 1-7.

Aitken RJ, Clarkson JS, Fishel S (1989a) Generation of reactive oxygen species, lipid peroxidation and human sperm function. Biol Reprod 40: 183–97.

Aitken RJ, Clarkson JS, Hargreave TB, Irvine DS, Wu FC (1989b) Analysis of the relationship between defective sperm function and the generation of reactive oxygen species in cases of oligozoospermia. J Androl 10: 214–20.

Aitken RJ, Buckingham DW, West KM (1992) Reactive oxygen species and human spermatozoa: analysis of the cellular mechanisms involved in luminol- and lucigenin-dependent chemiluminescence. J Cell Physiol 151: 466–77.

Aitken RJ, Paterson M, Fisher H, Buckingham DW, van Duin M (1995) Redox regulation of tyrosine phosphorylation in human spermatozoa and its role in the control of human sperm function. J Cell Sci 108: 2017–25.

Aitken RJ, Fisher HM, Fulton N, Gomez E, KnoxW, et al. (1997) Reactive oxygen species generated by human spermatozoa is induced by exogenous NADPH and inhibited by the flavoprotein inhibitors diphenylene iodonium and quinacrine. Mol Reprod Dev 47: 468–82.

Aitken RJ, Gordon E, Harkiss D, Twigg JP, Milne P, Jennings Z, Irvine DS (1998a) Relative impact of oxidative stress on the functional competence and genomic integrity of human spermatozoa. Biol Reprod 59: 1037–46.

Aitken RJ, Harkiss D, Knox W, Paterson M, Irvine DS (1998b) A novel signal transduction cascade in capacitating human spermatozoa characterised by a redox-regulated, cAMP-mediated induction of tyrosine phosphorylation. J Cell Sci 111: 645-56.

Aitken RJ, Sawyer D (2003) The human spermatozoon--not waving but drowning. Adv Exp Med Biol 518: 85-98.

Aitken RJ, Baker MA, Sawyer D (2003) Oxidative stress in the male germ line and its role in the aetiology of male infertility and genetic disease. Reprod Biomed Online 7: 65-70.

Aitken RJ, Ryan AL, Baker MA, McLaughlin EA (2004) Redox activity associated with the maturation and capacitation of mammalian spermatozoa. Free Radic Biol Med 36: 994–1010.

Aitken RJ, Wingate JK, De Iuliis GN, Koppers AJ, McLaughlin EA (2006) Cis-unsaturated fatty acids stimulate reactive oxygen species generation and lipid peroxidation in human spermatozoa. J Clin Endocrinol Metab 91: 4154-63.

Aitken RJ, Roman SD (2008) Antioxidant systems and oxidative stress in the testes. Oxid Med Cell Longev 1: 15–24.

Aitken RJ, Findlay JK, Hutt KJ, Kerr JB (2011) Apoptosis in the germ line. Reproduction 141:139-50.

Akahoshi T, Oppenheim JJ, Matsushima K (1988) Interleukin 1 stimulates its own receptor expression on human fibroblasts through the endogenous production of prostaglandin(s). Clin Invest 82: 1219-26.

Akamatsu Y, Ohno T, Hirota K, Kagoshima H, Yodoi J, Shigesada K (1997) Redox regulation of the DNA binding activity in transcription factor PEBP2. The roles of two conserved cysteine residues. J Biol Chem 272: 14497–500.

Akashi H (1994) Synonymous codon usage in Drosophila melanogaster: natural selection and translational accuracy. Genetics 136: 927–35.

Akashi H (1995) Inferring weak selection from patterns of polymorphism and divergence at ‘silent’ sites in Drosophila DNA. Genetics 139: 1067–76.

Akashi H (1997) Codon bias evolution in Drosophila. Population genetics of mutation-selection drift. Gene 205: 269–78.

Akashi H, Kliman RM, Eyre-Walker A (1998) Mutation pressure, natural selection and the evolution of base composition in Drosophila. Genetica 102/103: 49–60.

Akashi H, Gojobori T (2002) Metabolic efficiency and amino acid composition in the proteomes of Escherichia coli and Bacillus subtilis. Proc Natl Acad Sci USA 99: 3695–700.

Åkerfelt M, Trouillet D, Mezger V, Sistonen L (2007) Heat shock factors at a crossroad between stress and development. Ann NY Acad Sci 1113: 15-27.

Åkerfelt M, Henriksson E, Laiho A, Vihervaara A, Rautoma K, et al. (2008) Promoter ChIP-chip analysis in mouse testis reveals Y chromosome occupancy by HSF2. Proc Natl Acad Sci USA 105: 11224–9.

Åkerfelt M, Morimoto RI, Sistonen L (2010) Heat shock factors: integrators of cell stress, development and lifespan. Nat Rev Mol Cell Biol 11: 545–55.

Akerlund T, Nordström K, Bernander R (1995) Analysis of cell size and DNA content in exponentially growing and stationary-phase batch cultures of Escherichia coli. J Bacteriol 177: 6791-7.

Alatalo RV, Kotiaho J, Mappes J, Parri S (1998) Mate choice for offspring performance: major benefits or minor costs? Proc R Soc Lond B 265: 2297-301.

Albà MM, Castresana J (2005) Inverse relationship between evolutionary rate and age of mammalian genes. Mol Biol Evol 22: 598-606.

Albon SD, Mitchell B, Staines B (1983) Fertility and body weight in female red deer: a density-dependent relationship. J Anim Ecol 52: 969-80.

Albrecht AN, Kornak U, Boddrich A, Suring K, Robinson PN, et al. (2004) A molecular pathogenesis for transcription factor associated polyalanine tract expansions. Hum Mol Genet 13: 2351–9.

Albrecht A, Mundlos S (2005) The other trinucleotide repeat: polyalanine expansion disorders. Curr Opin Genet Dev 15: 285–93.

Albrechtsen N, Dornreiter I, Grosse F, Kim E, Wiesmüller L, Deppert W (1999) Maintenance of genomic integrity by p53: complementary roles for activated and non-activated p53. Oncogene 18: 7706-17.

Albu M, Kermany AR, Hickey DA (2012) Recombination reshuffles the genotypic deck, thus accelerating the rate of evolution. In: Singh RS, Xu J, Kulathinal RJ, eds. Rapidly Evolving Genes and Genetic Systems. Oxford, UK: Oxford University Press. pp 23-30.

Alcazar RM, Lin R, Fire AZ (2008) Transmission dynamics of heritable silencing induced by double-stranded RNA in Caenorhabditis elegans. Genetics 180: 1275–88.

Alcivar AA, Hake LE, Hecht NB (1992) DNA polymerase-beta and poly(ADP)ribose polymerase mRNAs are differentially expressed during the development of male germinal cells. Biol Reprod 46: 201-7.

Aldana M, Balleza E, Kauffman S, Resendiz O (2007) Robustness and evolvability in genetic regulatory networks. J Theor Biol 245: 433–48.

Alemán LM, Doench J, Sharp PA (2007) Comparison of siRNA-induced off-target RNA and protein effects. RNA 13: 385-95.

Alesci S, Manoli I, Michopoulos VJ, Brouwers FM, Le H, et al. (2006) Development of a human mitochondria-focused cDNA microarray (hMitChip) and validation in skeletal muscle cells: implications for pharmaco- and mitogenomics. Pharmacogenomics J 6: 333–42.

Alexander RD (1979) Darwinism and human affairs. Seattle, WA: University of Washington Press.

Alexander RD (1987) The biology of moral systems. New York, NY: Aldine de Gruyter.

Alexander RD, Marshall DC, Cooley JR (1997) Evolutionary perspectives on insect mating. In: Choe JC, Crespi BJ, eds. The Evolution of Mating Systems in Insects and Arachnids. Cambridge, UK: Cambridge University Press. pp 4–31.

Alexander RM (1982) Optima for animals. London, UK: Edward Arnold.

Alexandreanu IC, Lawson DM (2003) Heme oxygenase in the rat ovary: immunohistochemical localization and possible role in steroidogenesis. Exp Biol Med 228: 59-63.

Alfaradhi MZ, Ozanne SE (2011) Developmental programming in response to maternal overnutrition. Front Gene 2: 27.

Ali A, Li H, Schneider WL, Sherman DJ, Gray S, et al. (2006) Analysis of genetic bottlenecks during horizontal transmission of cucumber mosaic virus. J Virol 80: 8345–50.

Alié A, Leclère L, Jager M, Dayraud C, Chang P, et al. (2011) Somatic stem cells express Piwi and Vasa genes in an adult ctenophore: ancient association of “germline genes” with stemness. Dev Biol 350: 183–97.

Allan DJ, Harmon BV, Roberts SA (1992) Spermatogonial apoptosis has three morphologically recognizable phases and shows no circadian rhythm during normal spermatogenesis in the rat. Cell Prolif 25: 241-50.

Allegrucci C, Thurston A, Lucas E, Young L (2005) Epigenetics and the germline. Reproduction 129: 137–49.

Allen AP, Brown JH, Gillooly JF (2002) Global biodiversity, biochemical kinetics, and the energetic-equivalence rule. Science 297: 1545–8.

Allen A, Gillooly J, Savage V, Brown J (2006) Kinetic effects of temperature on rates of genetic divergence and speciation. Proc Natl Acad Sci USA 103: 9130–5.

Allen AP, Gillooly JF, Brown JH (2007) Recasting the species-energy hypothesis: the different roles of kinetic and potential energy in regulating biodiversity. In: Storch D, Marquet PA, Brown JH, eds. Scaling Biodiversity. Cambridge, UK: Cambridge University Press. pp 283–299.

Allen B, Scholes Rosenbloom DI (2012) Mutation rate evolution in replicator dynamics. Bull Math Biol 74: 2650-75.

Allen C, Garmestani A, Havlicek T, Marquet P, Peterson G, et al. (2006) Patterns in body mass distributions: sifting among alternative hypotheses. Ecol Lett 9: 630–43.

Allen D, Herbert DC, McMahan CA, Rotrekl V, Sobol RW, et al. (2008) Mutagenesis is elevated in male germ cells obtained from DNA polymerase-beta heterozygous mice. Biol Reprod 79: 824–31.

Allen DE, Lynch M (2008)Both costs and benefits of sex correlate with relative frequency of asexual reproduction in cyclically parthenogenic Daphnia pulicaria populations. Genetics 179: 1497-502.

Allen JA, Diemer T, Janus P, Hales KH, Hales DB (2004) Bacterial endotoxin lipopolysaccharide and reactive oxygen species inhibit Leydig cell steroidogenesis via perturbation of mitochondria. Endocrine 25: 265–75.

Allen JA, Shankara T, Janus P, Buck S, Diemer T, et al.(2006) Energized, polarized, and actively respiring mitochondria are required for acute Leydig cell steroidogenesis. Endocrinology 147: 3924-35.

Allen JF (1993) Control of gene expression by redox potential and the requirement for chloroplast and mitochondrial genomes. J Theor Biol 165: 609-31.

Allen JF (2003) The function of genomes in bioenergetic organelles. Philos Trans R Soc Lond B Biol Sci 358: 19–37.

Allen JM, Light JE, Perotti MA, Braig HR, Reed DL (2009) Mutational meltdown in primary endosymbionts: selection limits Muller’s ratchet. PLoS ONE 4: e4969.

Allen JS, Cheer SM (1996) The non-thrifty genotype. Curr Anthropol 37: 831–42.

Allen JW, Dix DJ, Collins BW, Merrick BA, He C, et al. (1996) HSP70-2 is part of the synaptonemal complex in mouse and hamster spermatocytes. Chromosoma 104: 414-21.

Allen RG (1991) Oxygen-reactive species and antioxidant responses during development: the metabolic paradox of cellular differentiation. Proc Soc Exp Biol Med 196: 117-29.

Allen RG (1998) Oxidative stress and superoxide dismutase in development, aging and gene regulation. Age 21: 47-76.

Allen RM, Buckley YM, Marshall DJ (2008) Offspring size plasticity in response to intraspecific competition: an adaptive maternal effect across life-history stages. Am Nat 171: 225–237.

Allendorf FW, Leary RF (1986) Heterozygosity and fitness in natural populations of animals. In: Soulé ME, ed. Conservation Biology: the science of scarcity and diversity. Sunderland, MA: Sinauer Associates. pp 57–76.

Al-Mehdi AB, Pastukh VM, Swiger BM, Reed DJ, Patel MR, et al. (2012)Perinuclear mitochondrial clustering creates an oxidant-rich nuclear domain required for hypoxia-induced transcription.Sci Signal 5: ra47.

Almeida FFL, Kristoffersen C, Taranger GL, Schulz RW (2008) Spermatogenesis in Atlantic cod (Gadus morhua): a novel model of cystic germ cell development. Biol Reprod 78:27–34.

al-Mukhtar KA, Webb AC (1971) An ultrastructural study of primordial germ cells, oogonia and early oocytes in Xenopus laevis. J Embryol Exp Morphol 26: 195–217.

Alonso-Alvarez C, Bertrand S, Devevey G, Prost J, Faivre B, Sorci G (2004a) Increased susceptibility to oxidative stress as a proximate cost of reproduction. Ecol Lett 7: 363–8.

Alonso-Alvarez C, Bertrand S, Devevey G, Gaillard M, Prost J, et al. (2004b) An experimental test of the dose-dependent effect of carotenoids and immune activation on sexual signals and antioxidant activity. Am Nat 164: 651–9.

Alonso-Alvarez C, Bertrand S, Devevey G, Prost J, Faivre B, et al. (2006)An experimental manipulation of life-history trajectories and resistance to oxidative stress.Evolution 60: 1913–24.

Alonso-Alvarez C, Bertrand S, Faivre B, Chastel O, Sorci G (2007)Testosterone and oxidative stress: the oxidation handicap hypothesis.Proc R Soc Lond B Biol Sci 274: 819-25.

Alpert P (2006) Constraints of tolerance: why are desiccation-tolerant organisms so small or rare? J Exp Biol 209: 1575–84.

Altenberg L, Feldman MW (1987) Selection, generalized transmission and the evolution of modifier genes. I. The reduction principle. Genetics 117: 559–72.

Altenhöfer S, Kleikers PWM, Radermacher KA, Scheurer P, Hermans JJR,et al. (2012) The NOX toolbox: validating the role of NADPH oxidases in physiology and disease. Cell Mol Life Sci 69: 2327–43.

Altiero T, Rebecchi L, Bertolani R (2006) Phenotypic variations in the life history of two clones of Macrobiotus richtersi (Eutardigrada, Macrobiotidae). Hydrobiologia 558: 33-40.

Altizer S, Harvell D, Friedle E (2003) Rapid evolutionary dynamics and disease threats to biodiversity. Trends Ecol Evol 18: 589–96.

Alvarez B, Demicheli V, Duran R, Trujillo M, Cervenansky C, et al. (2004) Inactivation of human Cu,Zn superoxide dismutase by peroxynitrite and formation of histidinyl radical. Free Radic Biol Med 37: 813–22.

Alvarez JG, Touchstone JC, Blasco L, Storey BT (1987) Spontaneous lipid peroxidation and production of hydrogen peroxide and superoxide in human spermatozoa. Superoxide dismutase as major enzyme protectant against oxygen toxicity. J Androl 8: 338–48.

Alves K, Canzian M, Delwart EL (2002) HIV type 1 envelope quasispecies in the thymus and lymph nodes of AIDS patients. Aids Res Hum Retrov 18: 161–5.

Alves MJ, Coelho MM, Collares-Pereira MJ (2001) Evolution in action through hybridisation and polyploidy in an Iberian freshwater fish: a genetic review. Genetica 111: 375-85.

Amagai A (2011) Ethylene as a potent inducer of sexual development. Dev Growth Differ 53: 617–23.

Amann RP (1970) Sperm production rates. In: Johnson AD, Gomes WR, Vandemark NL, eds. The Testis. Academic Press. pp 433–82.

Amann RP (2010) Evaluating testis function non-invasively: how epidemiologist–andrologist teams might better approach this task. Hum Reprod 25: 22-8.

Amann RP, Johnson L, Thompson DL Jr., Pickett BW (1976) Daily spermatozoal production, epididymal spermatozoal reserves and transit time of spermatozoa through the epididymis of the rhesus monkey. Biol Reprod 15: 586-92.

Amann RP, Howards SS (1980)Daily spermatozoal production and epididymal spermatozoal reserves of the human male.J Urol 124:211-5.

Amaral S, Mota P, Rodrigues AS, Martins L, Oliveira PJ, Ramalho-Santos J (2008)Testicular aging involves mitochondrial dysfunction as well as an increase in UCP2 levels and proton leak.FEBS Lett 582: 4191-6.

Amaravadi RK, Yu D, Lum JJ, Bui T, Christophorou MA, et al. (2007) Autophagy inhibition enhances therapy-induced apoptosis in a Myc-induced model of lymphoma. J Clin Invest 117: 326–36.

Amat R, Solanes G, Giralt M, Villarroya F (2007) SIRT1 is involved in glucocorticoid-mediated control of uncoupling protein-3 gene transcription. J Biol Chem 282: 34066-76.

Ambady S, Wu Z, Dominko T (2012)Identification of novel microRNAs in Xenopus laevis metaphase II arrested eggs.Genesis 50: 286-99.

Ambros V (2004) The functions of animal microRNAs. Nature 431: 350–5.

Ambrosio L, Schedl P (1984) Gene expression during Drosophila melanogaster oogenesis: Analysis by in situ hybridization to tissue sections. Dev Biol 105: 80-92.

Amelio I, Grespi F, Annicchiarico-Petruzzelli M, Melino G (2012)p63 the guardian of human reproduction.Cell Cycle 11: 4545-51.

Ames BN, Shigenaga MK, Hagen TM (1993)Oxidants, antioxidants, and the degenerative diseases of aging.Proc Natl Acad Sci USA 90: 7915-22.

Aminetzach YT, Macpherson JM, Petrov DA (2005) Pesticide resistance via transposition-mediated adaptive gene truncation in Drosophila. Science 309: 764–7.

Amiri A, Keiper BD, Kawasaki I, Fan Y, Kohara Y, et al. (2001) An isoform of eIF4E is a component of germ granules and is required for spermatogenesis in C.elegans. Development 128: 3899–912.

Amitin EG, Pitnick S (2007) Influence of developmental environment on male- and female-mediated sperm precedence in Drosophila melanogaster. J Evol Biol 20: 381–91.

Ammerer G (1994) Sex, stress and integrity: the importance of MAP kinases in yeast. Curr Opin Genet Dev 4: 90-5.

Ammons MC, Siemsen DW, Nelson-Overton LK, Quinn MT, Gauss KA (2007)Binding of pleomorphic adenoma gene-like 2 to the tumor necrosis factor (TNF)-alpha-responsive region of the NCF2 promoter regulates p67phox expression and NADPH oxidase activity.J Biol Chem 282: 17941-52.

Amon P, Haas E, Sumper M (1998) The sex-inducing pheromone and wounding trigger the same set of genes in the multicellular green alga Volvox. Plant Cell 10: 781-9.

Amsellem L, Noyer JL, Hossaert-McKey M (2001) Evidence for a switch in the reproductive biology of Rubus alceifolius (Rosaceae) towards apomixis, between its native range and its area of introduction. Am J Bot 88: 2243–51.

Amstad P, Moret R, Cerutti P (1994) Glutathione peroxidase compensates for the hypersensitivity of Cu,Zn-superoxide dismutase overproducers to oxidant stress. J Biol Chem 269: 1606-9.

Amundson R (1994) Two concepts of constraint: Adaptationism and the challenge from developmental biology. Philos Sci61: 556–78.

Amundson R (2001) Adaptation, development, and the quest for common ground. In: Orzack SH, SoberE, eds. Adaptationist and optimality. New York, NY: Cambridge University Press. pp 303–334.

An WG, Kanekal M, Simon MC, Maltepe E, Blagosklonny MV, Neckers LM (1998)Stabilization of wild-type p53 by hypoxia-inducible factor 1alpha.Nature 392: 405-8.

Anagnostopoulos T, Green PM, Rowley G, Lewis CM, Giannelli F (1999) DNA variation in a 5-Mb region of the X chromosome and estimates of sex-specific/type-specific mutation rates. Am J Hum Genet 64: 508-17.

Anand-Ivell R, Ivell R (2011) The special systems biology of the sperm. Biochem J 436:e3-5.

Ancel LW (1999) A quantitative model of the Simpson–Baldwin effect. J Theor Biol 196: 197–209.

Ancel LW (2000)Undermining the Baldwin expediting effect: does phenotypic plasticity accelerate evolution?Theor Popul Biol 58: 307-19.

Ancel LW, Fontana W (2000) Plasticity, evolvability, and modularity in RNA. J Exp Zool 288: 242–83.

Ancliff M, Park JM (2009) Maximum, minimum, and optimal mutation rates in dynamic environments. Phys Rev E Stat Nonlin Soft Matter Phys 80: 061910.

Anderbrant O, Schlyter F, Birgersson G (1985) Intraspecific competition affecting parents and offspring in the bark beetle Ips typographus. Oikos 45: 89-98.

Andersen DH (1926) Lymphatics and blood-vessels of the ovary of the sow. Contrib Embryol Carnegie Instn 88: 109-23.

Andersen R, Linnell JDC (2000) Irruptive potential in roe deer: Density-dependent effects on body mass and fertility. J Wildlife Manage 64: 698–706.

Anderson DJ (1989) The role of hatching asynchrony in siblicidal brood reduction of two booby species. Behav Ecol Sociobiol 25:363-8.

Anderson GM, Barrell GK (1998) Out-of-season breeding in thyroidectomized red deer hinds. Proc New Zeal Soc An 58: 20–4.

Anderson JB, Sirjusingh C (2004) Haploidy, diploidy and evolution of antifungal drug resistance in Saccharomyces cerevisiae. Genetics 168: 1915–23.

Anderson JF, Vossbrinck CR, Andreadis TG, Iton A, Beckwith WH III, Mayo DR(2001) A phylogenetic approach to following West Nile virus in Connecticut. Proc Natl Acad Sci USA 98: 12885–9.

Anderson MJ, Dixson AF (2002) Sperm competition: Motility and the midpiece in primates. Nature 416: 496.

Anderson MJ, Dixson AS, Dixson AF (2006) Mammalian sperm and oviducts are sexually selected: evidence for co-evolution. J Zool 270: 682–6.

Anderson P, Kedersha N (2002a)Visibly stressed: the role of eIF2, TIA-1, and stress granules in protein translation.Cell Stress Chaperones 7: 213-21.

Anderson P, Kedersha N (2002b) Stressful initiations. J Cell Sci 115: 3227–34.

Anderson P, Kedersha N (2006)RNA granules.J Cell Biol 172: 803-8.

Anderson P, Kedersha N (2008)Stress granules: the Tao of RNA triage.Trends Biochem Sci 33: 141-50.

Anderson P, Kedersha N (2009)Stress granules.Curr Biol 19: R397-8.

Andersson C (2011) Paleolithic punctuations and equilibria: did retention rather than invention limit technological evolution? PaleoAnthropology 2011: 243−59.

Andersson DI (2006) The biological cost of mutational antibiotic resistance: any practical conclusions? Curr Opin Microbiol 9: 461–5.

Andersson DI, Hughes D (1996) Muller’s ratchet decreases fitness of a DNA-based microbe. Proc Natl Acad Sci USA 93: 906–7.

Andersson DI, Slechta ES, Roth JR (1998) Evidence that gene amplification underlies adaptive mutability of the bacterial lac operon. Science 282: 1133–5.

Andersson DI, Levin BR (1999) The biological cost of antibiotic resistance. Curr Opin Microbiol 2: 489–93.

Andersson DI, Hughes D (2010)Antibiotic resistance and its cost: is it possible to reverse resistance?Nat Rev Microbiol 8: 260-71.

Andersson DI, Hughes D, Roth JR, eds. (2011) The origin of mutants under selection: Interactions of mutation, growth, and selection. Washington, DC: ASM Press.

Andersson M (1994) Sexual Selection. Princeton, NJ: Princeton University Press.

Andersson S (1993) The potential for selective seed maturation in Achillea ptarmica (Asteraceae). Oikos 66: 36-42.

Ando K, Hirao S, Kabe Y, Ogura Y, Sato I, et al. (2008)A new APE1/Ref-1-dependent pathway leading to reduction of NF-kappaB and AP-1, and activation of their DNA-binding activity.Nucleic Acids Res 36: 4327-36.

Andolfatto P (2005) Adaptive evolution of non-coding DNA in Drosophila. Nature 437: 1149–52.

André JB, Godelle B (2006) The evolution of mutation rate in finite asexual populations. Genetics 172: 611–26.

Andreev D, Kreitman M, Phillips TW, Beeman RW, ffrench-Constant RH (1999) Multiple origins of cyclodiene insecticide resistance in Tribolium castaneum (Coleoptera: Tenebrionidae). J Mol Evol 48: 615–24.

Andreyev AY, Kushnareva YE, Starkov AA (2005) Mitochondrial metabolism of reactive oxygen species. Biochemistry (Moscow) 70: 200–14.

Andric SA, Janjic MM, Stojkov NJ, Kostic TS. Protein kinase G-mediated stimulation of basal Leydig cell steroidogenesis. Am J Physiol Endocrinol Metab 293: E1399–E1408.

Andux S, Ellis RE (2008) Apoptosis maintains oocyte quality in aging Caenorhabditis elegans females. PLoS Genet 4: e1000295.

Angers B, Schlosser IJ (2007) The origin of Phoxinus eos-neogaeus unisexual hybrids. Mol Ecol 16: 4562–71.

Angers B, Castonguay E, Massicotte R (2010) Environmentally induced phenotypes and DNA methylation: how to deal with unpredictable conditions until the next generation and after. Mol Ecol 19: 1283–95.

Angert AL (2006) Demography of central and marginal populations of monkeyflowers (Mimulus cardinalis and M. lewisii). Ecology 87: 2014–25.

Angert AL, Schemske DW (2005) The evolution of species’ distributions: reciprocal transplants across the elevation ranges of Mimulus cardinalis and M. lewisii. Evolution 59: 1671–84.

Angilletta MJ Jr, Sears MW (1999) The metabolic cost of reproduction in an oviparous lizard. Funct Ecol 14: 39–45.

Angkeow P, Deshpande SS, Qi B, Liu YX, Park YC, Jeon BH, Ozaki M, Irani K (2002) Redox factor-1: an extra-nuclear role in the regulation of endothelial oxidative stress and apoptosis. Cell Death Differ 9: 717–25.

Angus RA, Schultz RJ (1979) Clonal diversity in the unisexual fish Poeciliopsis monacha-lucida: a tissue graft analysis. Evolution 33: 27-40.

Anisimov VN (2003) Effects of exogenous melatonin. A review. Toxicol Pathol 31: 589-603.

Ankel-Simons F, Cummins JM (1996) Misconceptions about mitochondria and mammalian fertilization: implications for theories on human evolution. Proc Natl Acad Sci USA 93: 13859–63.

Annesley SJ, Fisher PR (2009) Dictyostelium discoideum—a model for many reasons. Mol Cell Biochem329: 73–91.

Antequera F, Bird A (1993) CpG islands. Exs 64: 169-185.

Antolin MF, Strobeck C (1985) The population genetics of somatic mutations in plants. Am Nat 126: 52–62.

Antoni FA (1986) Hypothalamic control of adrenocorticotropin secretion: advances since the discovery of 41-residue corticotropin-releasing factor. Endocr Rev 7: 351-78.

Antonovics J (1976) The nature of limits to natural selection. Ann Missouri Bot Gard 63: 224–47.

Antunes F, Cadenas E (2000) Estimation of H2O2 gradients across biomembranes. FEBS Lett 475: 121–6.

Anway MD, Cupp AS, Uzumcu M, Skinner MK (2005) Epigenetic transgenerational actions of endocrine disruptors and male fertility. Science 308: 1466–9.

Anway MD, Skinner MK (2006) Epigenetic transgenerational actions of endocrine disruptors. Endocrinology 147: S43–S49.

Anway MD, Skinner MK (2008) Epigenetic programming of the germ line: effects of endocrine disruptors on the development of transgenerational disease. Reprod Biomed Online 16: 23–5.

Anxolabéhère D, Kidwell MG, Periquet G (1988) Molecular characteristics of diverse populations are consistent with the hypothesis of a recent invasion of Drosophila melanogaster by mobile P elements. Mol Biol Evol 5: 252–69.

Aoubala M, Murray-Zmijewski F, Khoury MP, Fernandes K, Perrier S, et al. (2011) p53 directly transactivates Δ133p53α, regulating cell fate outcome in response to DNA damage. Cell Death Differ 18: 248-58.

Apel K, Hirt H (2004) Reactive oxygen species: metabolism, oxidative stress, and signal transduction. Annu Rev Plant Biol 55: 373–99.

Aprelikova O, Pandolfi S, Tackett S, Ferreira M, Salnikow K, et al. (2009) Melanoma antigen-11 inhibits the hypoxia-inducible factor prolyl hydroxylase 2 and activates hypoxic response. Cancer Res 69: 616–24.

Aquilina G, Bignami M (2001)Mismatch repair in correction of replication errors and processing of DNA damage. J Cell Physiol 187: 145-54.

Aragonés J, Fraisl P, Baes M, Carmeliet P (2009) Oxygen sensors at the crossroad of metabolism. Cell Metab9:11–22.

Arakaki N, Miyoshi T, Noda H (2001) Wolbachia-mediated parthenogenesis in the predatory thrips Franklinothrips vespiformis (Thysanoptera: Insecta). Proc R Soc Lond B Biol Sci 268: 1011–6.

Araki H, Cooper B, Blouin MS (2007) Genetic effects of captive breeding cause a rapid, cumulative fitness decline in the wild. Science 318: 100–3.

Araki H, Cooper B, Blouin MS (2009) Carry-over effect of captive breeding reduces reproductive fitness of wild-born descendants in the wild. Biol Lett 5: 621–4.

Araripe LO, Klaczko LB, Moreteau B, David JR (2004) Male sterility thresholds in a tropical cosmopolitan drosophilid, Zaprionus indianus. J Thermal Biol 29: 73-80.

Aravin AA, Naumova NM, Tulin AV, Vagin VV, Rozovsky YM, Gvozdev VA (2001) Double-stranded RNA-mediated silencing of genomic tandem repeats and transposable elements in the D. melanogaster germline. Curr Biol 11: 1017–27.

Aravin AA, Lagos-Quintana M, Yalcin A, Zavolan M, Marks D, et al. (2003) The small RNA profile during Drosophila melanogaster development. Dev Cell 5: 337–50.

Aravin AA, Klenov MS, Vagin VV, Bantignies F, Cavalli G, Gvozdev VA (2004) Dissection of a natural RNA silencing process in the Drosophila melanogaster germ line. Mol Cell Biol 24: 6742–6750.

Aravin A, Gaidatzis D, Pfeffer S, Lagos-Quintana M, Landgraf P, et al. (2006) A novel class of small RNAs bind to MILI protein in mouse testes. Nature 442: 203–7.

Aravin AA, Sachidanandam R, Girard A, Fejes-Toth K, Hannon GJ (2007a) Developmentally regulated piRNA clusters implicate MILI in transposon control. Science 316: 744–7.

Aravin AA, Hannon GJ, Brennecke J (2007b) The Piwi-piRNA pathway provides an adaptive defense in the transposon arms race. Science 318: 761–4.

Aravin AA, Sachidanandam R, Bourc’his D, Schaefer C, Pezic D, et al. (2008) A piRNA pathway primed by individual transposons is linked to de novo DNA methylation in mice. Mol Cell 31: 785–99.

Aravin AA, Bourc'his D (2008) Small RNA guides for de novo DNA methylation in mammalian germ cells.Genes Dev 22: 970-5.

Aravin AA, van der Heijden GW, Castañeda J, Vagin VV, Hannon GJ, Bortvin A (2009) Cytoplasmic compartmentalization of the fetal piRNA pathway in mice. PLoS Genet 5: e1000764.

Arber W (1999) Involvement of gene products in bacterial evolution. Ann NY Acad Sci 870:36-44.

Arber W, Naas T, Blot M (1994) Generation of genetic diversity by DNA rearrangements in resting bacteria. FEMS Microbiol Ecol 15: 5-13.

Arbuthnott D, Rundle HD (2012) Sexual selection is ineffectual or inhibits the purging of deleterious mutations in Drosophila melanogaster.Evolution 66: 2127-37.

Ardanaz N, Pagano PJ (2006) Hydrogen peroxide as a paracrine vascular mediator: regulation and signaling leading to dysfunction. ExpBiol Med (Maywood) 231:237–51.

Arden SL, Lambert DM (1997) Is the black robin in genetic peril? Mol Ecol 6: 21–8.

Arendt J (1988) Melatonin. Clin Endocrinol (Oxf) 29: 205–29.

Argos P, Kamer G, Nicklin MJ, Wimmer E (1984) Similarity in gene organization and homology between proteins of animal picornaviruses and a plant comovirus suggest common ancestry of these virus families. Nucleic Acids Res 12: 7251–67.

Arias E, Cuervo AM (2011)Chaperone-mediated autophagy in protein quality control.Curr Opin Cell Biol 23: 184-9.

Arim M, Berazategui M, Barreneche JM, Ziegler L, Zarucki M, Abades SR (2011) Determinants of density-body size scaling within food webs and tools for their detection. Adv Ecol Res 45: 1–39.

Arimoto K, Fukuda H, Imajoh-Ohmi S, Saito H, Takekawa M (2008) Formation of stress granules inhibits apoptosis by suppressing stress-responsive MAPK pathways. Nat Cell Biol 10: 1324–32.

Aris-Brosou S, Yang Z (2002) Effects of models of rate evolution on estimation of divergence dates with special reference to the metazoan 18S ribosomal RNA phylogeny. Syst Biol 51: 703–14.

Aris-Brosou S, Yang Z (2003) Bayesian models of episodic evolution support a late precambrian explosive diversification of the metazoa. Mol Biol Evol 20: 1947–54.

Arita T, Suzuki R (2000) Interactions between learning and evolution: the outstanding strategy generated by the Baldwin effect. Artif Life 7: 196–205.

Arkhipova I, Meselson M (2000) Transposable elements in sexual and ancient asexual taxa. Proc Natl Acad Sci USA 97: 14473–7.

Arkhipova I, Meselson M (2005) Deleterious transposable elements and the extinction of asexuals. Bioessays 27: 76–85.

Armagan A, Uz E, Yilmaz HR, Soyupek S, Oksay T, Ozcelik N (2006)Effects of melatonin on lipid peroxidation and antioxidant enzymes in streptozotocin-induced diabetic rat testis.Asian J Androl 8: 595-600.

Armbrust EV, Galindo HM (2001) Rapid evolution of a sexual reproduction gene in centric diatoms of the genus Thalassiosira. Appl Environ Microbiol 67: 3501–13.

Armelin HA, Armelin MCS, Kelly K, Stewart T, Leder P, et al. (1984) Functional role for c-myc in mitogenic response to platelet-derived growth factor. Nature 310: 655-60.

Armengol X, Boronat L, Camacho A, Wurtsbaugh WA (2001) Grazing by a dominant rotifer Conochilus unicornis Rousselet in a mountain lake: in situ measurements with synthetic microspheres. Hydrobiologia 446: 107–14.

Armitage JA, Taylor PD, Poston L (2005)Experimental models of developmental programming: consequences of exposure to an energy rich diet during development.J Physiol 565: 3-8.

Armstrong DT (1986) Environmental stress and ovarian function. Biol. Reprod. 34: 29-9.

Armstrong JS, Bivalacqua TJ, Chamulitrat W, Sikka S, Hellstrom WJG (2002) A comparison of the NADPH oxidase in human sperm and white blood cells. Int J Androl 25: 223–9.

Armstrong R, Gilpin M (1977) Evolution in a time-varying environment. Science 195: 591-2.

Arnaud-Haond S, Duarte CM, Diaz-Almela E, Marbà N, Sintes T, Serrão EA (2012) Implications of extreme life span in clonal organisms: Millenary clones in meadows of the threatened seagrass Posidonia oceanica. PLoS ONE 7: e30454.

Arnault C, Dufournel I (1994) Genome and stresses : reactions against aggressions, behavior of transposable elements. Genetica 93: 149–60.

Arney KL, Bao S, Bannister AJ, Kouzarides T, Surani MA (2002) Histone methylation defines epigenetic asymmetry in the mouse zygote. Int J Dev Biol 46: 317–20.

Arnheim N, Calabrese P, Tiemann-Boege I (2007) Mammalian meiotic recombination hot spots. Annu Rev Genet 41: 369-99.

Arnheim N, Calabrese P (2009) Understanding what determines the frequency and pattern of human germline mutations. Nat Rev Genet 10: 478-88.

Arnold ML (2007) Evolution through genetic exchange. Oxford, UK: Oxford University Press.

Arnold SA, Brekken RA (2009) SPARC: a matricellular regulator of tumorigenesis. J Cell Commun Signal 3: 255–73.

Arnqvist G, Rowe L (1995) Sexual conflict and arms races between the sexes: a morphological adaptation for control of mating in a female insect. Proc R Soc Lond Ser B 261: 123–7.

Arnqvist G, Nilsson T (2000) The evolution of polyandry: multiple mating and female fitness in insects. Anim Behav 60: 145–64.

Arnqvist G, Edvardsson M, Friberg U, Nilsson T (2000) Sexual conflict promotes speciation in insects. Proc Natl Acad Sci USA 97: 10460-4.

Arnqvist G, Rowe L (2002)Antagonistic coevolution between the sexes in a group of insects.Nature 415: 787-9.

Arnqvist G, Rowe L (2005) Sexual conflict. Princeton, NJ: Princeton University Press.

Artenstein MS, Miller WS (1966) Air sampling for respiratory disease agents in army recruits. Bacteriol Rev 30: 571–72.

Arndt H (1993) Rotifers as predators on components of the microbial web (bacteria, heterotrophic flagellates, ciliates) – a review. Hydrobiologia 255: 231–46.

Arunkumar KP, Metta M, Nagaraju J (2006) Molecular phylogeny of silkmoths reveals the origin of domesticated silkmoth, Bombyx mori from Chinese Bombyx mandarina and paternal inheritance of Antheraea proylei mitochondrial DNA. Mol Phylogenet Evol 40: 419–27.

Aruoma OI, Kaur H, Halliwell B (1991a)Oxygen free radicals and human diseases.J R Soc Health 111: 172-7.

Aruoma OI, Halliwell B, Gajewski E, Dizdaroglu M (1991b)Copper-ion-dependent damage to the bases in DNA in the presence of hydrogen peroxide.Biochem J 273: 601-4.

Asa C, Miller P, Agnew M, Rebolledo JAR, Lindsey SL, et al. (2007) Relationship of inbreeding with sperm quality and reproductive success in Mexican gray wolves. Anim Conserv 10: 326–31.

Asahi KI, Sakurai A, Takahashi N, Kubohara Y, Okamoto K, Tanaka Y (1995) DIF-1, morphogen of Dictyostelium discoideum, induces the erythroid differentiation in murine and human leukemia cells. Biochem Biophys Res Commun 208: 1036–9.

Asayama K, Dobashi K, Hayashibe H, Megata Y, Kato K (1987) Lipid peroxidation and free radical scavengers in thyroid dysfunction in the rat: a possible mechanism of injury to heart and skeletal muscle in hyperthyroidism. Endocrinology 121: 2112–8.

Ash P (1980) The influence of radiation on fertility in man. Br. J Radiol 53: 271-8.

Ashburner M (1992) Drosophila, a laboratory handbook. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press.

Ashby WR (1954) Design for a brain. New York, NY: John Wiley & Sons.

Ashby WR (1956) An introduction to cybernetics. London, UK: Chapman & Hall.

Ashe A, Whitelaw E (2007) Another role for RNA: a messenger across generations. Trends Genet 23: 8-10.

Ashe A, Sapetschnig A, Weick EM, Mitchell J, Bagijn MP, et al. (2012) piRNAs can trigger a multigenerational epigenetic memory in the germline of C. elegans. Cell 150: 88–99.

Asher JH Jr (1970) Parthenogenesis and genetic variability. II. One-locus models for various diploid populations. Genetics 66: 369-91.

Ashley MV, Laipis PJ, Hauswirth WW (1989) Rapid segregation of heteroplasmic bovine mitochondria. Nucleic Acids Res 17: 7325–31.

Ashwood-Smith MJ, Edwards RG (1996) DNA repair by oocytes. Mol Hum Reprod 2: 46–51.

Asker SE, Jerling L (1992) Apomixis in plants. Boca Raton, FL: CRC Press.

Aslan JE, Thomas G (2009)Death by committee: organellar trafficking and communication in apoptosis.Traffic 10: 1390-404.

Aspbury AS, Gabor CR (2004) Discriminating males alter sperm production between species. Proc Natl Acad Sci USA 101: 15970-3.

Assis R, Zhou Q, Bachtrog D (2012)Sex-biased transcriptome evolution in Drosophila.Genome Biol Evol 4: 1189-200.

Asur RS, Thomas RA, Tucker JD (2009)Chemical induction of the bystander effect in normal human lymphoblastoid cells.Mutat Res 676: 11-6.

Asur R, Balasubramaniam M, Marples B, Thomas RA, Tucker JD (2010)Involvement of MAPK proteins in bystander effects induced by chemicals and ionizing radiation.Mutat Res 686: 15-29.

Atanassova NN, Walker M, McKinnell C, Fisher JS, Sharpe RM (2005) Evidence that androgens and oestrogens, as well as follicle-stimulating hormone, can alter Sertoli cell number in the neonatal rat. J Endocrinol 184: 107-17.

Atchley WR (1977a)Evolutionary consequences of parthogenesis: evidence from the Warramaba virgo complex.Proc Natl Acad Sci USA 74: 1130-4.

Atchley WR (1977b) Biological variability in the parthenogenetic grasshopper Warramaba virgo(Key) and its sexual relatives. I. The eastern Australian populations. Evolution 31: 782-99.

Atchley WR (1978) Biological variability in the parthenogenetic grasshopper Warramaba virgo (Key) and its sexual relatives. II. The western Australian taxa. Evolution 32: 375-88.

Atchley WR, Fitch WM (1995) Myc and Max: molecular evolution of a family of proto-oncogene products and their dimerization partner. Proc Natl Acad Sci USA 92: 10217-21.

Atessahin A, Sahna E, Türk G, Ceribasi AO, Yilmaz S, et al. (2006) Chemoprotective effect of melatonin against cisplatin-induced testicular toxicity in rats. J Pineal Res 41: 21-7.

Atkinson WD (1979) A field investigation of larval competition in domestic Drosophila. J Anim Ecol 48: 91–102.

Atkinson D (1991) Sexual showiness and parasite load: correlations without parasite coevolutionary cycles. J Theor Biol 150: 251-60.

Atkinson SR, Marguerat S, Bähler J (2012) Exploring long non-coding RNAs through sequencing. Semin Cell Dev Biol 23: 200-5.

At-Taras EE, Berger T, McCarthy MJ, Conley AJ, Nitta-Oda BJ, Roser JF (2006) Reducing estrogen synthesis in developing boars increases testis size and total sperm production. J Androl 27: 552-9.

Atwood KC, Schneider LK, Ryan FJ (1951) Periodic selection in Escherichia coli. Proc Natl Acad Sci USA 37: 146–55.

Aubin-Horth N, Renn SC (2009) Genomic reaction norms: using integrative biology to understand molecular mechanisms of phenotypic plasticity. Mol Ecol 18: 3763-80.

Audhya T, Hollander CS, Schlesinger DH, Hutchinson B (1989) Structural characterization and localization of corticotropin-releasing factor in testis. Biochim Biophys Acta 995: 10-6.

Aufsatz W, Mette MF, van der Winden J, Matzke AJM, Matzke M (2002) RNA-directed DNA methylation in Arabidopsis. Proc Natl Acad Sci USA 99: 16499–506.

Auld JR, Agrawal AA, Relyea RA (2010) Re-evaluating the costs and limits of adaptive phenotypic plasticity. Proc Biol Sci 277: 503-11.

Avery SV (2006) Microbial cell individuality and the underlying sources of heterogeneity. Nat Rev Microbiol 4: 577-87.

Avise JC (2008) Clonality. The genetics, ecology and evolution of sexual abstinence in vertebrate animals. Oxford, UK: Oxford University Press.

Avise JC, Quattro JM, Vrijenhoek RC (1992a) Molecular clones within organismal clones: mitochondrial DNA phylogenies and the evolutionary histories of unisexual vertebrates. Evol Biol 26: 225–46.

Avise JC, Bowen BW, Lamb T, Meylan AB, Bermingham E (1992b) Mitochondrial DNA evolution at a turtle’s pace: evidence for low genetic variability and reduced microevolutionary rate in the Testudines. Mol Biol Evol 9:457–73.

Avraham KB, Schickler M, Sapoznikov D, Yarom R, Groner Y (1988) Down's syndrome: abnormal neuromuscular junction in tongue of transgenic mice with elevated levels of human Cu/Zn-superoxide dismutase. Cell 54: 823-9.

Avraham KB, Sugarman H, Rotshenker S, Groner Y (1991) Down's syndrome: morphological remodelling and increased complexity in the neuromuscular junction of transgenic CuZn-superoxide dismutase mice. J Neurocytol 20: 208-15.

Awad H, Halawa F, Mostafa T, Atta H (2006) Melatonin hormone profile in infertile males. Int J Androl 29: 409-13.

Awadalla P, Gauthier J, Myers RA, Casals F, Hamdan FF, et al. (2011) Direct measure of the de novo mutation rate in autism and schizophrenia cohorts. Am J Hum Genet 87: 316–24.

Awe S, Renkawitz-Pohl R (2010) Histone H4 acetylation is essential to proceed from a histone- to a protamine-based chromatin structure in spermatid nuclei of Drosophila melanogaster. Syst Biol Reprod Med 56: 44–61.

Axelsson E, Smith NG, Sundström H, Berlin S, Ellegren H (2004) Male-biased mutation rate and divergence in autosomal, Z-linked and W-linked introns of chicken and turkey. Mol Biol Evol 21:1538-47.

Ayala FJ (1965) Evolution of fitness in experimental populations of Drosophila serrata. Science 150: 903-5.

Ayala FJ (1966) Evolution of fitness. I. Improvement in the productivity and size of irradiated populations of Drosophila serrata and Drosophila birchii. Genetics 53: 883-95.

Ayala F J (1967) Evolution of fitness. III. Improvement of fitness in irradiated populations of Drosophila serrata. Proc Natl Acad Sci USA 58: 1919-23.

Ayala FJ (1968a) Genotype, environment, and population numbers. Science 162: 1453-9.

Ayala FJ (1968b) Evolution of fitness. II. Correlated effects of natural selection on the productivity and size of experimental populations of Drosophila serrate. Evolution 22: 55-65.

Ayala FJ (1968c) Biology as an autonomous science. Am Sci 56: 207-21.

Ayala FJ (1969) Evolution of fitness. V. Rate of evolution in irradiated populations of Drosophila. Proc Natl Acad Sci USA 63: 790-3.

Ayala FJ (2000) Neutralism and selectionism: the molecular clock.Gene 261: 27-33.

Ayala FJ, Campbell CA (1974) Frequency-dependent selection. Annu Rev Ecol Syst 5: 115–38.

Aydilek N, Aksakal M, Karakilcik AZ (2004) Effects of testosterone and vitamin E on the antioxidant system in rabbit testis. Andrologia 36: 277–81.

Ayre EA, Pang SF (1994) 2-(125I)iodomelatonin binding sites in the testis and ovary: putative melatonin receptors in the gonads. Biol Signals 3: 71–84.

Azam S, Jouvet N, Jilani A, Vongsamphanh R, Yang X, et al. (2008) Human glyceraldehyde-3-phosphate dehydrogenase plays a direct role in reactivating oxidized forms of the DNA repair enzyme APE1. J Biol Chem 283: 30632-41.

Azoulay-Dupuis E, Rieux V, Rivier C, Trombe MC (1998) Pleiotropic mutations alter the kinetics of calcium transport, competence regulation, autolysis and experimental virulence in Streptococcus pneumoniae. Res Microbiol 149: 5-13.

Azuma Y, Kaji K, Katogi R, Takeshita S, Kudo A (2000) Tumor necrosis factor-alpha induces differentiation of and bone resorption by osteoclasts. J Biol Chem 275: 4858-64.

Azzu V, Jastroch M, Divakaruni AS, Brand MD (2010) The regulation and turnover of mitochondrial uncoupling proteins. Biochim Biophys Acta 1797: 785-91.

Baake E, Gabriel W (2000) Biological evolution through mutation, selection, and drift: an introductory review. Annu Rev Comp Phys 7: 203–64.

Baake E, Wagner H (2001) Mutation-selection models solved exactly with methods of statistical mechanics. Genet Res (Camb.) 78: 93–117.

Baali-Cherif D, Besnard G (2005) High genetic diversity and clonal growth in relict populations of Olea europaea subsp. laperrinei (Oleaceae) from Hoggar, Algeria. Ann Bot 96: 823–30.

Baarends WM, Roest HP, Grootegoed JA (1999) The ubiquitin system in gametogenesis. Mol Cell Endocrinol 151: 5–16.

Baarends WM, van der Laan R, Grootegoed JA (2000) Specific aspects of the ubiquitin system in spermatogenesis. J Endocrinol Invest 23: 597–604.

Baarends WM, van der Laan R, Grootegoed JA (2001) DNA repair mechanisms and gametogenesis. Reproduction 121: 31-9.

Babar IA, Slack FJ, Weidhaas JB (2008)miRNA modulation of the cellular stress response.Future Oncol 4: 289-98.

Babar IA, Czochor J, Steinmetz A,Weidhaas JB, Glazer PM, Slack FJ (2011) Inhibition of hypoxia-induced miR-155 radiosensitizes hypoxic lung cancer cells. Cancer Biol Ther 12: 908–14.

Babbs CF (1990) Free radicals and the etiology of colon cancer. Free Radic Biol Med 8: 191-200.

Babcock GT, Wikström M (1992)Oxygen activation and the conservation of energy in cell respiration.Nature 356: 301-9.

Baccarelli A, Bollati V (2009) Epigenetics and environmental chemicals. Curr Opin Pediatr 21: 243–51.

Bachl J, Carlson C, Gray-Schopfer V, Dessing M, Olsson C (2001) Increased transcription levels induce higher mutation rates in a hypermutating cell line. J Immunol 166: 5051–7.

Bachtrog D (2003) Adaptation shapes patterns of genome evolution on sexual and asexual chromosomes in Drosophila. Nat Genet 34: 215–9.

Bachmann K (1983) Evolutionary genetics and the control of morphogenesis in flowering plants. Evol Biol 16: 157-208.

Bachmann-Waldmann C, Jentsch S, Tobler H, Muller F (2004) Chromatin diminution leads to rapid evolutionary changes in the organization of the germline genomes of the parasitic nematodes A. suum and P. univalens. Mol Biochem Parasitol 134: 53–64.

Bachtrog D (2008) Evidence for male-driven evolution in Drosophila. Mol Biol Evol 25: 617–9.

Bachtrog D, Charlesworth B (2002) Reduced adaptation of a non-recombining neo-Y chromosome. Nature 416: 323–6.

Bacolla A, Larson JE, Collins JR, Li J, Milosavljevic A, et al. (2008) Abundance and length of simple repeats in vertebrate genomes are determined by their structural properties. Genome Res 18: 1545-53.

Bacolla A, Wells RD (2009) Non-B DNA conformations as determinants of mutagenesis and human disease. Mol Carcinog 48: 273-85.

Badyaev AV (2005a) Stress-induced variation in evolution: from behavioural plasticity to genetic assimilation. Proc R Soc Lond Ser B Biol Sci 272: 877–86.

Badyaev AV (2005b) Role of stress in evolution: from individual adaptability to evolutionary adaptation. In: Hallgrímsson B, Hall BK, eds. Variation: A Central Concept in Biology.Burlington, MA: Elsevier Academic Press. pp 277–302.

Badyaev AV (2009)Evolutionary significance of phenotypic accommodation in novel environments: an empirical test of the Baldwin effect.Philos Trans R Soc Lond B Biol Sci 364: 1125-41.

Badyaev AV, Ghalambor CK (1998) Does a trade-off exist between sexual ornamentation and ecological plasticity? Sexual dichromatism and occupied elevational range in finches. Oikos 82: 319–24.

Badyaev AV, Ghalambor, CK (2001) Evolution of life histories along elevational gradients: trade off between parental care and fecundity. Ecology 82: 2948-60.

Bae I, Fan S, Meng Q, Rih JK, Kim HJ, et al. (2004) BRCA1 induces antioxidant gene expression and resistance to oxidative stress. Cancer Res 64: 7893–909.

Bae YS, Kang SW, Seo MS, Baines IC, Tekle E, et al. (1997) Epidermal growth factor (EGF)-induced generation of hydrogen peroxide. J Biol Chem 272: 217–21.

Baek D, Villen J, Shin C, Camargo FD, Gygi SP, Bartel DP (2008) The impact of microRNAs on protein output. Nature 455: 64-71.

Baer CF (2008) Does mutation rate depend on itself? PLoS Biol 6(2): e52.

Baer CF, Miyamoto MM, Denver DR (2007) Mutation rate variation in multicellular eukaryotes: causes and consequences. Nat Rev Genet 8: 619–31.

Bailey JA, Liu G, Eichler EE (2003) An Alu transposition model for the origin and expansion of human segmental duplications. Am J Hum Genet 73: 823–34.

Bailly A, Gartner A (2013) Germ cell apoptosis and DNA damage responses. Adv Exp Med Biol 757: 249-76.

Baird DD, Wilcox AJ, Weinberg CR (1986) Use of time to pregnancy to study environmental exposures. Am J Epidemiol 124: 470–80.

Bajoria R, Chatterjee R (2011)Hypogonadotrophic hypogonadism and diminished gonadal reserve accounts for dysfunctional gametogenesis in thalassaemia patients with iron overload presenting with infertility.Hemoglobin 35: 636-42.

Bajpai M, Gupta G, Setty BS (1998)Changes in carbohydrate metabolism of testicular germ cells during meiosis in the rat.Eur J Endocrinol 138: 322-7.

Bak P (1993) in Thinking about Biology, Santa Fe Institute Studies in the Sciences of Complexity Vol. III, edited by W. Stein and F. J. Varela (Addison-Wesley, Reading, MA) pp. 255–268.

Bak P (1996) How nature works: The science of self-organized criticality New York, NY: Copernicus.

Bak P, Tang C, Wiesenfeld K (1987) Self-organized criticality: An explanation of the 1/f noise. Phys Rev Lett 59: 381-4.

Bak P, Tang C, Wiesenfeld K (1988) Self-organized criticality. Phys Rev A 38: 364-74.

Baker CD, Neill WH Jr, Strawn K (1970) Sexual difference in heat resistance of Ozark minnow, Dionda nubila (Forbes). T Am Fish Soc 99: 588-91.

Baker HG (1955) Self-compatibility and establishment after ‘‘long-distance’’ dispersal. Evolution Int J Org Evolution 9: 347–8.

Baker HG (1965) Characteristics and modes of origin of weeds. In: Baker HG, Stebbins GL, eds.Genetics of colonizing species. New York, NY: Academic Press. pp 147–172.

Baker MA, Taylor YC, Brown JM (1988) Radiosensitization, thiol oxidation, and inhibition of DNA repair by SR 4077. Radiat Res 113: 346-55.

Baker MA, Krutskikh A, Curry BJ, Hetherington L, Aitken RJ (2005) Identification of cytochrome-b5 reductase as the enzyme responsible for NADH-dependent lucigenin chemiluminescence in human spermatozoa. Biol Reprod 73: 334–42.

Baker MA, Aitken RJ (2004) The importance of redox regulated pathways in sperm cell biology. Mol Cell Endocrinol 216: 47–54.

Baker MA, Aitken RJ (2005) Reactive oxygen species in spermatozoa: methods for monitoring and significance for the origins of genetic disease and infertility. Reprod Biol Endocrinol 3:67.

Baker ME (2006) The genetic response to Snowball Earth: role of HSP90 in the Cambrian Explosion. Geobiology 4: 11–14.

Baker NE (2011) Cell competition. Curr Biol 21: R11-5.

Baker TG (1963) A quantitative and cytological study of germ cells in human ovaries. Proc R Soc Lond B Biol Sci 158: 417–33.

Bakken AH, McClanahan M (1978)Patterns of RNA synthesis in early meiotic prophase oocytes from fetal mouse ovaries.Chromosoma 67: 21-40.

Bakker K (1961) An analysis of factors which determine the success in competition of food among larvae of Drosophila melanogaster. Arch Netherl Zool 14: 200–81.

Bakst MR, Wishart GJ, Brillard JP (1994) Oviductal sperm selection transport and storage in poultry. Poultry Sci Rev 5: 117–43.

Balaban NQ, Merrin J, Chait R, Kowalik L, Leibler S (2004) Bacterial persistence as a phenotypic switch. Science 305: 1622-5.

Baldwin JM (1896) A new factor in evolution. Am Nat 30: 441-51.

Baldwin JM (1902) Development and evolution. London, UK: Macmillan.

Bale JS Ponder KL, Pritchard J (2007) Coping with stress. In: Emden HFV, Harrington R, eds. Aphids as crop pests. Oxford, UK: CABI. pp 287-310

Ball BA, Vo AT, Baumber J (2001) Generation of reactive oxygen species by equine spermatozoa. Am J Vet Res 62: 508–15.

Ball MA, Parker GA (1996) Sperm competition games: External fertilization and “adapative” infertility. J Theor Biol 180: 141-50.

Ballard JWO (2000a) Comparative genomics of mitochondrial DNA in members of the Drosophila melanogaster subgroup. J Mol Evol 51: 48-63.

Ballard JWO (2000b) Comparative genomics of mitochondrial DNA in Drosophila simulans. J Mol Evol 51: 64-75.

Ballard JWO, Kreitman M (1994) Unravel- ing selection in the mitochondrial genome of Drosophila. Genetics 138: 757-72.

Ballard JWO, Whitlock MC (2004) The incomplete natural history of mitochondria. Mol Ecol 13: 729-44.

Baltimore D (1970) Viral RNA-dependent DNA polymerase. Nature 226: 1209–11.

Baluška F, Mancuso S (2007)Plant neurobiology as a paradigm shift not only in the plant sciences.Plant Signal Behav 2: 205-7.

Balyaev DK, Borodin PM (1982) The influence of stress on variation and its role in evolution. Biol Zentbl 100: 705–14.

Bamberger CM, Schulte HM, Chrousos GP (1996) Molecular determinants of glucocorticoid receptor function and tissue sensitivity to glucocorticoids. Endocr Rev 17: 245-61.

Bambino TH, Hsueh AJW (1981) Direct inhibitory effect of glucocorticoids upon testicular luteinizing hormone receptor and steroidogenesis in vivo and in vitro. Endocrinology 108: 2142-8.

Bandaru B, Gopal J, Bhagwat AS (1996) Overproduction of DNA cytosine methyltransferases causes methylation and C→T mutations at noncanonical sites. J Biol Chem 271: 7851–9.

Bánfi B, Molnár G, Maturana A, Steger K, Hegedûs B, et al. (2001) A Ca2+-activated NADPH oxidase in testis, spleen, and lymph nodes. J Biol Chem 276: 37594–601.

Banks S, King SA, Irvine DS, Saunders PT (2005) Impact of a mild scrotal heat stress on DNA integrity in murine spermatozoa. Reproduction 129: 505-14.

Banning A, Deubel S, Kluth D, Zhou Z, Brigelius-Flohe R (2005) The GI-GPx gene is a target for Nrf2. Mol Cell Biol 25: 4914–23.

Banwell KM (2009)Oxygen concentration during oocyte maturation in the mouse.Thesis. Adelaide, Australia: University of Adelaide.

Bao J, Yan W (2012)Male germline control of transposable elements.Biol Reprod 86: 162, 1-14.

Baquero MR, Nilsson AI, Turrientes MC, Sandvang D, Galan JC, et al. (2004) Polymorphic mutation frequencies in Escherichia coli: emergence of weak mutators in clinical isolates. J Bacteriol 186: 5538–42.

Baranano DE, Rao M, Ferris CD, Snyder SH (2002) Biliverdin reductase: a major physiologic cytoprotectant. Proc Natl Acad Sci USA 99: 16093–8.

Barckmann B, Simonelig M (2013)Control of maternal mRNA stability in germ cells and early embryos.Biochim Biophys Acta 1829: 714-24.

Bardwell L, Cook JG, Inouye CJ, Thorner J (1994) Signal propagation and regulation in the mating pheromone response pathway of the yeast Saccharomyces cerevisiae. Dev Biol 166: 363-79.

Barford D (2004)The role of cysteine residues as redox-sensitive regulatory switches.Curr Opin Struct Biol 14: 679-86.

Barja G (2007) Mitochondrial oxygen consumption and reactive oxygen species production are independently modulated: implications for aging studies. Rejuv Res 10: 215-24.

Barkett M, Gilmore TD (1999) Control of apoptosis by Rel/NF-kappaB transcription factors. Oncogene 18: 6910-24.

Barlev NA, Sayan BS, Candi E, Okorokov AL (2010) The microRNA and p53 families join forces against cancer. Cell Death Differ 17: 373-5.

Barlow C, Liyanage M, Moens PB, Tarsounas M, Nagashima K, et al. (1998) Atm deficiency results in severe meiotic disruption as early as leptonema of prophase I. Development 125: 4007–17.

Barlow DP (1993) Methylation and imprinting: from host defense to gene regulation? Science 260: 309–10.

Barnes AI, Partridge L (2003) Costing reproduction. Anim Behav 66: 199–204.

Barnett L (1998) Ruggedness and neutrality: The NKp family of fitness landscapes. In: AdamiC, Belew RK, Kitano H, Taylor C, eds. Artificial Life VI: Proceedings of the Sixth International Conference on Artificial Life. Cambridge, MA: MIT Press.pp 18–27.

Baroiller JF, D’Cotta H, Bezault E, Wessels S, Hoerstgen-Schwark G (2008) Tilapia sex determination: Where temperature and genetics meet. Comp Biochem Physiol A Mol Integr Physiol 153: 30–8.

Barón SB, Bückle RF, Espina S (2002) Environmental factors and sexual differentiation in Poecilia sphenops Valenciennes (Pisces: Poeciliidae). Aquacult Res 33: 615–9.

Barr CM, Neiman M, Taylor DR (2005) Inheritance and recombination of mitochondrial genomes in plants, fungi and animals. New Phytol 168: 39–50.

Barraclough TG, Birky CW Jr, Burt A (2003) Diversification in sexual and asexual organisms. Evolution 57: 2166–72.

Barraclough TG, Fontaneto D, Ricci C, Herniou EA (2007) Evidence for inefficient selection against deleterious mutations in cytochrome oxidase I of asexual bdelloid rotifers. Mol Biol Evol 24: 1952–62.

Barres BA, Raff MC (1999) Axonal control of oligodendrocyte development. J Cell Biol 147: 1123–8.

Barreto G, Schäfer A, Marhold J, Stach D, Swaminathan SK, et al. (2007) Gadd45a promotes epigenetic gene activation by repair-mediated DNA demethylation. Nature 445: 671–5.

Barrett RDH, MacLean RC, Bell G (2005) Experimental evolution of Pseudomonas fluorescens in simple and complex environments. Am Nat 160: 470–80.

Barrett RDH, Bell G (2006)The dynamics of diversification in evolving Pseudomonas populations.Evolution 60: 484-90.

Barrett RDH, Schluter D (2008) Adaptation from standing genetic variation. Trends Ecol Evol 23: 38–44.

Barrett LG, Thrall PH, Dodds PN, van der Merwe M, Linde CC, et al. (2009) Diversity and evolution of effector loci in natural populations of the plant pathogen Melampsora lini. Mol Biol Evol 26: 2499–513.

Barrett SCH (2002) The evolution of plant sexual diversity. Nat Rev Genet 3: 274–84.

Barrick JE, Lenski RE (2009)Genome-wide mutational diversity in an evolving population of Escherichia coli. Cold Spring Harb Symp Quant Biol 74: 119–29.

Barrick JE, Yu DS, Yoon SH, Jeong H, Oh TK, et al. (2009) Genome evolution and adaptation in a long-term experiment with Escherichia coli.Nature 461: 1243-7.

Barrick JE, Kauth MR, Strelioff CC, Lenski RE (2010) Escherichia coli rpoB mutants have increased evolvability in proportion to their fitness defects. Mol Biol Evol 27: 1338–47.

Barritt JA, Brenner CA, Cohen J, Matt DW (1999) Mitochondrial DNA rearrangements in human oocytes and embryos. Mol Hum Reprod 5: 927-33

Barroso G, Morshedi M, Oehninger S (2000) Analysis of DNA fragmentation, plasma membrane translocation of phosphatidylserine and oxidative stress in human spermatozoa.Hum Reprod 15: 1338-44.

Barry G, Mattick JS (2012) The role of regulatory RNA in cognitive evolution. Trends Cogn Sci 16: 497–503.

Barry RG (1992) Mountain weather and climate. New York, NY: Routledge.

Bartek J, Lukas J (2003) Chk1 and Chk2 kinases in checkpoint control and cancer. Cancer Cell 3: 421–9.

Bartel DP (2004) MicroRNAs: genomics, biogenesis, mechanism, and function. Cell 116: 281–97.

Bartel DP, Chen CZ (2004) Micromanagers of gene expression: the potentially widespread influence of metazoan microRNAs. Nat Rev Genet 5: 396–400.

Barthelemy C, Khalfoun B, Guillaumin JM, Lecomte P, Bardos P (1988) Seminal fluid transferrin as an index of gonadal function in men. J Reprod Fertil 82: 113-8.

Bartke A (1995) Apoptosis of male germ cells, a generalized or cell type-specific phenomenon? Endocrinology 136:3–4.

Bartkova J, Horejsí Z, Koed K, Krämer A, Tort F, et al. (2005a)DNA damage response as a candidate anti-cancer barrier in early human tumorigenesis.Nature 434: 864-70.

Bartkova J, Bakkenist CJ, Rajpert-De Meyts E, Skakkebaek NE, Sehested M, et al. (2005b)ATM activation in normal human tissues and testicular cancer.Cell Cycle 4: 838-45.

Bartkova J, Rezaei N, Liontos M, Karakaidos P, Kletsas D, et al. (2006) Oncogene-induced senescence is part of the tumorigenesis barrier imposed by DNA damage checkpoints. Nature 444: 633–7.

Barto AG, Sutton RS, Anderson CW (1983) Neurolike adaptive elements that can solve difficult learning control problems. IEEE Trans Syst Man Cybern 13: 834-46.

Bartolome C, Maside X, Charlesworth B (2002) On the abundance and distribution of transposable elements in the genome of Drosophila melanogaster. Mol Biol Evol 19: 926–37.

Barton NH (1995) A general model for the evolution of recombination. Genet Res 65: 123–44.

Barton N (1998) The geometry of natural selection. Nature 395: 751–2.

Barton NH (2009)Why sex and recombination?Cold Spring Harb Symp Quant Biol 74:187-95.

Barton NH, Turelli M (1989) Evolutionary quantitative genetics: how little do we know? Annu Rev Genet 23: 337–70.

Barton NH, Charlesworth B (1998) Why sex and recombination? Science 281: 1986–90.

Barton N, Partridge L (2000) Limits to natural selection. BioEssays 22: 1075-84.

Barton NH, Keightley PD (2002) Understanding quantitative genetic variation. Nat Rev Genet 3: 11–21.

Barton NH, Turelli M (2004) Effects of genetic drift on variance components under a general model of epistasis. Evolution 58: 2111-32.

Barton NH, Otto SP (2005) Evolution of recombination due to random drift. Genetics 169: 2353–70.

Bartosch-Härlid A, Berlin S, Smith NGC,Møller AP, Ellegren H (2003) Life history and the male mutation bias. Evolution 57: 2398–406.

Bartosz G (2009)Reactive oxygen species: destroyers or messengers?Biochem Pharmacol 77: 1303-15.

Bartsh HE, Hietanen E, Malaveille C (1989) Carcinogenic nitrosamines: free radical aspects of their action. Free Rad Biol Med 7: 637-44.

Barzilai A, Yamamoto K (2004)DNA damage responses to oxidative stress.DNA Repair (Amst) 3:1109-15.

Basak S, Hoffmann A (2008) Crosstalk via the NF-κB signaling system. Cytokine Growth Factor Rev 19: 187–97.

Bashan N, Burdett E, Hundal HS, Klip A (1992) Regulation of glucose transport and glucose transport and Glut 1 glucose transporter expression by O2 in muscle cells in culture. Am J Physiol 262: 682–90.

Bashey F (2006) Cross-generational environmental effects and the evolution of offspring size in the Trinidadian guppy Poecilia reticulata. Evolution 60: 348–61.

Basini G, Bianco F, Grasselli F, Tirelli M, Bussolati S, Tamanini C (2004a) The effects of reduced oxygen tension on swine granulosa cell. Regul Pept 120:69–75.

Basini G, Grasselli F, Bianco F, Tirelli M, Tamanini C (2004b)Effect of reduced oxygen tension on reactive oxygen species production and activity of antioxidant enzymes in swine granulosa cells.Biofactors 20: 61-9.

Basini G, Bussolati S, Santini SE, Bianchi F, Careri M, et al.(2007)Antiangiogenesis in swine ovarian follicle: a potential role for 2-methoxyestradiol.Steroids 72: 660-5.

Bassar RD, Lopez-Sepulcre A, Walsh MR, Turcotte MM, Torres-Mejia M, Reznick DN (2010) Bridging the gap between ecology and evolution: integrating density regulation and life-history evolution. Ann NY Acad Sci 1206: 17–34.

Bassus S, Herkert O, Kronemann N, Görlach A, Bremerich D, et al. (2001) Thrombin causes vascular endothelial growth factor expression in vascular smooth muscle cells: role of reactive oxygen species. Arterioscler Thromb Vasc Biol 21: 1550-5.

Bataillon T (2000) Estimation of spontaneous genome-wide mutation rate parameters: whither beneficial mutations? Heredity 84: 497–501.

Bataillon T, Roumet P, Poirier S, David J (2000) Measuring genome-wide spontaneous mutation in Triticum durum: implications for long-term selection response and management of genetic resources. In: Gallais A, Dillmann C, Goldringer I, eds. Eucarpia: quantitative genetics and breeding methods—the way ahead. Versailles, France: INRA Editions. pp 251–257.

Bataillon T (2003) Shaking the ‘deleterious mutations’ dogma? Trends Ecol Evol 18: 315-7.

Batchelor TP, Briffa M (2010) Influences on resource-holding potential during dangerous group contests between wood ants. Anim Behav 80: 443-9.

Bateson G (1972) Steps to an ecology of mind. San Francisco, CA: Chandler Publishing Co.

Bateson P (2012)The impact of the organism on its descendants.Genet Res Int 2012: 640612.

Bateson PPG (1983) Mate choice. In: Bateson PPG, ed. Optimal Out-Breeding. Cambridge, UK: Cambridge University Press.pp. 257–277.

Batschelet E, Domingo E, Weissmann C (1976) The proportion of revertant and mutant phage in a growing population, as a function of mutation and growth rate. Gene 1: 27–32.

Battaglia C, Genazzani AD, Regnani G, Primavera MR, Petraglia F, Volpe A (2000)Perifollicular Doppler flow and follicular fluid vascular endothelial growth factor concentrations in poor responders.Fertil Steril 74: 809-12.

Battista JR (1997) Against all odds: the survival strategies of Deinococcus radiodurans. Ann Rev Microbiol 51: 203-24.

Bauché F, Fouchard MH, Jégou B (1994)Antioxidant system in rat testicular cells.FEBS Lett 349:392-6.

Baud V, Karin M (2001)Signal transduction by tumor necrosis factor and its relatives.Trends Cell Biol 11: 372-7.

Baudat F, Manova K, Yuen JP, Jasin M, Keeney S (2000) Chromosome synapsis defects and sexually dimorphic meiotic progression in mice lacking Spo11. Mol Cell 6: 989–98.

Baudat F, Buard J, Grey C, Fledel-Alon A, Ober C, et al. (2010)PRDM9 is a major determinant of meiotic recombination hotspots in humans and mice.Science 327: 836-40.

Baum JS, St George JP, McCall K (2005) Programmed cell death in the germline. Semin Cell Dev Biol 16: 245–59.

Baumann CG, Morris DG, Sreenan JM, Leese HJ (2007) The quiet embryo hypothesis: molecular characteristics favoring viability. Mol Reprod Dev 74: 1345–53.

Baur B, Hanselmann K, Schlimme W, Jenni B (1996) Genetic transformation in freshwater: Escherichia coli is able to develop natural competence. Appl Environ Microbiol 62: 3673-8.

Bavoux C, Hoffmann JS, Cazaux C (2005)Adaptation to DNA damage and stimulation of genetic instability: the double-edged sword mammalian DNA polymerase kappa.Biochimie 87: 637-46.

Bayliss CD, Field D, Moxon ER (2001) The simple sequence contingency loci of Haemophilus influenzae and Neisseria meningitidis. J Clin Invest 107: 657-62.

Bazin E, Glemin S, Galtier N (2006) Population size does not influence mitochondrial genetic diversity in animals. Science312: 570–2.

Bazykin GA, Kondrashov AS (2011)Detecting past positive selection through ongoing negative selection.Genome Biol Evol 3: 1006-13.

Beardmore J (1983) Extinction, survival, and genetic variation. In: Schonewald-Cox CM, Chambers SM, MacBryde F, Thomas L, eds. Genetics and conservation: a reference for managing wild animal and plant populations. Menlo Park, CA: Benjamin/Cummings.pp 125–151.

Beasley DWC, Davis CT, Guzman H, Vanlandingham DL, Travassos da Rosa APA, et al. (2003) Limited evolution of West Nile virus has occurred during its southwesterly spread in the United States. Virology 309: 190–5.

Beaton MJ, Hebert PDN (1988) Geographic parthenogenesis and polyploidy in Daphnia pulexLeydig. Am Nat132:837–45.

Beatty GE, McEvoy PM, Sweeney O, Provan J (2008) Range-edge effects promote clonal growth in peripheral populations of the one-sided wintergreen Orthilla secunda. Divers Distrib 14: 546–55.

Beatty J (1984) Chance and natural selection. Philos Sci 51: 183–211.

Beatty J, Finsen S (1989) Rethinking the propensity interpretation -- a peek inside Pandora’s box. In: Ruse M, ed. What the philosophy of biology is. Dordrecht, the Netherlands: Kluwer Publishers. pp 17–30.

Beaudet AL, Jiang YH (2002) A rheostat model for a rapid and reversible form of imprinting-dependent evolution. Am J Hum Genet 70: 1389–97.

Beaumont HJ, Gallie J, Kost C, Ferguson GC, Rainey PB (2009) Experimental evolution of bet hedging. Nature 462: 90–3.

Beaumont HM,Mandl AM (1961) A quantitative and cytological study of oogonia and oocytes in the foetal and neonatal rat. Proc R SocLond B 155: 557–79.

Beaupré CE, Tressler CJ, Beaupré SJ, Morgan JL, Bottje WG, Kirby JD(1997) Determination of testis temperature rhythms and effects of constant light on testicular function in the domestic fowl (Gallus domesticus). Biol Reprod 56: 570–5.

Beck ARP, Miller IJ, Anderson P, Streuli M (1998) RNA-binding protein TIAR is essential for primordial germ cell development. Proc Natl Acad Sci USA 95: 2331–6.

Beck R, Verrax J, Gonze T, Zappone M, Pedrosa RC, et al. (2009) Hsp90 cleavage by an oxidative stress leads to its client proteins degradation and cancer cell death. Biochem Pharmacol 77: 375-83.

Beck R, Dejeans N, Glorieux C, Pedrosa RC, Vásquez D, et al. (2011) Molecular chaperone Hsp90 as a target for oxidant-based anticancer therapies. Curr Med Chem 18: 2816-25.

Beck R, Dejeans N, Glorieux C, Creton M, Delaive E, et al. (2012) Hsp90 is cleaved by reactive oxygen species at a highly conserved N-terminal amino acid motif. PLoS One 7: e40795.

Beck V, Jaburek M, Demina T, Rupprecht A, Porter RK, et al. (2007) Polyunsaturated fatty acids activate human uncoupling proteins 1 and 2 in planar lipid bilayers. FASEB J 21: 1137-44.

Becker HJ (1974) Mitotic recombination maps in Drosophila melanogaster. Naturwissenschaften 61: 441-8.

Becker J, Schwaab R, Moller Taube A, Schwaab U, Schmidt W, et al. (1996) Characterization of the factor VIII defect in 147 patients with sporadic hemophilia A: family studies indicate a mutation type-dependent sex ratio of mutation frequencies. Am J Hum Genet 58: 657–70.

Beckman KB, Ames BN (1997) Oxidative decay of DNA. J Biol Chem 272: 13300-5.

Beckman RA (2009) Mutator mutations enhance tumorigenic efficiency across fitness landscapes. PLoS One 4: e5860.

Beckman RA, Loeb LA (2006) Efficiency of carcinogenesis with and without a mutator mutation. Proc Natl Acad Sci USA 103: 14140–5.

Becks L, Agrawal AF (2010) Higher rates of sex evolve in spatially heterogeneous environments. Nature 468: 89-92.

Becks L, Agrawal AF (2011) The effect of sex on the mean and variance of fitness in facultatively sexual rotifers. J Evol Biol 24: 656–64.

Becks L, Agrawal AF (2012) The evolution of sex is favoured during adaptation to new environments. PLoS Biol 10: e1001317.

Bedau MA, Packard NH (2003) Evolution of evolvability via adaptation of mutation rates. BioSystems 69: 143-62.

Bedford JM (1977) Evolution of the scrotum: the epididymis as the prime mover? In: Calaby JH, Tyndale-Biscoe CH, eds. Reproduction and evolution. Canberra, Australia: Australian Academy of Science. pp 171–182.

Bedford JM (2004) Enigmas of mammalian gamete form and function. Biol Rev 79: 429–60.

Bednarek J, Wysocki H, Sowinski J (2004) Oxidation products and antioxidant markers in plasma of patients with Graves’ disease and toxic multinodular goiter: effect of methimazole treatment. Free Radic Res 38: 659–64.

Beerenwinkel N, Antal T, Dingli D, Traulsen A, Kinzler KW, et al. (2007) Genetic progression and the waiting time to cancer. PLoS Comput Biol 3: e225.

Beermann S (1959) Chromatin-diminution bei Copepoden. Chromosoma 10: 504–14.

Begon M, Mortimer M (1986) Population Ecology: A Unified Study of Animals and Plants, 2nd edn. Boston, MA: Blackwell.

Begun DJ, Aquadro CF (1992) Levels of naturally occurring DNA polymorphism correlate with recombination rates in D. melanogaster. Nature 356: 519–20.

Begun DJ, Holloway AK, Stephens K, Hillier LW, Poh Y-P, et al. (2007) Population genomics: whole-genome analysis of polymorphism and divergence in Drosophila simulans. PLoS Biol 5: e310.

Behera N, Nanjundiah V (1995) An investigation into the role of phenotypic plasticity in evolution. J Theor Biol 172: 225–34.

Behera N, Nanjundiah V (2004) Phenotypic plasticity can potentiate rapid evolutionary change. J Theor Biol 226: 177–84.

Behl R, Pandey RS (2002) FSH induced stimulation of catalase activity in goat granulosa cells in vitro. Anim Reprod Sci 70: 215–21.

Behnke BJ, Kindig CA, Musch TI, Koga S, Poole DC (2001) Dynamics of microvascular oxygen pressure across the rest-exercise transition in rat skeletal muscle. Respir Physiol 126: 53–63.

Behrman HR, Aten RF (1991) Evidence that hydrogen peroxide blocks hormone-sensitive cholesterol transport into mitochondria of rat luteal cells. Endocrinology 128: 2958–66.

Behrman HR, Kodaman PH, Preston SL, Gao SP (2001) Oxidative stress and the ovary. J Soc Gynecol Invest 8 (Suppl. 1): S40–S42.

Beinert H, Kiley PJ (1999) Fe-S proteins in sensing and regulatory functions. Curr Opin Chem Biol 3: 152-7.

Beketov MA, Liess M (2007) Predation risk perception and food scarcity induce alterations of life-cycle traits of the mosquito Culex pipiens. Ecol Entomol 32: 405–10.

Bekker A, Holland HD, Wang PL, Rumble D III, Stein HJ, et al. (2004) Dating the rise of atmospheric oxygen. Nature 427: 117–20.

Bekker RM, Bakker JP, Grandin U, Kalamees R, Milberg P, et al. (1998) Seed size, shape and vertical distribution in the soil: indicators of seed longevity. Funct Ecol 12: 834-42.

BelAiba RS, Bonello S, Zähringer C, Schmidt S, Hess J, et al. (2007) Hypoxia up-regulates hypoxia-inducible factor-1alpha transcription by involving phosphatidylinositol 3-kinase and nuclear factor kappaB in pulmonary artery smooth muscle cells. Mol Biol Cell 18: 4691–7.

Belancio V, Roy-Engel AM (2011) Modulation of human mobile elements and genetic instability by environmental factors. In: Nriagu JO, ed. Encyclopedia of Environmental Health. Elsevier Science. pp 831–839.

Beldade P, Mateus AR, Keller RA (2011) Evolution and molecular mechanisms of adaptive developmental plasticity. Mol Ecol 20:1347-63.

Beletskii A, Bhagwat AS (1996) Transcription-induced mutations: increase in C to T mutations in the nontranscribed strand during transcription in Escherichia coli. Proc Natl Acad Sci USA 93: 13919–24.

Beletskii A, Bhagwat AS (2001) Transcription-induced cytosine-to-thymine mutations are not dependent on sequence context of the target cytosine. J Bacteriol 183: 6491–3.

Bell AW (1959) Enchytraeus fragmentosus, a new species of naturally fragmenting oligochaete worm. Science 129: 1278.

Bell EL, Klimova TA, Eisenbart J, Moraes CT, Murphy MP, et al.(2007) The Qo site of the mitochondrial complex III is required for the transduction of hypoxic signaling via reactive oxygen species production. J Cell Biol 177: 1029-36.

Bell G (1980) The costs of reproduction and their consequences. Am Nat 116: 45–76.

Bell G (1982) The masterpiece of nature: the evolution and genetics of sexuality. Berkeley, CA: The University of California Press.

Bell G (1985) Two theories of sex and variation.Experientia 41:1235-45.

Bell G (1988a) Sex and death in protozoa—the history of an obsession. Cambridge, UK: Cambridge University Press.

Bell G (1988b) Uniformity and diversity in the evolution of sex. In: Michod RE, Levin BR, eds. The evolution of sex. Sunderland, MA: Sinauer Associates. pp 126–138.

Bell G (2005) The evolution of evolution. Heredity 94: 1–2.

Bell G (2010) Fluctuating selection: the perpetual renewal of adaptation in variable environments. Phil Trans R Soc B 365: 87–97.

Bell G, Wolfe LM (1985) Sexual and asexual reproduction in a natural population of Hydra pseudoligactis. Can J Zool 63: 851-6.

Bell G, Koufopanou V (1986) The cost of reproduction. In: Dawkins R, Ridley M, eds. Oxford Surveys in Evolutionary Biology, vol. 3. Oxford, UK: Oxford University Press. pp 83–131.

Bell G, Koufopanou V (1991) The architecture of the life cycle in small organisms. Philos Trans R Soc Lond Ser B 332: 81-9.

Bell G, Reboud X (1997) Experimental evolution in Chlamydomonas. II. Genetic variation in strongly contrasted environments. Heredity 78: 498-506.

Bell G, Gonzalez A (2009) Evolutionary rescue can prevent extinction following environmental change. Ecol Lett 12: 942–8.

Bell JT, Spector TD (2011) A twin approach to unraveling epigenetics. Trends Genet 27: 116-25.

Bell PR (1996) Megaspore abortion: A consequence of selective apoptosis? Int J Plant Sci 157: 1–7.

Bellve AR, Zheng W (1989) Growth factors as autocrine and paracrine modulators of male gonadal functions. J ReprodFertil 85: 771–93.

Belyi VA, Ak P, Markert E, Wang H, Hu W, et al. (2010) The origins and evolution of the p53 family of genes. Cold Spring Harb Perspect Biol 2: a001198.

Ben-Ami F, Heller J (2005) Spatial and temporal patterns of parthenogenesis and parasitism in the freshwater snail Melanoides tuberculata. J Evol Biol 18: 138–46.

Ben-Ami F, Heller J (2007) Temporal patterns of geographic parthenogenesis in a freshwater snail. Biol J Linnean Soc 91: 711–8.

Ben-Ami F, Heller J (2008) Sex versus parasitism versus density. Biol J Linn Soc 93: 527–44.

Benda C (1891) Neue Mitteilungen über die Entwicklung der Genitaldrüsen und über die Metamorphose der Samenzellen (Histogenese der Spermatozoen). Arch Anat Physiol Physiol Abt: 549–52.

Bendall KE, Sykes BC (1995) Length heteroplasmy in the first hypervariable segment of the human mtDNA control region. Am J Hum Genet 57: 248–56.

Bendall KE, Macaulay VA, Baker JR, Sykes BC (1996) Heteroplasmic point mutations in the human mtDNA control region. Am J Hum Genet 59: 1276–87.

Bendall KE, Macaulay VA, Sykes BC (1997) Variable levels of a heteroplasmic point mutation in individual hair roots. Am J Hum Genet 61: 1303–8.

Benezra M, Chevallier N, Morrison DJ, MacLachlan TK, El-Deiry WS, Licht JD (2003) BRCA1 augments transcription by the NF-kappaB transcription factor by binding to the Rel domain of the p65/RelA subunit. J Biol Chem 278: 26333–41.

Bengtson S (2002) Origins and early evolution of predation. Paleont Soc Papers 8: 289–317.

Bengtson S, Morris SC (1992) Early radiation of biomineralizing phyla. In: Lipps JH, Signor PW, eds. Origin and Early Evolution of the Metazoa. New York, NY: Plenum Press. pp 447–481.

Bengtsson B (2003) Genetic variation in organisms with sexual and asexual reproduction. J Evol Biol 16: 189–199.

Bennett AF, Lenski RE (1993) Evolutionary adaptation to temperature. II. Thermal niches of experimental lines of Escherichia coli. Evolution 47: 1–12.

Bennett M, Monk N (2008) The flowering of systems approaches in plant and crop biology. New Phytol 179: 567-8.

Bennett WN, Boraas ME (1988) Isolation of a fast-growing strain of the rotifer Brachionus calyciflorus Pallas using turbidostat culture. Aquaculture 73: 27–36.

Bennet WN, Boraas ME (1989) A demographic profile of the fastest growing metazoan – a strain of Brachionus calyciflorus (Rotifera). Oikos 55: 365–9.

Benoit J (1936) Role of the thyroid in the gonado-stimulation by artificial light in the domestic duck. CR Soc Biol Paris 123: 243–6.

Bensaad K, Tsuruta A, Selak MA, Vidal MN, Nakano K, et al. (2006) TIGAR, a p53-inducible regulator of glycolysis and apoptosis. Cell 126: 107–20.

Bensaad K, Vousden KH (2007) p53: new roles in metabolism. Trends Cell Biol 17: 286–91.

Bentley GE, Van't Hof TJ, Ball GF (1999) Seasonal neuroplasticity in the songbird telencephalon: a role for melatonin. Proc Natl Acad Sci USA 96: 4674–4679.

Bentley GE, Ball GF (2000) Photoperiod-dependent and -independent regulation of melatonin receptors in the forebrain of songbirds. J Neuroendocrinol 12: 745–52.

Bentley GE, Perfito N, Ukena K, Tsutsui K, Wingfield JC (2003) Gonadotropin-inhibitory peptide in song sparrows (Melospiza melodia) in different reproductive conditions, and in house sparrows (Passer domesticus) relative to chicken-gonadotropin-releasing hormone. J Neuroendocrinol 15: 794–802.

Bentley SD, Vernikos GS, Snyder LA, Churcher C, Arrowsmith C, et al. (2007) Meningococcal genetic variation mechanisms viewed through comparative analysis of serogroup C strain FAM18. PLoS Genet 3: e23.

Benton MJ (2009) The Red Queen and the Court Jester: species diversity and the role of biotic and abiotic factors through time. Science 323: 728-32.

Bentwich I, Avniel A, Karov Y, Aharonov R, Gilad S, et al. (2005) Identification of hundreds of conserved and nonconserved human microRNAs. Nat Genet 37: 766–70.

Berenbaum MR (1996) Introduction to the symposium: on the evolution of specialization. Am Nat 148(suppl.): S78–S83.

Bérénos C, Wegner KM, Schmid-Hempel P (2011) Antagonistic coevolution with parasites maintains host genetic diversity: an experimental test. Proc Biol Sci 278: 218-24.

Bérénos C, Schmid-Hempel P, Wegner KM (2012) Antagonistic coevolution accelerates the evolution of reproductive isolation in tribolium castaneum. Am Nat 180: 520-8.

Berezikov E, Guryev V, van de Belt J, Wienholds E, Plasterk RH, Cuppen E (2005) Phylogenetic shadowing and computational identification of human microRNA genes. Cell 120: 21–4.

Berg H (2003) Factors influencing seed: ovule ratios and reproductive success in four cleistogamous species: a comparison between two flower types. Plant Biology 5: 194-202.

Berg JM (1992) Sp1 and the subfamily of zinc finger proteins with guanine-rich binding sites. Proc Natl Acad Sci USA 89: 11109–10.

Berg L, Pálsson SM, Lascoux M (2001) Fitness and sexual response to population density in Daphnia pulex. Freshwater Biol 46: 667–77.

Bergelson J, Purrington CB (1996) Surveying patterns in the cost of resistance in plants. Am Nat 148: 536–58.

Bergenius MAJ, Meekan MG, Robertson DR, McCormick MI (2002) Larval growth predicts the recruitment success of a coral reef fish. Oecologia 131: 521–5.

Berger SL (2007) The complex language of chromatin regulation during transcription. Nature 447: 407-12.

Bergerat A, de Massy B, Gadelle D, Varoutas PC, Nicolas A, Forterre P (1997) An atypical topoisomerase II from archaea with implication for meiotic recombination. Nature 386: 414–7.

Bergeron L, Perez GI, Macdonald G, Shi L, Sun Y, et al. (1998) Defects in regulation of apoptosis in caspase-2-deficient mice. Genes Dev 12: 1304–14.

Bergeron P, Careau V, Humphries MM, Réale D, Speakman JR, Garant D (2011) The energetic and oxidative costs of reproduction in a free-ranging rodent. Funct Ecol 25: 1063–71.

Berggren MI, Husbeck B, Samulitis B, Baker AF, Gallegos A, Powis G (2001) Thioredoxin peroxidase-1 (peroxiredoxin-1) is increased in thioredoxin-1 transfected cells and results in enhanced protection against apoptosis caused by hydrogen peroxide but not by other agents including dexamethasone, etoposide, and doxorubicin. Arch Biochem Biophys 392: 103–9.

Bergh A (1981) Morphological signs of a direct effect of experimental cryptorchidism on the Sertoli cells in rats irradiated as fetuses. Biol Reprod 24: 145–52.

Bergh A, Lissbrant E, Collin O (1999) Temporal variations in testicular microcirculation. J Androl 20: 724–30.

Bergh A, Collin O, Lissbrant E (2001) Effects of acute graded reductions in testicular blood flow on testicular morphology in the adult rat. Biol Reprod 64: 13-20.

Berglund A, Rosenqvist G (1986) Reproductive costs in the prawn Palaemon adspersus: effects on growth and predator vulnerability. Oikos 46: 349–54.

Berglund J, Pollard KS, Webster MT (2009) Hotspots of biased nucleotide substitutions in human genes. PLoS Biology 7: e26.

Bergman A, Siegal ML (2003) Evolutionary capacitance as a general feature of complex gene networks. Nature 424:549-52.

Bergstrom CT, Pritchard J (1998) Germline bottlenecks and the evolutionary maintenance of mitochondrial genomes. Genetics 149: 2135-46.

Bergstrom CT, McElhany P, Real LA (1999) Transmission bottlenecks as determinants of virulence in rapidly evolving pathogens. Proc Natl Acad Sci USA 96: 5095-100.

Berisha B, Schams D, Kosmann M, Amselgruber W, Einspanier R (2000) Expression and localisation of vascular endothelial growth factor and basic fibroblast growth factor during the final growth of bovine ovarian follicles. J Endocrinol 167: 371-82.

Berkner LV, Marshall LC (1965) On the origin and rise of oxygen concentration in the Earth’s atmosphere. J Atmos Sci 22: 225–61.

Berlin S, Tomaras D, Charlesworth B (2007) Low mitochondrial variability in birds may indicate Hill-Robertson effects on the W chromosome. Heredity 99: 389-96.

Berman J, Hadany L (2012) Does stress induce (para)sex? Implications for Candida albicans evolution. Trends Genet 28: 197-203.

Bernabucci U, Lacetera N, Baumgard LH, Rhoads RP, Ronchi B, Nardone A (2010) Metabolic and hormonal acclimation to heat stress in domesticated ruminants. Animal 4: 1167-83.

Bernal AJ, Dolinoy DC, Huang D, Skaar DA, Weinhouse C, Jirtle RL (2013)Adaptive radiation-induced epigenetic alterations mitigated by antioxidants.FASEB J 27: 665-71.

Bernard D, Quatannens B, Begue A, Vandenbunder B, Abbadie C (2001) Antiproliferative and antiapoptotic effects of crel may occur within the same cells via the up-regulation of manganese superoxide dismutase. Cancer Res 61: 2656–64.

Bernays EA (2001) Neural limitations in phytophagous insects: implications for diet breadth and evolution of host affiliation. Annu Rev Entomol 46: 703–27.

Bernays EA, Graham M (1988) On the evolution of host specificity in phytophagous arthropods. Ecology 69: 886–92.

Berndtson WE (1977)Methods for quantifying mammalian spermatogenesis: a review. J Anim Sci 44:818-33.

Berndtson WE (2011)The importance and validity of technical assumptions required for quantifying sperm production rates: a review.J Androl 32: 2-14.

Berns A (2010) Cancer: The blind spot of p53. Nature 468: 519-20.

Bernstein C (1987) Damage in DNA of an infecting phage T4 shifts reproduction from asexual to sexual allowing rescue of its genes. Genet Res 49: 183–9.

Bernstein C, Johns V (1989) Sexual reproduction as a response to H2O2 damage in Schizosaccharomyces pombe. J Bacteriol 171: 1893-7.

Bernstein C, Bernstein H (1991) Aging, sex, and DNA repair. San Diego, CA: Academic Press.

Bernstein H, Byers GS, Michod RE (1981) Evolution of sexual reproduction: importance of DNA repair, complementation, and variation. Am Nat 117: 537–549.

Bernstein H, Byerly HC, Hopf FA, Michod RE (1984) Origin of sex. J Theor Biol 110: 323–51.

Bernstein H, Byerly HC, Hopf FA, Michod RE (1985a) Genetic damage, mutation, and the evolution of sex. Science 229: 1277-81.

Bernstein H, Byerly HC, Hopf FA, Michod RE (1985b) The evolutionary role of recombinational repair and sex. Int Rev Cytol 96: 1-28.

Bernstein H, Hopf FA, Michod RE (1989) The evolution of sex: DNA repair hypothesis. In: Rasa A, Vogel C, Voland E, eds. Sociobiology of Reproduction, Strategies in Animal and Man. Bechenham, UK: Croom Helm Ltd. pp 3-18.

Beroukhim R, Mermel CH, Porter D, Wei G, Raychaudhuri S, et al. (2010) The landscape of somatic copy-number alteration across human cancers. Nature 463: 899–905.

Berruti G, Perego L, Borgonovo B, Martegani E (1998) MSJ-1, a new member of the DnaJ family of proteins, is a male germ cell-specific gene product. Exp Cell Res 239: 430-41.

Berry DR (1982) The biology of yeasts. London, UK: Edward Arnold.

Bersaglieri T, Sabeti PC, Patterson N, Vanderploeg T, Schaffner SF, et al.(2004) Genetic signatures of strong recent positive selection at the lactase gene. Am J Hum Genet 74:1111–20.

Bertin RI (1982) Paternity and fruit production in the trumpet creeper (Campsis radicans). AmNat 119: 694–709.

Bertin RI, Peters PJ (1992) Paternal effects on offspring quality in Campsis radicans. Am Nat 140: 166-78.

Bertoncini CR, Meneghini R (1995) DNA strand breaks produced by oxidative stress in mammalian cells exhibit 3’-phosphoglycolate termini. Nucleic Acids Res 23: 2995-3002.

Bertrand S, Alonso-Alvarez C, Devevey G, Faivre B, Prost J, Sorci G (2006) Carotenoids modulate the trade-off between egg production and resistance to oxidative stress in zebra finches. Oecologia 147: 576–84.

Besseau L, Faliex E (1994) Resorption of unemitted gametes in Lithognathus mormyrus (Sparidae, Teleostei): a possible synergic action of somatic and immune cells. Cell Tissue Res 276: 123–32.

Besset V, Magueresse-Battistoni B, Collette J, Benahmed M (1996) Tumor necrosis factor α stimulates insulin-like growth factor binding protein 3 expression in cultured porcine Sertoli cells. Endocrinology 137: 296–303.

Bestor TH (2003) Cytosine methylation mediates sexual conflict. Trends Genet 19: 185–90.

Bestor TH, Ingram VM (1983) Two DNA methyltransferases from murine erythroleukemia cells: purification, sequence specificity, and mode of interaction with DNA. Proc Natl Acad Sci USA 80: 5559–63.

Bestor T, Laudano A, Mattaliano R, Ingram V (1988) Cloning and sequencing of a cDNA encoding DNA methyltransferase of mouse cells. The carboxyl-terminal domain of the mammalian enzymes is related to bacterial restriction methyltransferases. J Mol Biol 203: 971–83.

Bestor TH, Bourc'his D (2004) Transposon silencing and imprint establishment in mammalian germ cells. Cold Spring Harb Symp Quant Biol 69: 381-7.

Betancourt A (2007) When the going gets tough, beneficial mutations get going. Heredity 99: 359–360.

Betancourt AJ (2009) Genomewide patterns of substitution in adaptively evolving populations of the RNA bacteriophage MS2. Genetics 181: 1535–44.

Betancourt AJ, Presgraves DC (2002) Linkage limits the power of natural selection in Drosophila. Proc Natl Acad SciUSA 99: 13616–20.

Betancourt M, Fereres A, Fraile A, Garcia-Arenal F (2008) Estimation of the effective number of founders that initiate an infection after aphid transmission of a multipartite plant virus. J Virol 82: 12416–21.

Betel D, Sheridan R, Marks DS, Sander C (2007) Computational analysis of mouse piRNA sequence and biogenesis. PLoS Comput Biol 3: e222.

Betrán E, Thornton K, Long M (2002) Retroposed new genes out of the X in Drosophila. Genome Res 12: 1854–9.

Beukeboom LW (2007) Sporadic sexual reproduction: Sex to some degree. Heredity 98: 123–4.

Beukeboom LW, Vrijenhoek RC (1998) Evolutionary genetics and ecology of sperm-dependent parthenogenesis. J Evol Biol 11:755–82.

Beukeboom LW, Zwaan BJ (2005) Genetics. In: Jervis MA, ed. Insect natural enemies. Dordrecht, The Netherlands: Springer. pp 167–218.

Beumer TL, Roepers-Gajadien HL, Gademan IS, van Buul PP, Gil-Gomez G, et al. (1998)The role of the tumor suppressor p53 in spermatogenesis.Cell Death Differ 5:669-77.

Beutler B (1995) TNF, immunity and inflammatory disease: lessons of the past decade. J Investig Med 43: 227-35.

Beutler B, Cerami A (1986) Cachectin and tumour necrosis factor as two sides of the same biological coin. Nature 320: 584-8.

Beutler B, Cerami A (1987) Cachectin: more than a tumor necrosis factor. N Engl J Med 316: 379-85.

Bevilacqua L, Ramsey JJ, Hagopian K, Weindruch R, Harper ME (2004) Effects of short- and medium-term calorie restriction on muscle mitochondrial proton leak and reactive oxygen species production. Am J Physiol Endocrinol Metab 286:E852-61.

Beyer S, Kristensen MM, Jensen KS, Johansen JV, Staller P (2008) The histone demethylases JMJD1A and JMJD2B are transcriptional targets of hypoxia-inducible factor HIF. J Biol Chem 283: 36542–52.

Beyer U, Moll-Rocek J, Moll UM, Dobbelstein M (2011)Endogenous retrovirus drives hitherto unknown proapoptotic p63 isoforms in the male germ line of humans and great apes.Proc Natl Acad Sci USA 108: 3624-9.

Bhakat KK, Mantha AK, Mitra S (2009) Transcriptional regulatory functions of mammalian AP-endonuclease (APE1/Ref-1), an essential multifunctional protein. Antioxid Redox Signal 11: 621-38.

Bhattacharyya NP, Skandalis A, Ganesh A, Groden J, Meuth M (1994) Mutator phenotypes in human colorectal carcinoma cell lines. Proc Natl Acad Sci USA 91: 6319-23.

Bhusari SS, Dobosy JR, Fu V, Almassi N, Oberley T, Jarrard DF (2010)Superoxide dismutase 1 knockdown induces oxidative stress and DNA methylation loss in the prostate.Epigenetics 5: 402-9.

Bichara M, Meier M, Wagner J, Cordonnier A, Lambert IB (2011) Postreplication repair mechanisms in the presence of DNA adducts in Escherichia coli. Mutat Res 727: 104–22.

Biebricher CK, Eigen M (2005) The error threshold. Virus Res 107: 117–27.

Biek R, Rodrigo AG, Holley D, Drummond A, Anderson CR Jr, et al. (2003)Epidemiology, genetic diversity, and evolution of endemic feline immunodeficiency virus in a population of wild cougars.J Virol 77:9578-89.

Biel SW, Hartl DL (1983) Evolution of transposons: natural selection for Tn5 in Escherichia coli K12. Genetics 103:581-92.

Bielas JH, Loeb KR, Rubin BP, True LD, Loeb LA (2006) Human cancers express a mutator phenotype. Proc Natl Acad Sci USA 103: 18238–42.

Biemont C, Vieira C (2006) Genetics: Junk DNA as an evolutionary force. Nature 443: 521–4.

Bienert GP, Schjoerring JK, Jahn TP (2006) Membrane transport of hydrogen peroxide. Biochim Biophys Acta 1758: 994–1003.

Bienert GP, Møller AL, Kristiansen KA, Schulz A, Møller IM, et al. (2007) Specific aquaporins facilitate the diffusion of hydrogen peroxide across membranes. J Biol Chem 282: 1183-92.

Bierne N, Eyre-Walker A (2004) The genomic rate of adaptive amino acid substitution in Drosophila. Mol Biol Evol 21: 1350–60.

Bierzychudek P (1985) Patterns in plant parthenogenesis. Experimentia 41: 1255-64.

Bierzychudek P (1987) Pattern in plant parthenogenesis. In: Stearns SC, ed. The Evolution of Sex and Its Consequences. Boston, MA: Birkhaeuser. pp 197–218.

Bijlsma R, Loeschcke V (2005) Environmental stress, adaptation and evolution: an overview. J Evol Biol 18: 744-9.

Bijlsma R, Loeschcke V (2012) Genetic erosion impedes adaptive responses to stressful environments. Evol Appl 5: 117-29.

Billard R (1969) La spermatogenèse de Poecilia reticulata. I–Estimation du nombre degénérations goniales et rendement de la spermatogenèse. Ann Biol Anim Biochim Biophys 9:251–71.

Billard R (1986) Spermatogenesis and spermatology of some teleost fish species. Reprod Nutr Dev 26: 877–920.

Billard R, Takashima F (1983) Resorption of spermatozoa in the sperm duct of rainbow trout during the post-spawning period. Bull Japan Soc Sci Fish 49: 387–92.

Billi AC, Alessi AF, Khivansara V, Han T, Freeberg M, et al. (2012) The Caenorhabditis elegans HEN1 ortholog, HENN-1, methylates and stabilizes select subclasses of germline small RNAs. PLoS Genet 8: e1002617.

Billig H, Furuta I, Rivier C, Tapanainen J, Parvinen M, Hsueh AJW (1995) Apoptosis in testis germ cells: developmental changes in gonadotropin dependence and localization to selective tubule stages. Endocrinology 136: 5–12.

Billingham ME (1987) Cytokines as inflammatory mediators. Br Med Bull 43: 350-70.

Bindra RS, Schaffer PJ, Meng A, Woo J, Maseide K, et al. (2004) Downregulation of Rad51 and decreased homologous recombination in hypoxic cancer cells. Mol Cell Biol 24: 8504–18.

Bindra RS, Schaffer PJ, Meng A, Woo J, Måseide K, et al. (2005a) Alterations in DNA repair gene expression under hypoxia: elucidating the mechanisms of hypoxia-induced genetic instability. Ann NY Acad Sci 1059: 184-95.

Bindra RS, Gibson SL, Meng A, Westermark U, Jasin M, et al. (2005b) Hypoxia-induced down-regulation of BRCA1 expression by E2Fs. Cancer Res 65: 11597–604.

Bindra RS, Crosby ME, Glazer PM (2007) Regulation of DNA repair in hypoxic cancer cells. Cancer Metastasis Rev 26: 249–60.

Bindra RS, Glazer PM (2007a) Repression of RAD51 gene expression by E2F4/p130 complexes in hypoxia. Oncogene 26: 2048-57.

Bindra RS, Glazer PM (2007b) Co-repression of mismatch repair gene expression by hypoxia in cancer cells: role of the Myc/Max network. Cancer Lett 252: 93–103.

Bingaman EW, Magnuson DJ, Gray TS, Handa RJ (1994) Androgen inhibits the increases in hypothalamic corticotropin-releasing hormone (CRH) and CRH-immunoreactivity following gonadectomy. Neuroendocrinology 59: 228-34.

Bingham PM, Kidwell MG, Rubin GM (1982) The molecular basis of P-M hybrid dysgenesis - the role of the P element, a P-strain-specific transposon family. Cell 29: 995–1004.

Binelli M, Murphy BD (2010) Coordinated regulation of follicle development by germ and somatic cells. Reprod Fertil Dev 22: 1–12.

Birchler JA, Yao H, Chudalayandi S, Vaiman D, Veitia RA (2010) Heterosis. Plant Cell 22: 2105-12.

Bird AP (1980) DNA methylation and the frequency of CpG in animal DNA. Nucleic Acids Res 8: 1499-504.

Bird A (2002) DNA methylation patterns and epigenetic memory. Genes Dev 16: 6–21.

Bird AP, Wolffe AP (1999) Methylation-induced repression--belts, braces, and chromatin. Cell 99: 451-4.

Bird CP, Stranger BE, Liu M, Thomas DJ, Ingle CE, et al. (2007) Fast-evolving noncoding sequences in the human genome. Genome Biol 8: R118.

Bird JM, Hodkinson ID (2005) What limits the altitudinal distribution of Craspedolepta species (Sternorrhyncha: Psylloidea) on fireweed? Ecol Entomol 30: 510–20.

Birdsell JA (2002) Integrating genomics, bioinformatics, and classical genetics to study the effects of recombination on genome evolution. Mol Biol Evol 19: 1181–97.

Birdsell J, Wills C (1996) Significant competitive advantage conferred by meiosis and syngamy in the yeast Saccharomyces cerevisiae. Proc Natl Acad Sci USA 93: 908-12.

Birdsell JA, Wills C (2003) The evolutionary origin and maintenance of sexual recombination: A review of contemporary models. In: MacIntyre RJ, Clegg MT, eds. Evolutionary BiologyVol. 33. New York, NY: Kluwer Academic Publishers. pp 27–138.

Birkhead TR (2000) Promiscuity: An Evolutionary History of Sperm Competition and Sexual Conflict. London, UK: Faber & Faber.

Birkhead TR (2010) How stupid not to have thought of that: Post copulatory sexual selection. J Zool (Lond) 281: 78–93.

Birkhead TR, Briskie JV, Møller AP (1993a) Male sperm reserves and copulation frequency in birds. Behav Ecol Sociobiol 32: 85–93.

Birkhead TR, Møller AP, Sutherland WJ (1993b) Why do females make it so difficult for males to fertilize their eggs? J Theor Biol 161: 51–60.

Birkhead TR, Møller AP (1998) Sperm competition and sexual selection. London, UK: Academic Press.

Birkhead TR, Martinez JG, Burke T, Froman DP (1999) Sperm mobility determines the outcome of sperm competition in the domestic fowl. Proc R Soc B-Biol Sci 266: 1759–64.

Birkhead TR, Pizzari T (2002) Postcopulatory sexual selection. Nat Rev Gene 3: 262–73.

Birkhead TR, Brillard JP (2007) Reproductive isolation in birds: postcopulatory prezygotic barriers. Trends Ecol Evol 22: 266-72.

Birky CW Jr (1995) Uniparental inheritance of mitochondrial and chloroplast genes: mechanisms and evolution. Proc Natl Acad Sci USA 92: 11331–8.

Birky CW Jr (1996) Heterozygosity, heteromorphy, and phylogenetic trees in asexual eukaryotes. Genetics 144: 427–37.

Birky CW (2004) Bdelloid rotifers revisited. Proc Natl Acad Sci USA 101: 2651-2.

Birky CW Jr (2009) Sex and evolution in eukaryotes. [Internet]. Oxford, UK: Encyclopedia of Life Support Systems (EOLSS) Publishers. Available from: http://www.eolss.net.

Birky CW Jr (2010) Positively negative evidence for asexuality. J Hered 101: S42–S45.

Birky CW Jr, Walsh JB (1988) Effects of linkage on rates of molecular evolution. Proc Natl Acad Sci USA 85: 6414–8.

Birky CW Jr, Wolf C, Maughan H, Herbertson L, Henry E (2005) Speciation and selection without sex. Hydrobiologia 546: 29–45.

Birky CW, Jr., Adams J, Gemmel M, Perry J (2010) Using population genetic theory and DNA sequences for species detection and identification in asexual organisms. PLoS ONE 5: e10609.

Birmingham A, Anderson EM, Reynolds A, Ilsley-Tyree D, Leake D, et al. (2006) 3' UTR seed matches, but not overall identity, are associated with RNAi off-targets. Nat Methods 3: 199-204.

Birney E, Stamatoyannopoulos JA,Dutta A, Guigó R, Gingeras TR, et al. (2007) Identification and analysis of functional elements in 1% of the human genome by the ENCODE pilot project. Nature 447: 799–816.

Biswas G, Guha M, Avadhani NG (2005) Mitochondria-to-nucleus stress signaling in mammalian cells: nature of nuclear gene targets, transcription regulation, and induced resistance to apoptosis. Gene 354: 132-9.

Bixler A, Tang-Martinez Z (2006) Reproductive performance as a function of inbreeding in prairie voles (Microtus ochrogaster). J Mammal 87: 944–9.

Bizé P, Devevey G, Monaghan P, Doligez B, Christe P (2008) Fecundity and survival in relation to resistance to oxidative stress in a free-living bird. Ecology 89: 2584–93.

Bizuayehu TT, Babiak J, Norberg B, Fernandes JM, Johansen SD, Babiak I (2012) Sex-biased miRNA expression in Atlantic halibut (Hippoglossus hippoglossus) brain and gonads. Sex Dev 6: 257-66.

Bjedov I, Tenaillon O, Gerard B, Souza V, Denamur E, et al. (2003) Stress-induced mutagenesis in bacteria. Science 300: 1404–9.

Bjelakovic G, Beninati S, Pavlovic D, Kocic G, Jevtovic T, et al. (2007) Glucocorticoids and oxidative stress. J Basic Clin Physiol Pharmacol 18: 115-27.

Björk JK, Sistonen L (2010) Regulation of the members of the mammalian heat shock factor family. FEBS J 277: 4126–39.

Björk JK, Sandqvist A, Elsing AN, Kotaja N, Sistonen L (2010) miR-18, a member of Oncomir-1, targets heat shock transcription factor 2 in spermatogenesis. Development 137: 3177–84.

Björkholm B, Sjölund M, Falk PG, Berg OG, Engstrand L, Andersson DI (2001) Mutation frequency and biological cost of antibiotic resistance in Helicobacter pylori. Proc Natl Acad Sci USA 98: 14607–12.

Bjørnstad ON, Fromentin JM, Stenseth NC, Gjøsaeter J (1999) Cycles and trends in cod populations. Proc Natl Acad Sci USA 96: 5066–71.

Black JC, Van Rechem C, Whetstine JR (2012) Histone lysine methylation dynamics: establishment, regulation, and biological impact. Mol Cell 48: 491–507.

Blackburn EH (2000) Telomere states and cell fates. Nature 408: 53–6.

Blackman RL (1971) Variation in the photoperiodic response within natural populations of Myzus persicae (Sulz). Bull Entomol Res 60: 533–46.

Blackman RL, Spence JM (1996) Ribosomal DNA is frequently concentrated on only one X chromosome in permanently apomitic aphids, but this does not inhibit male determination. Chromosome Res 109: 314–20.

Blackman RL, Eastop VF (1994) Aphids on the world's trees, an identification and information guide. Wallingford, UK: CAB International.

Blackman RL, Eastop VF (2000) Aphids on the world’s crops: An identification and information guide (2nd edition). Chichester, UK: John Wiley and Sons Ltd.

Blackman RL, Eastop VF (2006) Aphids on the world’s herbaceous plants and shrubs. Chichester, UK: John Wiley and Sons Ltd.

Blackstone NW (1999) Redox control in development and evolution: evidence from colonial hydroids. J Exp Biol 202: 3541–53.

Blackstone NW (2006)Multicellular redox regulation: integrating organismal biology and redox chemistry.Bioessays 28: 72-7.

Blackstone N (2009)Mitochondria and the redox control of development in cnidarians.Semin Cell Dev Biol 20: 330-6.

Blackstone NW, Green DR (1999) The evolution of a mechanism of cell suicide. Bioessays 21: 84-8.

Blagosklonny MV (2011) Cell cycle arrest is not senescence. Aging (Albany NY) 3: 94-101.

Blagosklonny MV, An WG, Romanova LY, Trepel J, Fojo T, Neckers L (1998)p53 inhibits hypoxia-inducible factor-stimulated transcription.J Biol Chem 273: 11995-8.

Blair AC, Wolfe LM (2004) The evolution of an invasive plant: an experimental study with Silene latifolia. Ecology 85: 3035–42.

Blais JD, Addison CL, Edge R, Falls T, Zhao H, et al. (2006) Perk-dependent translational regulation promotes tumor cell adaptation and angiogenesis in response to hypoxic stress. Mol Cell Biol 26: 9517–32.

Blakely WF, Fuciarelli AF, Wegher BJ, Dizdaroglu M (1990)Hydrogen peroxide-induced base damage in deoxyribonucleic acid.Radiat Res 121: 338-43.

Blanc G, Ngwamidiba M, Ogata H, Fournier PE, Claverie JM, Raoult D (2005) Molecular evolution of Rickettsia surface antigens: evidence of positive selection. Mol Biol Evol 22: 2073–83.

Blanckenhorn WU (1998) Adaptive phenotypic plasticity in growth, development, and body size in the yellow dung fly. Evolution 52: 1394-407.

Blanckenhorn WU, Hosken DJ, Martin OY, Reim C, Teuschl Y, Ward PI (2002) The costs of copulating in the dung fly Sepsis cynipsea. Behav Ecol 13: 353–8.

Blanckenhorn WU, Heyland A (2004) The quantitative genetics of two life history tradeoffs in the yellow dung fly in abundant and limited food environments. EvolEcol 18: 385-402.

Blanco D, Vicent S, Fraga MF, Fernandez-Garcia I, Freire J, et al. (2007) Molecular analysis of a multistep lung cancer model induced by chronic inflammation reveals epigenetic regulation of p16 and activation of the DNA damage response pathway. Neoplasia 9: 840–52.

Blanco M, Herrera G, Urios A (1995) Increased mutability by oxidative stress in OxyR-deficient Escherichia coli and Salmonella typhimurium cells: clonal occurrence of the mutants during growth on nonselective media. Mutat Res Lett 346: 215-20.

Blanco R, Gerhardt H (2013)VEGF and Notch in tip and stalk cell selection.Cold Spring Harb Perspect Med 3: a006569.

Blanco-Rodríguez J (1998) A matter of death and life: the significance of germ cell death during spermatogenesis. Int J Androl 21: 236-48.

Blanco-Rodríguez J, Martinez-Garcia C, Porras A (2003) Correlation between DNA synthesis in the second, third and fourth generations of spermatogonia and the occurrence of apoptosis in both spermatogonia and spermatocytes. Reproduction 126: 661–8.

Blasco MA, Serrano M, Fernandez-Capetillo O (2011)Genomic instability in iPS: time for a break.EMBO J 30: 991-3.

Blaylock BG, Shugart HH Jr (1972) The effect of radiation-induced mutations on the fitness of Drosophila populations. Genetics 72: 469-74.

Blázquez J (2003) Hypermutation as a factor contributing to the acquisition of antimicrobial resistance. Clin Infect Dis 37: 1201–9.

Bleiweiss R (1998) Slow rate of molecular evolution in high-elevation hummingbirds. Proc Natl Acad Sci USA 95: 612–6.

Bleuyard J-Y, Gallego ME, White CI (2006) Recent advances in understanding of the DNA double-strand break repair machinery of plants. DNA Repair 5:1–12.

Blewitt ME, Vickaryous NK, Paldi A, Koseki H, Whitelaw E (2006) Dynamic reprogramming of DNA methylation at an epigenetically sensitive allele in mice. PLoS Genet 2: 399-405.

Blish CA (2012) HIV-1 transmission goes retro (steps back).J Infect Dis 206: 1336-8.

Bloch GJ, Gorski RA (1988) Estrogen/progesterone treatment in adulthood affects the size of several components of the medial preoptic area in the male rat. J Comp Neurol 275: 613-22.

Bloch Qazi MC, Heifetz Y, Wolfner MF (2003) The developments between gametogenesis and fertilization: ovulation and female sperm storage in Drosophila melanogaster. Dev Biol 256: 195–211.

Block E (1952) Quantitative morphological investigations of the follicular system in women. Variations at different ages. Acta Anat 14: 108–23.

Blok RB, Gook DA, Thorburn DR, Dahl HH (1997) Skewed segregation of the mtDNA nt 8993 (TRG) mutation in human oocytes. Am J Hum Genet 60: 1495–501.

Blondin P, Coenen K, Sirard MA (1997) The impact of reactive oxygen species on bovine sperm fertilizing ability and oocyte maturation. J Androl 18: 454–60.

Bloom JD, Labthavikul ST, Otey CR, Arnold FH (2006) Protein stability promotes evolvability. Proc Natl Acad Sci USA 103:5869–74.

Bloom JD, Arnold FH (2009)In the light of directed evolution: pathways of adaptive protein evolution.Proc Natl Acad Sci USA 106 Suppl 1: 9995-10000.

Blottner S, Hingst O, Meyer HH (1995) Inverse relationship between testicular proliferation and apoptosis in mammalian seasonal breeders. Theriogenology 44:321–8.

Blouin MS (2000) Neutrality tests on mtDNA: unusual results from nematodes. J Hered 91: 156-8.

Blueweiss L, Fox H, Kudzma V, Nakashima D, Peters R, Sams S (1978) Relationship between body size and some life history parameters. Oecologia 37: 257-72.

Blumenstiel JP (2007)Sperm competition can drive a male-biased mutation rate.J Theor Biol 249: 624-32.

Bock WJ (2003) Ecological aspects of the evolutionary processes.Zool Sci 20: 279-89.

Bock WJ (2010) Multiple explanations in Darwinian evolutionary theory. Acta Biotheor 58: 65-79.

Bode AM, Dong Z (2005)Inducible covalent posttranslational modification of histone H3.Sci STKE 2005: re4.

Bode SN, Adolfsson S, Lamatsch DK, Martins MJ, Schmit O, et al. (2010) Exceptional cryptic diversity and multiple origins of parthenogenesis in a freshwater ostracod. Mol Phylogenet Evol 54: 542–52.

Bodey B (2002)Cancer-testis antigens: promising targets for antigen directed antineoplastic immunotherapy.Expert Opin Biol Ther 2:577-84.

Boe L, Danielsen M, Knudsen S, Petersen JB, Maymann J, Jensen PR (2000) The frequency of mutators in populations of Escherichia coli. Mutat Res 448: 47–55.

Boekelheide K, Lee J, Shipp EB, Richburg JH, Li G (1998) Expression of Fas system-related genes in the testis during development and after toxicant exposure. Toxicol Lett 102–103: 503–8.

Bogenhagen DF (1999) Repair of mtDNA in vertebrates. Am J Hum Genet 64: 1276–81.

Boggs C (1992) Resource allocation: exploring connections between foraging and life history. FunctEcol 6:508–18.

Bohler HCL, Zoeller RT, King JC, Rubin BS, Weber R, Merriam GR (1990) Corticotropin-releasing hormone mRNA is elevated on the afternoon of proestrous in the parvocellular paraventricular nuclei of the female rat. Mol Brain Res 8: 259-62.

Böhm J, Munk-Schulenburg S, Felscher S, Kohlhase J (2006)SALL1 mutations in sporadic Townes-Brocks syndrome are of predominantly paternal origin without obvious paternal age effect.Am J Med Genet A 140:1904-8.

Böhne A, Brunet F, Galiana-Arnoux D, Schultheis C, Volff JN (2008)Transposable elements as drivers of genomic and biological diversity in vertebrates.Chromosome Res 16: 203-15.

Bohr VA, Smith CA, Okumoto DS, Hanawalt PC (1985) DNA repair in an active gene: removal of pyrimidine dimers from the DHFR gene of CHO cells is much more efficient than in the genome overall. Cell 40: 359–69.

Bohr VA, Stevnsner T, de Souza-Pinto NC (2002) Mitochondrial DNA repair of oxidative damage in mammalian cells. Gene 286: 127–34.

Boissinot S, Davis J, Entezam A, Petrov D, Furano AV (2006) Fitness cost of LINE-1 (L1) activity in humans. Proc Natl Acad Sci USA 103: 9590–4.

Boissonneault G (2002) Chromatin remodeling during spermiogenesis: a possible role for the transition proteins in DNA strand break repair. FEBS Lett 514: 111–4.

Boiteux S, Radicella JP (2000) The human OGG1 gene: structure, functions, and its implication in the process of carcinogenesis. Arch Biochem Biophys 377: 1–8.

Boland NI, Humpherson PG, Leese HJ, Gosden RG (1993) Pattern of lactate production and steroidogenesis during growth and maturation of mouse ovarian follicles in vitro. Biol Reprod 48: 798-806.

Boldajipour B, Raz E (2007) What is left behind—quality control in germ cell migration. Sci STKE 2007: pe16.

Boldt DH (1999) New perspectives on iron: an introduction. Am J Med Sci 318: 207-12.

Boldyrev AA, Carpenter DO, Huentelman MJ, Peters CM, Johnson P (1999)Sources of reactive oxygen species production in excitotoxin- stimulated cerebellar granule cells.Biochem Biophys Res Commun 256: 320-4.

Bolis PF, Soro V, Martinetti Bianchi M, Belvedere M (1985) HLA compatibility and human reproduction. Clin Exp Obstet Gynecol 12: 9–12.

Bollback, J. P., and J. P. Huelsenbeck, 2007 Clonal interference is alleviated by high mutation rates in large populations. Mol. Biol. Evol. 24(6): 1397–1406.

Bolnick DI (2001)Intraspecific competition favours niche width expansion in Drosophila melanogaster.Nature 410: 463-6.

Bolnick DI (2004) Can intraspecific competition drive disruptive selection? An experimental test in natural populations of sticklebacks. Evolution 58: 608–18.

Bolnick DI, Preisser EL (2005) Resource competition modifies the strength of trait-mediated predator-prey interactions: a meta-analysis. Ecology 86: 2771–9.

Bonaiti-Pellie C, Pelet A, Ogier H, Nelson JR, Largilliere C, et al.(1990) A probable sex difference in mutation rates in ornithine transcarbamylase deficiency. Hum Genet 84: 163-6.

Bonavida B, Granger G, eds. (1990) Tumor Necrosis Factor: Structure, Mechanism of Action, Role in Disease and Therapy. Basel, Switzerland: Karger.

Bondar T, Medzhitov R (2010)p53-mediated hematopoietic stem and progenitor cell competition.Cell Stem Cell 6:309-22.

Bonduriansky R (2009) Reappraising sexual coevolution and the sex roles. PLoS Biol 7: e1000255.

Bonduriansky R, Day T (2009) Nongenetic inheritance and its evolutionary implications. Annu Rev Ecol Evol Syst 40: 103–25.

Bonduriansky R, Crean AJ, Day T (2012) The implications of nongenetic inheritance for evolution in changing environments. Evol Appl 5: 192-201.

Bonello S, Zähringer C, BelAiba RS, Djordjevic T, Hess J, et al. (2007) Reactive oxygen species activate the HIF-1α promoter via a functional NFκB site. Arterioscler Thromb Vasc Biol 27: 755–61.

Bonenfant C, Gaillard J, Coulson T, Festa-Bianchet M, Loison A, Garel M, Loe LE, et al. (2009) Empirical evidence of density-dependence in populations of large herbivores. In: Caswell H, ed. Advances in ecological research. Vol. 41. New York, NY: Academic Press.pp 313–357.

Bonhoeffer S, Chappey C, Parkin NT, Whitcomb JM, Petropoulos CJ (2004) Evidence for positive epistasis in HIV-1. Science 306: 1547–50.

Bonifield HR, Hughes KT (2003) Flagellar phase variation in Salmonella enterica is mediated by a posttranscriptional control mechanism. J Bacteriol 185: 3567-74.

Bonizzi G, Piette J, Merville MP, Bours V (2000)Cell type-specific role for reactive oxygen species in nuclear factor-kappaB activation by interleukin-1.Biochem Pharmacol 59: 7-11.

Bonneau KR, Mullens BA, MacLachlan NJ (2001) Occurrence of genetic drift and founder effect during quasispecies evolution of the VP2 and NS3/NS3A genes of bluetongue virus upon passage between sheep, cattle, and Culicoides sonorensis. J Virol 75: 8298–305.

Bonner JT (2003) On the origin of differentiation. J Biosci 28: 523–8.

Bonner JT (2009) The social amoebae: The biology of cellular slime molds. Princeton, NJ: Princeton University Press.

Bonner-Weir S (2000)Life and death of the pancreatic beta cells.Trends Endocrinol Metab 11: 375-8.

Bookman SS (1984) Evidence for selective fruit production in Asclepias. Evolution 38: 72–86.

Boominathan L (2010) The tumor suppressors p53, p63, and p73 are regulators of microRNA processing complex. PLoS One 5: e10615.

Boonstra R, McColl CJ, Karels TJ (2001) Reproduction at all costs: the adaptive stress response of male arctic ground squirrels. Ecology 82: 1930-46.

Boonyaprakob U, Gadsby JE, Hedgpeth V, Routh PA, Almond GW (2005) Expression and localization of hypoxia inducible factor-1 alpha mRNA in the porcine ovary. Can J Vet Res 69: 215–22.

Booth IR (2002)Stress and the single cell: intrapopulation diversity is a mechanism to ensure survival upon exposure to stress.Int J Food Microbiol 78:19-30.

Boots M, Sasaki A (2000) The evolutionary dynamics of local infection and global reproduction in host-parasite interactions. Ecol Lett 3: 181–5.

Booy D, Hendriks RJJ, Smulders MJM, Van Groenendael JM, Vosman B (2000) Genetic diversity and the survival of the populations. Plant Biol 2: 379–95.

Boraas ME (1983) Population dynamics of food-limited rotifers in two-stage chemostat culture. Limnol Oceanogr 28: 546–63.

Borde V, Robine N, Lin W, Bonfils S, Géli V, Nicolas A (2009) Histone H3 lysine 4 trimethylation marks meiotic recombination initiation sites. EMBO J 28: 99–111.

Borenstein E, Meilijson I, Ruppin E (2006)The effect of phenotypic plasticity on evolution in multipeaked fitness landscapes.J Evol Biol 19: 1555-70.

Borger DR, Essig DA (1998) Induction of HSP 32 gene in hypoxic cardiomyocytes is attenuated by treatment with N-acetyl-L-cysteine. Am J Physiol Heart Circ Physiol 274: H965–H973.

Borgia G (1979) Sexual selection and the evolution of mating systems. In: Blum MS, Blum NA, eds. Sexual selection and reproductive systems in insects. New York, NY: Academic Press. pp 19-80.

Borkovich KA, Farrelly FW, Finkelstein DB, Taulien J, Lindquist S (1989) hsp82 is an essential protein that is required in higher concentrations for growth of cells at higher temperatures. Mol Cell Biol 9: 3919–30.

Bornstein SR, Rutkowski H, Vrezas I (2004)Cytokines and steroidogenesis.Mol Cell Endocrinol 215: 135-41.

Borrello ME (2005)The rise, fall and resurrection of group selection.Endeavour 29: 43-7.

Borrello ME (2012) Evolutionary Restraints: The Contentious History of Group Selection. Chicago, IL: University of Chicago Press.

Borstnik B, Pumpernik D (2002) Tandem repeats in protein coding regions of primate genes. Genome Res 12: 909–15.

Borum K (1961) Oogenesis in the mouse: a study of the meiotic prophase. Exp Cell Res 24: 495–507.

Boschetto C, Gasparini C, Pilastro A (2011) Sperm number and velocity affect sperm competition success in the guppy (Poecilia reticulata). Behav Ecol Sociobiol65: 813-21.

Boss O, Hagen T, Lowell BB (2000)Uncoupling proteins 2 and 3: potential regulators of mitochondrial energy metabolism.Diabetes 49: 143-56.

Bossdorf O, Richards CL, Pigliucci M (2008) Epigenetics for ecologists. Ecol Lett 11: 106–15.

Boswell RE, Mahowald AP (1985) Tudor, a gene required for assembly of the germ plasm in Drosophila melanogaster. Cell 43: 97–104.

Bott RC, McFee RM, Clopton DT, Toombs C, Cupp AS (2006) Vascular endothelial growth factor and kinase domain region receptor are involved in both seminiferous cord formation and vascular development during testis morphogenesis in the rat. Biol Reprod 75: 56–67.

Boveris A, Chance B (1973) The mitochondrial generation of hydrogen peroxide. General properties and effect of hyperbaric oxygen. Biochem J 134: 707–16.

Boucot AJ (1990) Evolutionary Paleobiology of Behavior and Coevolution. Amsterdam, The Netherlands: Elsevier.

Bouguenec V, Giani N (1989) Biological studies upon Enchytraeus variatus in breeding cultures. Hydrobiologia 180: 151–65.

Boul KE, Funk WC, Darst CR, Cannatella DC, Ryan MJ (2007) Sexual selection drives speciation in an Amazonian frog. Proc Biol Sci 274: 399-406.

Bourc’his D, Bestor TH (2004) Meiotic catastrophe and retrotransposon reactivation in male germ cells lacking Dnmt3L. Nature 431: 96-9.

Bourguet D, Guillemaud T, Chevillon C, Raymond M (2004) Fitness costs of insecticide resistance in natural breeding sites of the mosquito Culex pipiens. Evolution 58: 128–35.

Bourque G, Pevzner PA, Tesler G (2004) Reconstructing the genomic architecture of ancestral mammals: lessons from human, mouse, and rat genomes. Genome Res 14: 507–16.

Bourque G (2009) Transposable elements in gene regulation and in the evolution of vertebrate genomes. Curr Opin Genet Dev 19: 607-12.

Boussouar F, Grataroli R, Ji J, Benahmed M (1999) Tumor necrosis factor-α stimulates lactate dehydrogenase A expression in porcine cultured Sertoli cells: mechanisms of action. Endocrinology 140: 3054–62.

Boussouar F, Benahmed M (2004) Lactate and energy metabolism in male germ cells. Trends Endocrinol Metab 15: 345–50.

Bouvet GF, Jacobi V, Plourde KV, Bernier L (2008)Stress-induced mobility of OPHIO1 and OPHIO2, DNA transposons of the Dutch elm disease fungi.Fungal Genet Biol 45: 565-78.

Bouwman KM, Van Dijk RE, Wijmenga JJ, Komdeur J (2007) Older male reed buntings are more successful at gaining extrapair fertilizations. Anim Behav 73: 15–27.

Boveris A (1977) Mitochondrial production of superoxide radical and hydrogen peroxide. Adv Exp Med Biol 78: 1-82.

Bowater RP, Jaworski A, Larson JE, Parniewski P, Wells RD (1997) Transcription increases the deletion frequency of long CTG.CAG triplet repeats from plasmids in Escherichia coli. Nucleic Acids Res 25: 2861-8.

Bowen NJ, Jordan IK (2002)Transposable elements and the evolution of eukaryotic complexity.Curr Issues Mol Biol 4: 65-76.

Bowen NJ, Jordan IK (2007)Exaptation of protein coding sequences from transposable elements.Genome Dyn 3: 147-62.

Bowles D (2004)A radical idea: men and women are different.Cardiovasc Res 61: 5-6.

Boyce MS (1984) Restitution of r- and K-selection as a model of density-dependent natural selection. Annu Rev Ecol Syst 15: 427–447.

Boyce MS, Perrins CM (1987) Optimizing Great Tit clutch size in a fluctuating environment. Ecology 68: 142-53.

Boyd JM, Drevland RM, Downs DM, Graham DE (2009) Archaeal ApbC/Nbp35 homologs function as iron-sulfur cluster carrier proteins. J Bacteriol 191: 1490–7.

Boyer JC, Umar A, Risinger JI, Lipford JR, Kane M, et al. (1995) Microsatellite instability, mismatch repair deficiency, and genetic defects in human cancer cell lines. Cancer Res 55: 6063-70.

Boyko A, Kovalchuk I (2008) Epigenetic control of plant stress response. Environ Mol Mutagen 49: 61–72.

Boyko A, Blevins T, Yao Y, Golubov A, Bilichak A, et al. (2010) Transgenerational adaptation of arabidopsis to stress requires DNA methylation and the function of Dicer-like proteins. PLoS ONE 5: e9514.

Boyko A, Kovalchuk I (2011) Genome instability and epigenetic modification—heritable responses to environmental stress? Curr Opin Plant Biol 14: 260-6.

Bracht JR, Fang W, Goldman AD, Dolzhenko E, Stein EM, Landweber LF (2013)Genomes on the edge: programmed genome instability in ciliates.Cell 152: 406-16.

Brackett BG (2004) Male reproduction in mammals. In: Reece WO, ed. Dukes' Physiology of Domestic Animals, 12th edn. Ithaca, NY: Cornell University Press. pp 670–691.

Brackney DE, Brown IK, Nofchissey RA, Fitzpatrick KA, Ebel GD (2010)Homogeneity of Powassan virus populations in naturally infected Ixodes scapularis.Virology 402: 366-71.

Bradbury JW, Vehrencamp SL, Gibson R (1985) Leks and the unanimity of female choice. In: Greenwood PJ, Harvey PH, Slatkin M, eds. Evolution: Essays in Honour of John Maynard Smith. Cambridge, UK: Cambridge University Press. pp 301-314.

Bradshaw AD (1965) Evolutionary significance of phenotypic plasticity in plants. Adv Genet 13: 115–55.

Bradshaw AD, McNeilly T (1991) Evolutionary response to global climatic change. Ann Bot 67: 5–14.

Bradshaw WE, Holzapfel CM (2007) Evolution of animal photoperiodism. Annu Rev Ecol Evol Syst 38: 1-25.

Bradshaw VA, McEntee K (1989) DNA damage activates transcription and transposition of yeast Ty retrotransposons. Mol Gen Genet 218: 465–74.

Brady N, Gaymes TJ, Cheung M, Mufti GJ, Rassool FV (2003) Increased error-prone NHEJ activity in myeloid leukemias is associated with DNA damage at sites that recruit key nonhomologous end-joining proteins. Cancer Res 63: 1798–805.

Braendle C, Flatt T (2006) A role for genetic accommodation in evolution?Bioessays 28:868-73.

Branch P, Aquilina G, Bignami M, Karran P (1993) Defective mismatch binding and a mutator phenotype in cells tolerant to DNA damage. Nature 362: 652-4.

Branch P, Hampson R, Karran P (1995) DNA mismatch binding defects, DNA damage tolerance, and mutator phenotypes in human colorectal carcinoma cell lines. Cancer Res 55: 2304-9.

Branciforte D, Martin SL (1994) Developmental and cell type specificity of LINE-1 expression in mouse testis: implications for transposition. Mol Cell Biol 14: 2584–92.

Brand MD, Couture P, Hulbert AJ (1994) Liposomes from mammalian liver mitochondria are more polyunsaturated and leakier to protons than those from reptiles. Comp Biochem Physiol 108B: 181–8.

Brand MD, Affourtit C, Esteves TC, Green K, Lambert AJ, et al. (2004) Mitochondrial superoxide: production, biological effects, and activation of uncoupling proteins.Free Radic Biol Med 37: 755-67.

Brand MD, Esteves TC (2005) Physiological functions of the mitochondrial uncoupling proteins UCP2 and UCP3. Cell Metab 2: 85–93.

Brandeis M, Kafri T, Ariel M, Chaillet JR, McCarrey J, Razin A, Cedar H (1993) The ontogeny of allele-specific methylation associated with imprinted genes in the mouse. EMBO J 12: 3669–77.

Brandes RP, Kreuzer J (2005) Vascular NADPH oxidases: molecular mechanisms of activation. Cardiovasc Res 65: 16–27.

Brandriff B, Pedersen RA (1981) Repair of the ultraviolet-irradiated male genome in fertilized mouse eggs. Science 211: 1431–3.

Brandsma I, Gent DC (2012)Pathway choice in DNA double strand break repair: observations of a balancing act.Genome Integr 3: 9.

Brasier MD, Lindsay JF (2001) Did supercontinental amalgamation trigger the “Cambrian Explosion”? In: Zhuralev AY, Riding R, eds. The Ecology of the Cambrian Radiation. New York, NY: Columbia University Press. pp 69–89.

Braun RE (1998) Every sperm is sacred--or is it? Nat Genet 18: 202–4.

Bravard A, Vacher M, Gouget B, Coutant A, de Boisferon FH, et al.(2006) Redox regulation of human OGG1 activity in response to cellular oxidative stress. Mol Cell Biol 26: 7430-6.

Bravard A, Vacher M, Moritz E, Vaslin L, Hall J, et al. (2009)Oxidation status of human OGG1-S326C polymorphic variant determines cellular DNA repair capacity.Cancer Res 69: 3642-9.

Brawer J, Schipper H, Robaire B (1983)Effects of long term androgen and estradiol exposure on the hypothalamus.Endocrinology 112: 194-9.

Brawer JR, Beaudet A, Desjardins GC, Schipper HM (1993) Pathological effect of estradiol on the hypothalamus. Biol Reprod 49: 647-52.

Braydich-Stolle L, Kostereva N, Dym M, Hofmann MC (2007)Role of Src family kinases and N-Myc in spermatogonial stem cell proliferation.Dev Biol 304: 34-45.

Breen MS, Kemena C, Vlasov PK, Notredame C, Kondrashov FA (2012)Epistasis as the primary factor in molecular evolution.Nature 490: 535-8.

Breivik J (2001) Don’t stop for repairs in a war zone: Darwinian evolution unites genes and environment in cancer development. Proc Natl Acad Sci USA 98: 5379–81.

Breivik J, Gaudernack G (1999) Genomic instability, DNA methylation, and natural selection in colorectal carcinogenesis. Semin Cancer Biol 9: 245–54.

Brekken RA, Puolakkainen P, Graves DC, Workman G, Lubkin SR, Sage EH (2003) Enhanced growth of tumors in SPARC null mice is associated with changes in the ECM. J Clin Invest 111: 487–95.

Brengues M, Teixeira D, Parker R (2005) Movement of eukaryotic mRNAs between polysomes and cytoplasmic processing bodies. Science 310: 486–9.

Brennan LJ, Keddie BA, Braig HR, Harris HL (2008)The endosymbiont Wolbachia pipientis induces the expression of host antioxidant proteins in an Aedes albopictus cell line.PLoS One 3: e2083.

Brennan LJ, Haukedal JA, Earle JC, Keddie B, Harris HL (2012)Disruption of redox homeostasis leads to oxidative DNA damage in spermatocytes of Wolbachia-infected Drosophila simulans.Insect Mol Biol 21: 510-20.

Brennan RJ, Schiestl RH (1998) Free radicals generated in yeast by the Salmonella test-negative carcinogens benzene, urethane, thiourea and auramine O. Mutat Res 403:65–73.

Brennan ST, Lowenstein TK, Horita J (2004) Seawater chemistry and the advent of biocalcification. Geology 32: 473–6.

Brennecke J, Aravin AA, Stark A, Dus M, Kellis M, Sachidanandam R, Hannon GJ (2007) Discrete small RNA-generating loci as master regulators of transposon activity in Drosophila. Cell 128: 1089–103.

Brennecke J,Malone CD, Aravin AA, Sachidanandam R, Stark A, Hannon GJ (2008) An epigenetic role for maternally inherited piRNAs in transposon silencing. Science 322: 1387–92.

Brenner CA, Wolny YM, Barritt JA, Matt DW, Munne S, Cohen J (1998) Mitochondrial DNA deletion in human oocytes and embryos. Mol Hum Reprod 4: 887-92.

Bréque C, Surai P, Brillard JP (2003)Roles of antioxidants on prolonged storage of avian spermatozoa in vivo and in vitro.Mol Reprod Dev 66: 314-23.

Breuner CW, Patterson SH, Hahn TP (2008) In search of relationships between the acute adrenocortical response and fitness. Gen Comp Endocrinol 157: 288–95.

Brevini TA, Vassena R, Francisci C, Gandolfi F (2005) Role of adenosine triphosphate, active mitochondria, and microtubules in the acquisition of developmental competence of parthenogenetically activated pig oocytes. Biol Reprod 72: 1218–23.

Brick K, Smagulova F, Khil P, Camerini-Otero RD, Petukhova GV (2012)Genetic recombination is directed away from functional genomic elements in mice.Nature 485: 642-5.

Bridges CB (1919) Specific modifiers of eosin eye color in Drosophila melanogaster. J Exp Zool 28: 337-84.

Brigelius-Flohe R (2006) Glutathione peroxidases and redox-regulated transcription factors. Biol Chem 387: 1329–35.

Briggs D, Miller D, Gosden RG (1999)Molecular biology of female gametogenesis. In:Fauser B,

Rutherford A, Strauss J, Van Steirteghem A, ed. Molecular Biology in Reproductive Medicine.New York, NY: Parthenon. pp 251–269.

Brinkworth MH, Weinbauer GF, Schlatt S, Nieschlag E (1995) Identification of male germ cells undergoing apoptosis in adult rats. J Reprod Fertil 105: 25–33.

Brinkworth MH, Weinbauer GF, Bergmann M, Nieschlag E (1997) Apoptosis as a mechanism of germ cell loss in elderly men. Int J Androl 20: 222–8.

Briones V, Muegge K (2012) The ghosts in the machine: DNA methylation and the mystery of differentiation. Biochim Biophys Acta 1819: 757–62.

Briskie JV, Montgomerie R (1992) Sperm size and sperm competition in birds. Proc R Soc Lond B 247: 89–95.

Briskie JV, Montgomerie R (1993) Patterns of sperm storage in relation to sperm competition in passerine birds. Condor 95: 442–54.

Briskie JV, Montgomerie R, Birkhead TR (1997) The evolution of sperm size in birds. Evolution 51: 937–45.

Brison DR, Houghton FD, Falconer D, Roberts SA, Hawkhead J, et al. (2004) Identification of viable embryos in IVF by non-invasive measurement of amino acid turnover. Hum Reprod 19: 2319–2324.

Brisson D (2003)The directed mutation controversy in an evolutionary context. Crit Rev Microbiol 29: 25-35.

Bristow RG, Hill RP (2008) Hypoxia and metabolism. Hypoxia, DNA repair and genetic instability. Nat Rev Cancer 8:180-92.

Brock GJ, Anderson NH, Monckton DG (1999) Cis-acting modifiers of expanded CAG/CTG triplet repeat expandability: association with flanking GC content and proximity to CpG islands. Hum Mol Genet 8: 1061-7.

Brock MA, Nielsen DN, Shiel RJ, Green JD, Langley JD (2003) Drought and aquatic community resilience: the role of eggs and seeds in sediments of temporary wetlands. Freshwater Biol 48: 1207–18.

Brock RD (1971) The role of induced mutations in plant improvement. Radiat Bot 11: 181-96.

Brockes JP, Kumar A (2005)Appendage regeneration in adult vertebrates and implications for regenerative medicine.Science 310: 1919-23.

Brockhurst MA, Morgan AD, Rainey PB, Buckling A (2003) Population mixing accelerates coevolution. Ecol Lett 6: 975–9.

Brodie EI, Brodie EJ (1999) Predator–prey arms races. Bioscience 49: 557–68.

Broekmans FJ, Soules MR, Fauser BC (2009) Ovarian aging: mechanisms and clinical consequences. Endocr Rev 30: 465-93.

Brokelmann J (1963) Fine structure of germ cells and Sertoli cells during the cycle of the seminiferous epithelium in the rat. Z Zellforsch Mikrosk Anat 59: 820–50.

Bromage N, Porter M, Randall C (2001) The environmental regulation of maturation in farmed finfish with special reference to the role of photoperiod and melatonin. Aquaculture 197: 63–98.

Brönmark C, Miner JG (1992) Predator-induced phenotypical change in body morphology in crucian carp. Science 258: 1348–50.

Brook BW, Bradshaw CJA (2006) Strength of evidence for density dependence in abundance time series of 1198 species. Ecology 87: 1445–51.

Brook JD, McCurrach ME, Harley HG, Buckler AJ, Church D, et al. (1992) Molecular basis of myotonic dystrophy: expansion of a trinucleotide (CTG) repeat at the 3' end of a transcript encoding a protein kinase family member. Cell 68: 799-808.

Brookes PS, Buckingham JA, Tenreiro AM, Hulbert AJ, Brand MD (1998)The proton permeability of the inner membrane of liver mitochondria from ectothermic and endothermic vertebrates and from obese rats: correlations with standard metabolic rate and phospholipid fatty acid composition.Comp Biochem Physiol B Biochem Mol Biol 119: 325-34.

Brookes P, Darley-Usmar VM (2002) Hypothesis: the mitochondrial NO signaling pathway, and the transduction of nitrosative to oxidative cell signals: an alternative function for cytochrome C oxidase. Free Radic Biol Med 32: 370–4.

Brookes PS, Yoon Y, Robotham JL, Anders MW, Sheu SS (2004)Calcium, ATP, and ROS: a mitochondrial love-hate triangle.Am J Physiol Cell Physiol 287: C817-33.

Brookfield JFY (2004)Evolutionary genetics: Mobile DNAs as sources of adaptive change?Curr Biol 14: R344-5.

Brookfield JFY (2005) The ecology of the genome—Mobile DNA elements and their hosts. Nat Rev Genet 6: 128–36.

Brosius J (1999) RNAs from all categories generate retrosequences that may be exapted as novel genes or regulatory elements. Gene 238: 115–34.

Brosius J (2003) The contribution of RNAs and retroposition to evolutionary novelties. Genetica 118: 99–116.

Brosius J (2005) Waste not, want not–transcript excess in multicellular eukaryotes. Trends Genet 21: 287–8.

Brown GC (2001)Regulation of mitochondrial respiration by nitric oxide inhibition of cytochrome c oxidase.Biochim Biophys Acta 1504: 46-57.

Brown GC, Cooper CE (1994) Nanomolar concentrations of nitric oxide reversibly inhibit synaptosomal respiration by competing with oxygen at cytochrome oxidase. FEBS Lett 356: 295–8.

Brown JA, Roberts TL, Richards R, Woods R, Birrell G, et al. (2011) A novel role for hSMG-1 in stress granule formation. Mol Cell Biol 31: 4417–29.

Brown JH (1981) Two decades of homage to Santa Rosalia: toward a general theory of diversity. Am Zool 21: 877–88.

Brown JH (1984) On the relationship between abundance and distribution of species. Am Nat 124: 255-79.

Brown JH (2001) Mammals on mountainsides: elevational patterns of diversity. Global Ecol Biogeogr10:101–9.

Brown JH, Marquet PA, Taper ML (1993) Evolution of body size: consequences of an energetic definition of fitness. Am Nat 142: 573–584.

Brown JH, Lomolino MV (1998) Biogeography. 2nd edn. Sunderland, MA: Sinauer.

Brown JH, Gillooly JF, Allen AP, Savage VM, West GB (2004) Toward a metabolic theory of ecology. Ecology 81: 1771–89.

Brown JL (1997) A theory of mate choice based on heterozygosity. Behav Ecol 8: 60–5.

Brown JS, Vincent TL (1992) Organisation of predator-prey communities as an evolutionary game. Evolution 46: 1269–83.

Brown MJF, Loosli R, Schmid-Hempel P (2000) Condition-dependent expression of virulence in a trypanosome infecting bumblebees. Oikos 91: 421-7.

Brown SG, Kwan S, Shero S (1995) The parasitic theory of sexual reproduction: parasitism in unisexual and bisexual geckos. Proc R Soc Lond B 260: 317–20.

Brown TR, Greene FE, Bardin CW (1976)Androgen receptor dependent and independent activities of testosterone on hepatic microsomal drug metabolism.Endocrinology 99: 1353-62.

Brown WM, George M, Wilson AC (1979) Rapid evolution of animal mitochondrial DNA. Proc Natl Acad Sci USA 76: 1967-71.

Brown WM, Prager EM, Wang A, Wilson AC (1982) Mitochondrial DNA sequences of primates: tempo and mode of evolution. J Mol Evol 18: 225–39.

Browne RA, Hoopes CW (1990) Genotype diversity and selection in asexual brine shrimp (Artemia). Evolution 44: 1035–51.

Broyde S, Wang L, Rechkoblit O, Geacintov NE, Patel DJ (2008)Lesion processing: high-fidelity versus lesion-bypass DNA polymerases. Trends Biochem Sci 33: 209-19.

Bruggeman J, Debets AJ, Wijngaarden PJ, de Visser JA, Hoekstra RF (2003) Sex slows down the accumulation of deleterious mutations in the homothallic fungus Aspergillus nidulans. Genetics 164: 479–85.

Bruggeman J, Debets AJ, Hoekstra RF (2004) Selection arena in Aspergillus nidulans. Fungal Genet Biol 41:181–8.

Bruggeman V, Van As P, Decuypere E (2002) Developmental endocrinology of the reproductive axis in the chicken embryo. Comp Biochem Physiol A 131: 839–46.

Brumer Y, Michor F, Shakhnovich EI (2006) Genetic instability and the quasispecies model. J Theor Biol 241: 216–22.

Brune B, Zhou J (2003) The role of nitric oxide (NO) in stability regulation of hypoxia inducible factor-1 alpha (HIF-1alpha). Curr Med Chem 10: 845–55.

Brunelle JK, Bell EL, Quesada NM, Vercauteren K, Tiranti V, et al. (2005)Oxygen sensing requires mitochondrial ROS but not oxidative phosphorylation.Cell Metab 1: 409-14.

Brunet E, Rouzine IM, Wilke CO (2008)The stochastic edge in adaptive evolution.Genetics 179:603-20.

Brunner KA, Klauninger B (2003) An integrative image of causality and emergence. In: Arshinov V, Fuchs C, eds. Causality, Emergence, Self-Organisation. Moscow, Russia: NIA-Priroda. pp 23-35.

Bruno JF, Edmunds PJ (1997) Clonal variation for phenotypic plasticity in the coral Madracis mirabilis. Ecology 78: 2177-90.

Bruskov VI, Malakhova LV, Masalimov ZK, Chernikov AV (2002) Heat-induced formation of reactive oxygen species and 8-oxoguanine, a biomarker of damage to DNA. Nucleic Acids Res 30: 1354–63.

Bruvo R, Schulenburg H, Storhas M, Michiels NK (2007) Synergism between mutational meltdown and Red Queen in parthenogenetic biotypes of the freshwater planarian Schmidtea polychroa. Oikos 116: 313–23.

Bryant EH, Meffert LM (1986) The effect of serial founder-flush cycles on quantitative genetic variation in the housefly. Heredity 70: 122-9.

Bryant EH, Reed DH (1999) Fitness decline under relaxed selection in captive populations. Conserv Biol 7: 122–31.

Buard J, Barthès P, Grey C, de Massy B (2009) Distinct histone modifications define initiation and repair of meiotic recombination in the mouse. EMBO J 28: 2616–24.

Bubici C, Papa S, Dean K, Franzoso G (2006)Mutual cross-talk between reactive oxygen species and nuclear factor-kappa B: molecular basis and biological significance.Oncogene 25: 6731-48.

Bubliy OA, Loeschcke V(2002) Effects of low stressful temperature on genetic variation of five quantitative traits in Drosophila melanogaster. Heredity 89: 70-5.

Buccione R, Schroeder AC, Eppig JJ (1990) Interactions between somatic cells and germ cells throughout mammalian oogenesis. Biol Reprod 43: 543–7.

Buch JP, Lamb DJ, Lipshultz LI, Smith RG (1988) Partial characterization of a unique growth factor secreted by human Sertoli cells. Fertil Steril 49: 658–65.

Buchan JR, Parker R (2009)Eukaryotic stress granules: the ins and outs of translation.Mol Cell 36: 932-41.

Buchanan A, Triant M, Bedau M (2004) The flexible balance of evolutionary novelty and memory in the face of environmental catastrophes. Artificial Life IX. MIT Press, Cambridge, MA, 297-302.

Buchanan KL (2000)Stress and the evolution of condition-dependent signals.Trends Ecol Evol 15: 156-60.

Buchberger A, Bukau B, Sommer T (2010)Protein quality control in the cytosol and the endoplasmic reticulum: brothers in arms.Mol Cell 40: 238-52.

Bucher E, Reinders J, Mirouze M (2012)Epigenetic control of transposon transcription and mobility in Arabidopsis. Curr Opin Plant Biol 15: 503-10.

Bucheton A, Paro R, Sang HM, Pelisson A, Finnegan DJ (1984) The molecular basis of I-R hybrid dysgenesis in Drosophila melanogaster - identification, cloning, and properties of the I-Factor. Cell 38: 153–63.

Buchholz JT (1922) Developmental selection in vascular plants. Bot Gaz 73: 249-86.

Buchner H (1987) Untersuchungen über die Bedingungen der heterogonen Fortpflanzungsarten bei den Rädertieren III. Über den Verlust der miktischen Potenz bei Brachionus urceolaris. Arch Hydrobiol 109: 333–54.

Buchner J (1999)Hsp90 & Co. - a holding for folding.Trends Biochem Sci 24: 136-41.

Buckling A, Rainey PB (2002a) The role of parasites in sympatric and allopatric host diversification. Nature 420: 496–9.

Buckling A, Rainey PB (2002b) Antagonistic coevolution between a bacterium and a bacteriophage. Proc R Soc Lond B 269: 931–6.

Buckling A, Wei Y, Massey RC, Brockhurst MA, Hochberg ME (2006) Antagonistic coevolution with parasites increases the cost of host deleterious mutations. Proc RSoc B Biol Sci 273: 45–9.

Budanov AV, Sablina AA, Feinstein E, Koonin EV, Chumakov PM (2004) Regeneration of peroxiredoxins by p53-regulated sestrins, homologs of bacterial AhpD. Science 304: 596–600.

Budanov AV (2011)Stress-responsive sestrins link p53 with redox regulation and mammalian target of rapamycin signaling.Antioxid Redox Signal 15: 1679-90.

Budde LM, Wu C, Tilman C, Douglas I, Ghosh S (2002)Regulation of IkappaBbeta expression in testis.Mol Biol Cell 13: 4179-94.

Budnik LT, Jahner D, Mukhopadhyay AK (1999) Inhibitory effects of TNFα on mouse tumor Leydig cells: possible role of ceramide in the mechanism of action. Mol Cell Endocrinol 150:39–46.

Bueno J, Pfaff DW (1976) Single unit recording in hypothalamus and preoptic area of estrogen-treated and untreated ovariectomized female rats. Brain Res 101: 67-78.

Buettner GR (1993)The pecking order of free radicals and antioxidants: lipid peroxidation, alpha-tocopherol, and ascorbate.Arch Biochem Biophys 300: 535-43.

Buettner GR, Ng CF, Wang M, Rodgers VG, Schafer FQ (2006) A new paradigm: manganese superoxide dismutase influences the production of H2O2 in cells and thereby their biological state. Free Radic Biol Med 41: 1338–50.

Buggisch M, Ateghang B, Ruhe C, Strobel C, Lange S, et al. (2007) Stimulation of ES-cell-derived cardiomyogenesis and neonatal cardiac cell proliferation by reactive oxygen species and NADPH oxidase. J Cell Sci 120: 885–94.

Bühler M (2009) RNA turnover and chromatin-dependent gene silencing. Chromosoma 118:141-51.

Bui LT, Abbass HA, Branke J (2005) Multiobjective optimization for dynamic environments.In: Proceedings of the IEEE Congress on evolutionary computation, vol 3. pp 2349–2356.

Bukau B, Weissman J, Horwich A (2006)Molecular chaperones and protein quality control.Cell 125: 443-51.

Bukovsky A, Caudle MR, Svetlikova M, Wimalasena J, Ayala ME, Dominguez R (2005) Oogenesis in adult mammals, including humans: a review. Endocrine 26: 301-16.

Bull JJ, Vogt RC (1979) Temperature-dependent sex determination in turtles. Science 206: 1186–8.

Bull JJ, Meyers LA, Lachmann M (2005) Quasispecies made simple. PLoS Comput Biol 1: e61.

Bull L (1999) On the Baldwin effect. Artif Life 5: 241–6.

Bull RA, Luciani F, McElroy K, Gaudieri S, Pham ST, et al. (2011) Sequential bottlenecks drive viral evolution in early acute hepatitis C virus infection. PLoS Pathog 7: e1002243.

Buller HA, Kothe MJC, Goldman DA, Grubman SA, Sasak WV, et al. (1990) Coordinate expression of lactase-phlorizin hydrolase mRNA and enzyme levels in rat intestine during development. J Biol Chem 265: 6978–83.

Bulletti C, Flamigni C, Giacomucci E (1996) Reproductive failure due to spontaneous abortion and recurrent miscarriage. Hum Reprod Update 2: 118-36.

Bullock JM, Edwards RJ, Carey PD, Rose RJ 2000) Geographical separation of two Ulex species at three spatial scales: Does competition limit species’ ranges? Ecography 23: 257–71.

Bulmer MG (1980) The Mathematical Theory of Quantitative Genetics. Oxford, UK: Oxford University Press.

Bulmer MG (1985) Selection for iteroparity in a variable environment. Am Nat 126: 63–71.

Bulmer M (1986) Neighboring base effects on substitution rates in pseudogenes. Mol Biol Evol 3: 322–9.

Bunn HF, Poyton RO (1996) Oxygen sensing and molecular adaptation to hypoxia. Physiol Rev 76: 839–85.

Burch CL, Chao L (2000) Evolvability of an RNA virus is determined by its mutational neighbourhood. Nature 406: 625-8.

Burch CL, Chao L (2004) Epistasis and its relationship to canalization in the RNA virus phi 6. Genetics 167: 559–67.

Burch LH, Yang Y, Sterling JF, Roberts SA, Chao FG, et al. (2011) Damage-induced localized hypermutability. Cell Cycle 10: 1073-85.

Burda H (2003) Adaptations for subterranean life. In: Grzimek´s Animal life encyclopaedia, Vol 12. New York, NY: Gale Inc. pp 69–78.

Burdge GC, Slater-Jefferies J, Torrens C, Phillips ES, Hanson MA, Lillycrop KA (2007) Dietary protein restriction of pregnant rats in the F0 generation induces altered methylation of hepatic gene promoters in the adult male offspring in the F1 and F2 generations. Br J Nutr 97: 435-9.

Burdon JJ, Thrall PH (2003) The fitness costs to plants of resistance to pathogens. Genome Biol 4: 227.

Burdon RH (1986) Heat shock and the heat shock proteins. Biochem J 240: 313-24.

Burdon RH (1995) Superoxide and hydrogen peroxide in relation to mammalian cell proliferation. Free Radic Biol Med 18: 775–94.

Burdon RH, Rice-Evans C (1989) Free radicals and the regulation of mammalian cell proliferation. Free Radical Res Commun 6: 345–58.

Burger HG, Dudley EC, Hopper JL, Shelley JM, Green A, et al. (1995) The endocrinology of the menopausal transition: a cross-sectional study of a population-based sample. J Clin Endocrinol Metab 80: 3537–45.

Bürger R (1999) Evolution of genetic variability and the advantage of sex and recombination in changing environments. Genetics 153: 1055-69.

Bürger R (2000) The Mathematical Theory of Selection, Recombination, and Mutation. Chichester, UK: Wiley.

Bürger R (2002) Additive genetic variation under intraspecific competition and stabilizing selection: a two-locus study. Theor Popul Biol 61: 197-213.

Bürger R, Gimelfarb A (2002) Fluctuating environments and the role of mutation in maintaining quantitative genetic variation. Genet Res 80: 31-46.

Burgess R, Yang Z (2008) Estimation of hominoid ancestral population sizes under Bayesian coalescent models incorporating mutation rate variation and sequencing errors. Mol Biol Evol 25: 1979–94.

Burgoyne PS, Baker TG (1981) Oocyte depletion in XO mice and their XX sibs from 12 to 200 days post partum. J Reprod Fertil 61: 207-12.

Burgoyne PS, Baker TG (1985) Perinatal oocyte loss in XO mice and its implications for the aetiology of gonadal dysgenesis in XO women. J Reprod Fertil 75: 633–45.

Burhans WC, Heintz NH (2009) The cell cycle is a redox cycle: linking phase-specific targets to cell fate. Free Radic Biol Med 47: 1282-93.

Burke MK, Dunham JP, Shahrestani P, Thornton KR, Rose MR, Long AD (2010) Genome-wide analysis of a longterm evolution experiment with Drosophila. Nature 467: 587–90.

Burley N (1986) Sexual selection for aesthetic traits in species with biparental care. Am Nat 127: 415–45.

Burley N (1988) The differential-allocation hypothesis: an experimental test. Am Nat 132: 611–28.

Burness G, Casselman SJ, Schulte-Hostedde AI, Moyes CD, Montgomerie R (2004) Sperm swimming speed and energetics vary with sperm competition risk in bluegill (Lepomis macrochirus). Behav Ecol Sociobiol 56:65–70.

Burnet FM (1957) A modification of Jerne’s theory of antibody production using the concept of clonal selection. Aust J Sci 20: 67-9.

Burnett AL, Diehl NA (1963) The nervous system of Hydra. III. The initiation of sexuality with special reference to the nervous system. J Exp Zool 157: 237-50.

Burney S, Caulfield JL, Niles JC, Wishnok JS, Tannenbaum SR (1999) The chemistry of DNA damage from nitric oxide and peroxynitrite. Mutat Res 424: 37-49.

Burnicka-Turek O, Shirneshan K, Paprotta I, Grzmil P, Meinhardt A, et al. (2009) Inactivation of insulin-like factor 6 disrupts the progression of spermatogenesis at late meiotic prophase. Endocrinology 150: 4348-57.

Burns JG, Mery F (2010) Transgenerational memory effect of ageing in Drosophila. J Evol Biol 23: 678-86.

Burns JH, Blomberg SP, Crone EE, Ehrlén J, Knight TM, et al. (2010) Empirical tests of life-history evolution theory using phylogenetic analysis of plant demography. J Ecol 98: 334-44.

Burt A, Trivers RL (2006) Genes in conflict: the biology of selfish genetic elements. Cambridge, MA: Belknapp.

Burt A (2000) Perspective: sex, recombination, and the efficacy of selection--was Weismann right? Evolution 54: 337-51.

Burton NO, Burkhart KB, Kennedy S (2011) Nuclear RNAi maintains heritable gene silencing in Caenorhabditis elegans. Proc Natl Acad Sci USA 108: 19683-8.

Burzynski A, Zbawicka M, Skibinski DOF, Wenne R (2003) Evidence for recombination of mtDNA in the marine mussel Mytilus trossulus from the baltic. Mol Biol Evol 20: 388–92.

Busch JW, Neiman M, Koslow JM (2004) Evidence for maintenance of sex by pathogens in plants. Evolution 58: 2584-90.

Bushati N, Cohen SM (2007) microRNA functions. Annu Rev Cell Dev Biol 23: 175-205.

Buss LW (1982) Somatic cell parasitism and the evolution of somatic tissue compatibility. Proc Natl Acad Sci USA 79: 5337-41.

Buss LW (1983) Evolution, development, and the units of selection. Proc Natl Acad Sci USA 80: 1387-91.

Buss LW (1985) The uniqueness of the individual revisited. In: Jackson JBC, Buss LW, Cook RE, eds. Population Biology and Evolution of Clonal Organisms. New Haven, CT: Yale University Press. pp 467-505.

Buss LW (1987) The evolution of individuality. Princeton, NJ: Princeton University Press.

Busso CS, Liu CJ, Hash CT, Witcombe JR, Devos KM, et al. (1995) Analysis of recombination rate in female and male gametogenesis in pearl millet (Pennisetum glaucum) using RFLP markers. Theor Appl Genet 90: 242–6.

Bussolati B, Moggio A, Collino F, Aghemo G, D’Armento G, et al. (2012) Hypoxia modulates the undifferentiated phenotype of human renal inner medullary CD133+ progenitors through Oct4/miR-145 balance. Am J Physiol Renal Physiol 302: F116–F128.

Busson D, Gans M, Komitopoulou K, Masson M (1983) Genetic analysis of three dominant female sterile mutations located on the X chromosome of Drosophila melanogaster. Genetics 105: 309-25.

Bustamante CD, Fledel-Alon A,Williamson S, Nielsen R, Hubisz MT, et al. (2005) Natural selection on protein-coding genes in the human genome. Nature 437: 1153–7.

Bustamante-Marín X, Quiroga C, Lavandero S, Reyes JG, Moreno RD (2012) Apoptosis, necrosis and autophagy are influenced by metabolic energy sources in cultured rat spermatocytes. Apoptosis 17: 539–50.

Bustos-Obregón E, González JR, Espinoza O (2005) Melatonin as protective agent for the cytotoxic effects of diazinon in the spermatogenesis in the earthworm Eisenia foetida. Ital J Anat Embryol 110(2 Suppl 1): 159-65.

Buszczak M, Cooley L (2000) Eggs to die for: cell death during Drosophila oogenesis. Cell Death Differ 7: 1071-4.

Butlin RK (2000) Virgin rotifers. Trends Ecol Evol 15: 389–90.

Butlin (2002) Evolution of sex: The costs and benefits of sex: new insights from old asexual lineages. Nat Rev Genet 3: 311-7.

Butlin RK, Tregenza T (1998) Levels of genetic polymorphism: marker loci versus quantitative traits. Phil Trans R Soc Lond B 353: 187–98.

Butow RA, Avadhani NG (2004) Mitochondrial signaling: the retrograde response. Mol Cell 14: 1–15.

Buttke IM, Sandstrom PA (1994) Oxidative stress as a mediator of apoptosis. Immunol Today 15: 7-10.

Buzzard JJ, Morrison JR, O'Bryan MK, Song Q, Wreford NG (2000) Development expression of thyroid hormone receptors in the rat testis. Biol Reprod 62: 664-9.

Byers JA, Waits L (2006) Good genes sexual selection in nature. Proc Natl Acad Sci USA 103: 16343-5.

Bygren LO, Kaati G, Edvinsson S (2001) Longevity determined by paternal ancestors’ nutrition during their slow growth period. Acta Biotheor 49: 53–59.

Byrne PG, Roberts JD, Simmons LW (2002) Sperm competition selects for increased testes mass in Australian frogs. J Evol Biol 15: 347–55.

Byskov AG (1986) Differentiation of mammalian embryonic gonad. Physiol Rev 66: 71-117.

Caballero A, Toro MA, López-Fanjul C (1991) The response to artificial selection from new mutations in Drosophila melanogaster. Genetics 128: 89-102.

Cabrero J, Palomino-Morales RJ, Camacho JPM (2007) The DNA-repair Ku70 protein is located in the nucleus and tail of elongating spermatids in grasshoppers. Chromosome Res 15: 1093–100.

Cáceres CE (1997) Temporal variation, dormancy, and coexistence: a field test of the storage effect. Proc Natl Acad Sci USA 94: 9171–5.

Cáceres CE, Hairston Jr NG (1998) Benthic-pelagic coupling in planktonic crustaceans: the role of the benthos. In: Brendonck L, De Meester L, Hairston Jr NG, eds. Evolutionary and ecological aspects of crustacean diapause. Arch Hydrobiol Beih Ergebn Limnol 52. pp 163–174.

Cáceres CE, Tessier AJ (2004) Incidence of diapause varies among populations of Daphnia pulicaria. Oecologia 141: 425–31.

Cadenas E, Davies KJ (2000) Mitochondrial free radical generation, oxidative stress, and aging. Free Radic Biol Med 29: 222–30.

Cadet J, Douki T, Gasparutto D, Ravanat JL (2003) Oxidative damage to DNA: formation, measurement and biochemical features. Mutat Res 531: 5–23.

Caires KC, de Avila J, McLean DJ (2009) Vascular endothelial growth factor regulates germ cell survival during establishment of spermatogenesis in the bovine testis. Reproduction 138: 667–77.

Cairns BR (2009) The logic of chromatin architecture and remodelling at promoters. Nature 461: 193–8.

Cairns J (2000) The contribution of bacterial hypermutators to mutation in stationary phase. Genetics 156: 923–6.

Cairns J, Overbaugh J, Miller S (1988) The origin of mutants. Nature 335: 142–5.

Cairns J, Foster PL (1991) Adaptive reversion of a frameshift mutation in Escherichia coli. Genetics 128: 695–701.

Calamita G, Mazzone A, Cho YS, Valenti G, Svelto M (2001a) Expression and localization of the aquaporin-8 water channel in rat testis. Biol Reprod 64: 1660-6.

Calamita G, Mazzone A, Bizzoca A, Svelto M (2001b) Possible involvement of aquaporin-7 and -8 in rat testis development and spermatogenesis. Biochem Biophys Res Commun 288: 619-25.

Calamita G, Ferri D, Gena P, Liquori GE, Cavalier A, et al. (2005) The inner mitochondrial membrane has aquaporin-8 water channels and is highly permeable to water. J Biol Chem 280: 17149–53.

Calamita G, Gena P, Meleleo D, Ferri D, Svelto M (2006) Water permeability of rat liver mitochondria: A biophysical study. Biochim Biophys Acta 1758: 1018-24.

Calamita G, Moreno M, Ferri D, Silvestri E, Roberti P, et al. (2007) Triiodothyronine modulates the expression of aquaporin-8 in rat liver mitochondria. J Endocrinol 192: 111–20.

Calderón Guzmán D, Mejía GB, Vázquez IE, García EH, del Angel DS, Olguín HJ (2005) Effect of testosterone and steroids homologues on indolamines and lipid peroxidation in rat brain. J Steroid Biochem Mol Biol 94: 369–73.

Calero C, Ibáñez O, Mayol M, Rosselló JA (1999) Random amplified polymorphic DNA (RAPD) markers detect a single phenotype in Lysimachia minoricensis J.J.Rodr. (Primulaceae), a wild extinct plant. Mol Ecol 8: 2133–6.

Calhim S, Immler S, Birkhead TR (2007) Postcopulatory sexual selection is associated with reduced variation in sperm morphology. PLoS ONE 2: e413.

Calisi RM, Bentley GE (2009) Lab and field experiments: are they the same animal? Horm Behav 56: 1–10.

Calisi RM, Díaz-Muñoz SL, Wingfield JC, Bentley GE (2011) Social and breeding status are associated with the expression of GnIH. Genes Brain Behav 10: 557–64.

Callaghan A, Holloway GJ (1999) The relationship between environmental stress and variance. Ecol Appl 9: 456–62.

Callard GV, Jorgensen JC, Redding JM (1995) Biochemical analysis of programmed cell death during premeiotic stages of spermatogenesis in vivo and in vitro. Dev Genet 16: 140–7.

Callard GV, McClusky LM, Betka M (1998) Apoptosis as a normal mechanism of growth control and target of toxicant actions during spermatogenesis. In: Halvorson LGA, ed. New Developments in Marine Biotechnology. New York, NY: Plenum Press. pp 125–128.

Calo S, Billmyre RB, Heitman J (2013) Generators of phenotypic diversity in the evolution of pathogenic microorganisms. PLoS Pathog 9: e1003181.

Calow P (1979) The cost of reproduction – A physiological approach. Biol Rev 54: 23-40.

Calsbeek R, Gosden TP, Kuchta SR, Svensson EI (2012) Fluctuating selection and dynamic adaptive landscapes. In: Svensson EI, Calsbeek R, eds. The adaptive landscape in evolutionary biology. Oxford, UK: Oxford University Press. pp 89–109.

Cam HP, Noma K, Ebina H, Levin HL, Grewal SIS (2008) Host genome surveillance for retrotransposons by transposon-derived proteins. Nature 451: 431-6.

Camello-Almaraz C, Gomez-Pinilla PJ, Pozo MJ, Camello PJ (2006) Mitochondrial reactive oxygen species and Ca2+ signaling. Am J Physiol Cell Physiol 291: C1082-8.

Cameron TC, Wearing HJ, Rohani P, Sait SM (2007) Two-species asymmetric competition: effects of age structure on intra- and interspecific interactions. J Anim Ecol 76: 83–93.

Campbell EM, Ball A, Hoppler S, Bowman AS (2008) Invertebrate aquaporins: a review. J Comp Physiol B 178: 935-55.

Campbell PM, Pottinger TG, Sumpter JP (1992) Stress reduces the quality of gametes produced by rainbow trout.Biol Reprod 47: 1140-50.

Campisi J, Gray HE, Pardee AB, Dean M, Sonenshein GE (1984) Cell-cycle control of c-Myc but not c-ras expression is lost following chemical transformation. Cell 36: 241–7.

Campos AC, Molognoni F, Melo FH, Galdieri LC, Carneiro CR, et al. (2007) Oxidative stress modulates DNA methylation during melanocyte anchorage blockade associated with malignant transformation. Neoplasia 9: 1111-21.

Campos PRA, de Oliveira VM (2004) Mutational effects on the clonal interference phenomenon. Evolution 58: 932–7.

Campos PRA, Adami C, Wilke CO (2004) Modeling stochastic clonal interference. In: Ciobanu G, Rozenberg G, eds. Modeling in molecular biology. Springer Series in Natural Computing. Berlin, Germany: Springer. pp 21–39.

Campos PRA, Neto PSCA, de Oliveira VM, Gordo I (2008) Environmental heterogeneity enhances clonal interference. Evolution 62: 1390–9.

Campos PR, Wahl LM (2010) The adaptation rate of asexuals: deleterious mutations, clonal interference and population bottlenecks. Evolution 64:1973-83.

Camps C, Buffa FM, Colella S, Moore J, Sotiriou C, et al. (2008) hsa-miR-210 is induced by hypoxia and is an independent prognostic factor in breast cancer. Clin Cancer Res 14: 1340–8.

Camps M, Herman A, Loh E, Loeb LA (2007) Genetic constraints on protein evolution. Crit Rev Biochem Mol Biol 42: 313-26.

Canfield DE, Teske A (1996) Late Proterozoic rise in atmospheric oxygen concentration inferred from phylogenetic and sulphur-isotope studies. Nature 382: 127–32.

Canfield DE, Poulton SW, Narbonne GM (2007) Late-Neoproterozoic deep-ocean oxygenation and the rise of animal life. Science 315: 92–5.

Cannino G, Di Liegro CM, Rinaldi AM (2007) Nuclear–mitochondrial interaction. Mitochondrion 7: 359–66.

Cannon B, Shabalina IG, Kramarova TV, Petrovic N, Nedergaard J (2006) Uncoupling proteins: A role in protection against reactive oxygen species—or not? Biochim Biophys Acta 1757: 449–58.

Cano CE, Gommeaux J, Pietri S, Culcasi M, Garcia S, et al. (2009) Tumor protein 53-induced nuclear protein 1 is a major mediator of p53 antioxidant function. Cancer Res 69: 219–26.

Cano L, Escarre J, Fleck I, Blanco-Moreno JM, Sans FX (2008) Increased fitness and plasticity of an invasive species in its introduced range: a study using Senecio pterophorus. J Ecol 96: 468–76.

Cant MA (2006) A tale of two theories: parent-offspring conflict and reproductive skew. Anim Behav 71: 255–63.

Cant MA (2012) Suppression of social conflict and evolutionary transitions to cooperation. Am Nat 179: 293-301.

Cao L, Shitara H, Horii T, Nagao Y, Imai H, et al. (2007) The mitochondrial bottleneck occurs without reduction of mtDNA content in female mouse germ cells. Nat Genet 39: 386–90.

Cao L, Shitara H, Sugimoto M, Hayashi J-I, Abe K, et al. (2009) New evidence confirms that the mitochondrial bottleneck is generated without reduction of mitochondrial DNA content in early primordial germ cells of mice. PLoS Genet 5: e1000756.

Cao X, Kambe F, Ohmori S, Seo H (2002) Oxidoreductive modification of two cysteine residues in paired domain by Ref-1 regulates DNA-binding activity of Pax-8. Biochem Biophys Res Commun 297: 288–93.

Caporale LH (1999) Chance favors the prepared genome. Ann NY Acad Sci 870:1-21.

Caporale LH (2003a) Darwin in the genome: molecular strategies in biological evolution. New York, NY: McGraw-Hill.

Caporale LH (2003b) Natural Selection and the emergence of a mutation phenotype: An update of the evolutionary synthesis considering mechanisms that affect genome variation. Annu Rev Microbiol 57: 467–85.

Caporale LH (2009) Putting together the pieces: evolutionary mechanisms at work within genomes. BioEssays 31: 700–2.

Capp JP (2010) Noise-driven heterogeneity in the rate of genetic-variant generation as a basis for evolvability. Genetics 185: 395-404.

Capra JA, Pollard KS (2011) Substitution patterns are GC-biased in divergent sequences across the metazoans. Genome Biol Evol 3:516-27.

Capy P, Gasperi G, Biémont C, Bazin C (2000) Stress and transposable elements: co-evolution or useful parasites? Heredity 85: 101–6.

Cardillo M (1999) Latitude and rates of diversification in birds and butterflies. Proc R Soc Lond B 266: 1221-5.

Cardillo M, Orme CDL, Owens IPF (2005) Testing for latitudinal bias in diversification rates: an example using New World birds. Ecology 86: 2278-87.

Cardoso AA, Jiang Y, Luo M, Reed AM, Shahda S, et al. (2012) APE1/Ref-1 regulates STAT3 transcriptional activity and APE1/Ref-1–STAT3 dual-targeting effectively inhibits pancreatic cancer cell survival. PLoS ONE 7: e47462.

Carey CC, Gorman KF, Rutherford S (2006) Modularity and intrinsic evolvability of Hsp90-buffered change. PLoS ONE 1: e76.

Cargill SL, Carey JR, Müller HG, Anderson G (2003) Age of ovary determines remaining life expectancy in old ovariectomized mice. Aging Cell 2: 185-90.

Carlborg O, Jacobsson L, Ahgren P, Siegel P, Andersson L (2006) Epistasis and the release of genetic variation during long-term selection. Nat Genet 38: 418-20.

Carlson KM, Bracamontes J, Jackson CE, Clark R, Lacroix A, et al. (1994) Parent-of-origin effects in multiple endocrine neoplasia type 2B. Am J Hum Genet 55: 1076–82.

Carlson LL, Page AW, Bestor TH (1992) Properties and localization of DNA methyltransferase in preimplantation mouse embryos: implications for genomic imprinting. Genes Dev 6: 2536–41.

Carrillo C, Lu Z, Borca MV, Vagnozzi A, Kutish GF, Rock DL (2007) Genetic and phenotypic variation of foot-and-mouth disease virus during serial passages in a natural host. J Virol 81: 11341–51.

Carlsen E, Giwercman A, Keiding N, Skakkebaek NE (1992) Evidence for decreasing quality of semen during the past 50 years. Br Med J 305: 609-13.

Carlson CA, Pierson LS, Rosen JJ, Ingraham JL (1983) Pseudomonas stutzeri and related species undergo natural transformation. J Bacteriol 153: 93-9.

Carlson KM, Bracamontes J, Jackson CE, Clark R, Lacroix A, Wells SA Jr, Goodfellow PJ (1994) Parent-of-origin effects in multiple endocrine neoplasia type 2B. Am J Hum Genet 55:1076-82.

Carmeliet P, Dor Y, Herbert JM, Fukumura D, Brusselmans K, et al. (1998) Role of HIF-1a in hypoxia-mediated apoptosis, cell proliferation and tumour angiogenesis. Nature 394: 485–90.

Carmell MA, Girard A, van de Kant HJ, Bourc’his D, Bestor TH, et al. (2007) MIWI2 is essential for spermatogenesis and repression of transposons in the mouse male germline. Dev Cell 12: 503–14.

Carmichael AN, Fridolfsson AK, Halverson J, Ellegren H (2000) Male-biased mutation rates revealed from Z and W chromosome-linked ATP synthase alpha-subunit (ATP5A1) sequences in birds. J Mol Evol 50: 443-7.

Carneiro M, Hartl DL (2010) Adaptive landscapes and protein evolution. Proc Natl Acad Sci USA 107 (suppl. 1): 1747–51.

Carninci P, Kasukawa T, Katayama S, Gough J, Frith MC, et al. (2005) The transcriptional landscape of the mammalian genome. Science 309: 1559–63.

Caro AA, Cederbaum AI (2004) Antioxidant properties of S-adenosyl-L-methionine in Fe2+-initiated oxidations. Free Radic Biol Med 36: 1303-16.

Carone BR, Fauquier L, Habib N, Shea JM, Hart CE, et al. (2010) Paternally induced transgenerational environmental reprogramming of metabolic gene expression in mammals. Cell 143: 1084–96.

Carragher JF, Sumpter JP, Pottinger TG, Pickering AD (1989) The deleterious effects of cortisol implantation on reproductive function in two species of trout, Salmo trutta L. and Salmo gairdneri Richardson. Gen Comp Endocrinol 76: 310-21.

Carrasco AJ, Dzeja PP, Alekseev AE, Pucar D, Zingman LV, et al. (2001) Adenylate kinase phosphotransfer communicates cellular energetic signals to ATP-sensitive potassium channels. Proc Natl Acad Sci USA 98: 7623-8.

Carré D, Djediat C, Sardet C (2002) Formation of a large Vasa-positive granule and its inheritance by germ cells in the enigmatic Chaetognaths. Development 129: 661-70.

Carrera-Bastos P, Fontes M, O’Keefe JH, Lindeberg S, Cordain L (2011) The Western Diet and lifestyle and diseases of civilization. Res Reports Clin Cardiol 2:15–35.

Carrick FN, Setchell BP (1977) The evolution of the scrotum. In: Calaby JH, Tyndale-Biscoe CH, eds. Reproduction and evolution.Canberra, Australia: Australian Academy of Science. pp 165–170.

Carroll SB (2005) Endless Forms Most Beautiful. New York, NY: Norton.

Carroll SP, Hendry AP, Reznick DN, Fox CW (2007) Evolution on ecological time-scales. Funct Ecol 21: 387-93.

Carson HL (1990) Increased genetic variance after a population bottleneck. Trends Ecol Evol 5: 228-30.

Carter AJR, Hermisson J, Hansen TF (2005) The role of epistatic gene interactions in the response to selection and the evolution of evolvability. Theor Popul Biol 68:179–96.

Carton Y, Nappi AJ, Poirie M (2005) Genetics of anti-parasite resistance in invertebrates. Dev Comp Immunol 29: 9–32.

Carvalho GR, Hughes RN (1983) The effect of food availability female culture density and photoperiod on ephippia production in Daphnia magna (Crustacea Cladocera). Freshwater Biol 13: 37–46.

Casaccia-Bonnefil P, Gu A, Khursigara G, Chao MV (1999) The p75 neurotrophin receptor as a modulator of survival and death decisions. Microsc Res Tech 45: 217-24.

Casacuberta E, González J (2013) The impact of transposable elements in environmental adaptation. Mol Ecol 22: 1503-17.

Cases-González C, Arribas M, Domingo E, Lázaro E (2008) Beneficial effects of population bottlenecks in an RNA virus evolving at increased error rate. J Mol Biol 384: 1120-9.

Cash TP, Pan Y, Simon MC (2007) Reactive oxygen species and cellular oxygen sensing. Free Radic Biol Med 43: 1219–25.

Cassone VM (1990) Effects of melatonin on vertebrate circadian systems. Trends Neurosci 13: 457-64.

Castagnone-Sereno P (2006) Genetic variability and adaptive evolution in parthenogenetic root-knot nematodes.Heredity 96: 282–9.

Castel SE, Martienssen RA (2013) RNA interference in the nucleus: roles for small RNAs in transcription, epigenetics and beyond. Nat Rev Genet 14: 100–12.

Castellana S, Vicario S, Saccone C (2011) Evolutionary patterns of the mitochondrial genome in Metazoa: exploring the role of mutation and selection in mitochondrial protein-coding genes. Genome Biol Evol 3: 1067–79.

Castello PR, David PS, McClure T, Crook Z, Poyton RO (2006)Mitochondrial cytochrome oxidase produces nitric oxide under hypoxic conditions: implications for oxygen sensing and hypoxic signaling in eukaryotes.Cell Metab 3: 277-87.

Castello PR, Woo DK, Ball K, Wojcik J, Liu L, Poyton RO (2008)Oxygen-regulated isoforms of cytochrome c oxidase have differential effects on its nitric oxide production and on hypoxic signaling.Proc Natl Acad Sci USA 105: 8203-8.

Castillo-Davis CI, Mekhedov SL, Hartl DL, Koonin EV, Kondrashov FA (2002) Selection for short introns in highly expressed genes. Nat Genet 31: 415-8.

Castonguay E, Angers B (2012) The key role of epigenetics in the persistence of asexual lineages. Genet Res Int 2012: 534289.

Castrillon DH, Quade BJ, Wang TY, Quigley C, Crum CP (2000) The human VASA gene is specifically expressed in the germ cell lineage. Proc Natl Acad Sci USA 97: 9585–90.

Catling DC, Glein CR, Zahnle KJ, McKay CP (2005) Why O2 is required by complex life on habitable planets and the concept of planetary ‘‘oxygenation time’’. Astrobiology 5: 415–38.

Causevic A, Gentil MV, Delaunay A, El-Soud WA, Garcia Z, et al. (2006) Relationship between DNA methylation and histone acetylation levels, cell redox and cell differentiation states in sugarbeet lines. Planta 224:812-27.

Cavalier-Smith T (2002) Origins of the machinery of recombination and sex. Heredity 88: 125–41.

Ceccarelli D, Gallesi D, Giovannini F, Ferrali M, Masini A (1995) Relationship between free iron level and rat liver mitochondrial dysfunction in experimental dietary iron overload. Biochem Biophys Res Commun 209: 53-9.

Cek M, Cûrgûl S, Ertas M, Sözer T, Alpacar Z (1992) Seminal transferrin levels in seminal plasma of fertile and infertile men. Urol Int 49: 218-21.

Celerin M, Merino ST, Stone JE, Menzie AM, Zolan ME (2000) Multiple roles of Spo11 in meiotic chromosome behavior. EMBO J 19: 2739–50.

Celeste A, Petersen S, Romanienko PJ, Fernandez-Capetillo O, Chen HT, et al. (2002) Genomic instability in mice lacking histone H2AX. Science 296: 922-7.

Celino FT, Yamaguchi S, Miura C, Ohta T, Tozawa Y, et al. (2011) Tolerance of spermatogonia to oxidative stress is due to high levels of Zn and Cu/Zn superoxide dismutase. PLoS ONE 6: e16938.

Cerda S, Weitzman SA (1997) Influence of oxygen radical injury on DNA methylation. Mutat Res 386: 141–52.

Cerutti H, Casas-Mollano JA (2006) On the origin and functions of RNA-mediated silencing: from protists to man. Curr Genet 50: 81–99.

Cerutti PA (1985) Prooxidant states and tumor promotion. Science 227: 375-81.

Cervera MT, Ruiz-Garcia L, Martinez-Zapater JM (2002) Analysis of DNA methylation in Arabidopsis thaliana based on methylation-sensitive AFLP markers. Mol Genet Genomics 268: 543–52.

César CE, Álvarez L, Bricio C, van Heerden E, Littauer D, Berenguer J (2011) Unconventional lateral gene transfer in extreme thermophilic bacteria. Int Microbiol 14: 187-99.

Chabory E, Damon C, Lenoir A, Henry-Berger J, Vernet P, et al. (2010) Mammalian glutathione peroxidases control acquisition and maintenance of spermatozoa integrity. J Anim Sci 88: 1321-31.

Chabot B, Stephenson DA, Chapman VM, Besmer P, Bernstein A (1988) The proto-oncogene c-kit encoding a transmembrane tyrosine kinase receptor maps to the mouse W locus. Nature 335: 88-9.

Chacinska A, Koehler CM, Milenkovic D, Lithgow T, Pfanner N (2009) Importing mitochondrial proteins: machineries and mechanisms. Cell 138: 628–44.

Chain BM, Myers R (2005) Variability and conservation in hepatitis B virus core protein. BMC Microbiol 5:33.

Chainy GB, Samantaray S, Samanta L (1997) Testosterone-induced changes in testicular antioxidant system. Andrologia 29: 343-9.

Chaki SP, Misro MM, Ghosh D, Gautam DK, Srinivas M (2005) Apoptosis and cell removal in the cryptorchid rat testis. Apoptosis 10:395-405.

Chakir M, Chafik A, Moreteau B, Gibert P, David JR (2002) Male sterility thresholds in Drosophila: D. simulans appears more cold-adapted than its sibling D. melanogaster. Genetica 114: 195-205.

Chakraborty R, Nei M (1982) Genetic differentiation of quantitative characters between populations or species. I. Mutation and random genetic drift. Genet Res 39: 303-14.

Chakravarthy MV, Booth FW (2004) Eating, exercise, and “thrifty” genotypes: connecting the dots toward an evolutionary understanding of modern chronic diseases. J Appl Physiol 96:3–10.

Chalmers DJ (1990) The evolution of learning: an experiment in genetic connectionism. In: Touretzky DS, Elman JL, Sejnowski TJ, Hinton GE, eds. Proceedings of the 1990 Connectionist Models Summer School. San Mateo, CA: Morgan Kaufmann.

Chamary JV, Parmley JL, Hurst LD (2006) Hearing silence: non-neutral evolution at synonymous sites in mammals. Nat Rev Genet 7:98–108.

Chamberlain NL, Driver ED, Miesfeld RL (1994) The length and location of CAG trinucleotide repeats in the androgen receptor N-terminal domain affect transactivation function. Nucleic Acids Res 22: 3181-6.

Chambeyron S, Popkova A, Payen-Groschene G, Brun C, Laouini D, et al. (2008) piRNA-mediated nuclear accumulation of retrotransposon transcripts in the Drosophila female germline. Proc Natl Acad Sci USA 105: 14964–9.

Champagne FA (2008) Epigenetic mechanisms and the transgenerational effects of maternal care. Front Neuroendocrinol 29: 386–97.

Chan MF, van Amerongen R, Nijjar T, Cuppen E, Jones PA, Laird PW (2001) Reduced rates of gene loss, gene silencing, and gene mutation in Dnmt1-deficient embryonic stem cells. Mol Cell Biol 21: 7587-600.

Chan SY, Zhang YY, Hemann C, Mahoney CE, Zweier JL, Loscalzo J (2009) MicroRNA-210 controls mitochondrial metabolism during hypoxia by repressing the iron-sulfur cluster assembly proteins ISCU1/2. Cell Metab 10: 273–84.

Chan SY, Loscalzo J (2010) MicroRNA-210: a unique and pleiotropic hypoxamir. Cell Cycle (Georgetown, Tex.) 9: 1072-83.

Chan YC, Banerjee J, Choi SY, Sen CK (2012) miR-210: the master hypoxamir. Microcirculation 19: 215-23.

Chance B, Sies H, Boveris A (1979) Hydroperoxide metabolism in mammalian organs. Physiol Rev 59: 527-605.

Chand D, Lovejoy DA (2011) Stress and reproduction: controversies and challenges. Gen Comp Endocrinol 171: 253-7.

Chandel NS, Maltepe E, Goldwasser E, Mathieu CE, Simon MC, Schumacker PT (1998) Mitochondrial reactive oxygen species trigger hypoxia-induced transcription. Proc Natl Acad Sci USA 95: 11715-20.

Chandel NS, McClintock DS, Feliciano CE, Wood TM, Melendez JA, et al. (2000) Reactive oxygen species generated at mitochondrial complex III stabilize hypoxia-inducible factor-1alpha during hypoxia: a mechanism of O2 sensing. J Biol Chem 275: 25130-8.

Chandler VL (2007) Paramutation: from maize to mice. Cell 128: 641-5.

Chandler VL, Stam M (2004) Chromatin conversations: mechanisms and implications of paramutation. Nat Rev Genet 5: 532–44.

Chang BH, Shimmin LC, Shyue SK, Hewett-Emmett D, Li WH (1994) Weak male-driven molecular evolution in rodents. Proc Natl Acad Sci USA 91: 827-31.

Chang BH, Li WH (1995) Estimating the intensity of male-driven evolution in rodents by using X-linked and Y-linked Ube 1 genes and pseudogenes. J Mol Evol 40: 70-7.

Chang CL, Marra G, Chauhan DP, Ha HT, Chang DK, et al. (2002) Oxidative stress inactivates the human DNA mismatch repair system. Am J Physiol Cell Physiol 283: C148–C154.

Chang DK, Metzgar D, Wills C, Boland CR (2001) Microsatellites in the eukaryotic DNA mismatch repair genes as modulators of evolutionary mutation rate. Genome Res 11: 1145-6.

Chang IY, Kim SH, Cho HJ, Lee DY, Kim MH, et al. (2005) Human AP endonuclease suppresses DNA mismatch repair activity leading to microsatellite instability. Nucleic Acids Res 33: 5073-81.

Chang S, Wang RH, Akagi K, Kim KA, Martin BK, et al. (2011) Tumor suppressor BRCA1 epigenetically controls oncogenic microRNA-155. Nat Med 17: 1275-82.

Chang S, Sharan SK (2012) Epigenetic control of an oncogenic microRNA, miR-155, by BRCA1. Oncotarget 3: 5-6.

Chang TC, Wentzel EA, Kent OA, Ramachandran K, Mullendore M, et al. (2007) Transactivation of miR-34a by p53 broadly influences gene expression and promotes apoptosis. Mol. Cell 26: 745–52.

Changeux JP, Courrege P, Danchin A (1973) A theory of the epigenesis of neuronal networks by selective stabilization of synapses. Proc Natl Acad Sci USA 70: 2974-8.

Changeux JP, Danchin A (1976) Selective stabilisation of developing synapses as a mechanism for the specification of neuronal networks. Nature 264: 705-12.

Chao HT, Lee SY, Lee HM, Liao TL, Wei YH, Kao SH (2005) Repeated ovarian stimulations induce oxidative damage and mitochondrial DNA mutations in mouse ovaries. Ann NY Acad Sci 1042: 148-56.

Chao L (1990) Fitness of RNA virus decreased by Muller’s ratchet. Nature 348: 454–5.

Chao L, Cox EC (1983) Competition between high and low mutating strains of Escherichia coli. Evolution 37: 125-34.

Chao L, Vargas C, SpearBB, Cox EC (1983) Transposable elements as mutator genes in evolution. Nature 303: 633-5.

Chao L, Tran TT, Tran TT (1997) The advantage of sex in the RNA virus phi6. Genetics 147: 953–9.

Chao L, Hanley KA, Burch CL, Dahlberg C, Turner PE (2000) Kin selection and parasite evolution: higher and lower virulence with hard and soft selection. Q Rev Biol 75: 261–75.

Chapin RE, Phelps J (1990) Recent advances in testicular cell culture: Implications for toxicology. Toxicol In Vitro 4:543-59.

Chapman LJ, Galis F, Shinn J (2000) Phenotypic plasticity and the possible role of genetic assimilation: hypoxia induced trade-offs in the morphological traits of an African cichlid. Ecol Lett 3: 387–93.

Chapman T (2006) Evolutionary conflicts of interest between males and females. Curr Biol 16: R744-54.

Chapman T, Liddle LF, Kalb JM, Wolfner MF, Partridge L (1995) Costs of mating in Drosophila melanogaster females is mediated by male accessory gland products. Nature 373: 241–4.

Chapman T, Partridge L (1996) Sexual conflict as fuel for evolution. Nature 381: 189–90.

Charaabi K, Carletto J, Chavigny P, Marrakchi M, Makni M, Vanlerberghe-Masutti F (2008) Genotypic diversity of the cotton-melon aphid Aphis gossypii (Glover) in Tunisia is structured by host plants. Bull Entomol Res 98: 333-41.

Chardard D, Dournon C (1999) Sex reversal by aromatase inhibitor treatment in the newt Pleurodeles waltl. J Exp Zool 283: 43–50.

Charlesworth B (1976) Recombination modification in a fluctuating environment. Genetics 83: 181–95.

Charlesworth B (1994) The effect of background selection against deleterious mutations on weakly selected, linked variants. Genet Res 63: 213–27.

Charlesworth B (2009) Fundamental concepts in genetics: effective population size and patterns of molecular evolution and variation. Nat Rev Genet 10:195-205.

Charlesworth B (2012) The effects of deleterious mutations on evolution at linked sites. Genetics 190: 5-22.

Charlesworth B, Langley CH (1989) The population genetics of Drosophila transposable elements. Annu Rev Genet 23: 251-87.

Charlesworth B, Barton NH (1996) Recombination load associated with selection for increased recombination. Genet Res 67: 27–41.

Charlesworth B, Charlesworth D (1997) Rapid fixation of deleterious alleles can be caused by Muller's ratchet. Genet Res 70: 63-73.

Charlesworth D, Charlesworth B (1987) Inbreeding depression and its evolutionary consequences. Annu Rev Ecol Syst 18: 237–68.

Charlesworth D, Morgan MT, Charlesworth B (1993) Mutation accumulation in finite outbreeding and inbreeding populations. Genet Res 61: 39–56.

Charlesworth J, Eyre-Walker A (2006) The rate of adaptive evolution in enteric bacteria. Mol Biol Evol 23: 1348–56.

Charmantier A, Blondel J, Perret P, Lambrechts MM (2004) Do extra-pair paternities provide genetic benefits for female blue tits Parus caeruleus? J Avian Biol 35: 524–32.

Charmantier A, Garant D (2005) Environmental quality and evolutionary potential: lessons from wild populations. Proc R Soc Lond Ser B 272: 1415-25.

Charnov EL, ErnestSKM (2006) The offspring-size⁄clutch-size trade-off in mammals. Am Nat 167: 578–82.

Chasnov JR (2000) Mutation-selection balance, dominance and the maintenance of sex. Genetics 156:1419-25.

Chastain PD 2nd, Nakamura J, Swenberg J, Kaufman D (2006) Nonrandom AP site distribution in highly proliferative cells. FASEB J 20: 2612-4.

Chatterjee S, Grosshans H (2009) Active turnover modulates mature microRNA activity in Caenorhabditis elegans. Nature 461: 546-9.

Chavan GV, Chopde PR (1982) Polygenic variability, heritability and genetic advance in irradiated sesame. J Marathwada Agric Univ 7: 17-9.

Chavarrias D, López-Fanjul C, García-Dorado A (2001) The rate of mutation and the homozygous and heterozygous mutational effects for competitive viability: a long-term experiment with Drosophila melanogaster. Genetics 158: 681-93.

Chawla S, Bading H (1998) Function of nuclear and cytoplasmic calcium in the control of gene expression. In: Verkhratsky A, Toescu EC, eds. Integrative Aspects of Calcium Signalling. New York, NY:Plenum Press. pp 59-78.

Chekanova JA, Gregory BD, Reverdatto SV, Chen H, Kumar R, et al. (2007) Genome-wide high-resolution mapping of exosome substrates reveals hidden features in the Arabidopsis transcriptome. Cell 131: 1340-53.

Chemes H (1986) The phagocytic function of Sertoli cells: a morphological, biochemical, and endocrinological study of lysosomes and acid phosphatase localization in the rat testis. Endocrinology 119: 1673–81.

Chen B, Walser JC, Rodgers TH, Sobota RS, Burke MK, et al. (2007) Abundant, diverse, and consequential P elements segregate in promoters of small heat-shock genes in Drosophila populations. J Evol Biol 20: 2056-66.

Chen B, Wagner A (2012) Hsp90 is important for fecundity, longevity, and buffering of cryptic deleterious variation in wild fly populations. BMC Evol Biol 12: 25.

Chen C, Nott TJ, Jin J, Pawson T (2011) Deciphering arginine methylation: Tudor tells the tale. Nat Rev Mol Cell Biol 12: 629–42.

Chen CL, Schroeder MC, Kango-Singh M, Tao C, Halder G (2012) Tumor suppression by cell competition through regulation of the Hippo pathway. Proc Natl Acad Sci USA 109: 484-9.

Chen D, Ma H, Hong H, Koh SS, Huang SM, et al. (1999) Regulation of transcription by a protein methyltransferase. Science 284: 2174–7.

Chen D, Toone WM, Mata J, Lyne R, Burns G, et al. (2003a) Global transcriptional responses of fission yeast to environmental stress. Mol Biol Cell 14: 214–229.

Chen D, Li M, Luo J, Gu W (2003b) Direct interactions between HIF-1 alpha and Mdm2 modulate p53 function. J Biol Chem 278: 13595–8.

Chen D, Thomas EL, Kapahi P (2009) HIF-1 modulates dietary restriction-mediated lifespan extension via IRE-1 in Caenorhabditis elegans. PLoS Genet5:e1000486.

Chen G, Bradford WD, Seidel CW, Li R (2012) Hsp90 stress potentiates rapid cellular adaptation through induction of aneuploidy. Nature 482: 246-50.

Chen H, Zirkin BR (1999) Long-term suppression of Leydig cell steroidogenesis prevents Leydig cell aging. Proc Natl Acad Sci USA 96: 14877-81.

Chen H, Cangello D, Benson S, Folmer J, Zhu H, et al. (2001) Age-related increase in mitochondrial superoxide generation in the testosterone-producing cells of Brown Norway rat testes: relationship to reduced steroidogenic function? Exp Gerontol 36: 1361–73.

Chen H, Liu J, Luo L, Baig MU, Kim JM, Zirkin BR (2005) Vitamin E, aging and Leydig cell steroidogenesis. Exp Gerontol 40: 728–36.

Chen H,YanY, Davidson TL, Shinkai Y, CostaM(2006) Hypoxic stress induces dimethylated histone H3 lysine 9 through histone methyltransferase G9a in mammalian cells. Cancer Res 66: 9009–16.

Chen J, Tomkinson AE, Ramos W, Mackey ZB, Danehower S, et al. (1995) Mammalian DNA ligase III: Molecular cloning, chromosomal localization and expression in spermatocytes undergoing meiotic recombination. Mol Cell Biol 15: 5412-22.

Chen JM, Cooper DN, Chuzhanova N, Férec C, Patrinos GP (2007) Gene conversion: mechanisms, evolution and human disease. Nat Rev Genet 8:762-75.

Chen K, Albano A, Ho A, Keaney JF Jr (2003) Activation of p53 by oxidative stress involves platelet-derived growth factor-beta receptor-mediated ataxia telangiectasia mutated (ATM) kinase activation. J Biol Chem 278: 39527-33.

Chen K (2004) Organization of MAO A and MAO B promoters and regulation of gene expression. Neurotoxicology 25: 31-6.

Chen K, Rajewsky N (2006) Natural selection on human microRNA binding sites inferred from SNP data. Nat Genet 38: 1452–6.

Chen K, Ou XM, Wu JB, Shih JC (2011) Transcription factor E2F-associated phosphoprotein (EAPP), RAM2/CDCA7L/JPO2 (R1), and simian virus 40 promoter factor 1 (Sp1) cooperatively regulate glucocorticoid activation of monoamine oxidase B. Mol Pharmacol 79: 308-17.

Chen MJ, Sepramaniam S, Armugam A, Choy MS, Manikandan J, et al. (2008) Water and ion channels: crucial in the initiation and progression of apoptosis in central nervous system? CurrNeuropharmacol 6: 102–16.

Chen Q, Duan EK (2011) Aquaporins in sperm osmoadaptation: an emerging role for volume regulation. Acta Pharmacol Sin 32: 721-4.

Chen R, Holmes EC (2006) Avian influenza virus exhibits rapid evolutionary dynamics. Mol Biol Evol 23: 2336-41.

Chen RS, McDonald BA (1996) Sexual reproduction plays a major role in the genetic structure of populations of the fungus Mycosphaerella graminicola. Genetics 142: 1119-27.

Chen RZ, Pettersson U, Beard C, Jackson-Grusby L, Jaenisch R (1998)DNA hypomethylation leads to elevated mutation rates.Nature 395: 89-93.

Chen T, Li E (2006) Establishment and maintenance of DNA methylation patterns in mammals. Curr Top Microbiol Immunol 301: 179–201.

Chen W, Jinks-Robertson S (1998) Mismatch repair proteins regulate heteroduplex formation during mitotic recombination in yeast. Mol Cell Biol 18: 6525–37.

Chen W, Sun Z, Wang XJ, Jiang T, Huang Z, et al. (2009) Direct interaction between Nrf2 and p21(Cip1/WAF1) upregulates the Nrf2-mediated antioxidant response. Mol Cell 34: 663–73.

Chen Y, Roxby R (1997) Identification of a functional CT-element in the Phytophthora infestans piypt1 gene promoter. Gene 198: 159-64.

Chen YF, Meng QM (1991) Sexual dimorphism of blood pressure in spontaneously hypertensive rats is androgen dependent. Life Sci 48: 85–96.

Chen Z, Li Y, Zhang H, Huang P, Luthra R (2010) Hypoxia-regulated microRNA-210 modulates mitochondrial function and decreases ISCU and COX10 expression.Oncogene 29: 4362-8.

Chen Z, Zhao T, Xu Y (2012) The genomic stability of induced pluripotent stem cells.Protein Cell 3: 271-7.

Chen ZJ (2007)Genetic and epigenetic mechanisms for gene expression and phenotypic variation in plant polyploids.Annu Rev Plant Biol 58: 377-406.

Chen ZJ (2010)Molecular mechanisms of polyploidy and hybrid vigor.Trends Plant Sci 15: 57-71.

Cheng J, Kapranov P, Drenkow J, Dike S, Brubaker S, et al. (2005) Transcriptional maps of 10 human chromosomes at 5-nucleotide resolution. Science 308: 1149–54.

Cheng X, Blumenthal RM, eds. (1999) S-Adenosylmethionine-dependent Methyltransferases: Structures and Functions.Singapore: World Scientific.

Cheng YH, Wong EW, Cheng CY (2011)Cancer/testis (CT) antigens, carcinogenesis and spermatogenesis.Spermatogenesis 1: 209-20.

Cheo DL, Bayles KW, Yasbin RE (1993) Elucidation of regulatory elements that control damage induction and competence induction of the Bacillus subtilis SOS system. J Bacteriol 175: 5907-15.

Cheok CF, Verma CS, Baselga J, Lane DP (2011) Translating p53 into the clinic. Nat Rev Clin Oncol 8: 25-37.

Cherry LM, Case SM, Kunkel JG, Wyles JS, Wilson AC (1982) Body-shape metrics and organismal evolution. Evolution 36: 914-33.

Chesson PL, Warner RR (1981) Environmental variability promotes coexistence in lottery competitive systems. Am Nat 117: 923–43.

Chesson P, Huntly N (1997) The roles of harsh and fluctuating conditions in the dynamics of ecological communities. Am Nat 150: 519–53.

Cheung P, Allis CD, Sassone-Corsi P (2000) Signaling to chromatin through histone modifications. Cell 103:263-71.

Chevallet M, Wagner E, Luche S, van Dorsselaer A, Leize-Wagner E, Rabilloud T (2003) Regeneration of peroxiredoxins during recovery after oxidative stress: only some overoxidized peroxiredoxins can be reduced during recovery after oxidative stress. J Biol Chem 278: 37146–53.

Chevin LM, Hospital F (2008) Selective sweep at a quantitative trait locus in the presence of background genetic variation. Genetics 180: 1645–60.

Chevion M (1988) A site-specific mechanism for free radical induced biological damage: the essential role of redox-active transition metals. Free Radic Biol Med 5: 27-37.

Chiacchiera F, Piunti A, Pasini D (2013) Epigenetic methylations and their connections with metabolism.Cell Mol Life Sci 70: 1495–508.

Chiang SM, Schellhorn HE (2010) Evolution of the RpoS regulon: origin of RpoS and the conservation of RpoS-dependent regulation in bacteria. J Mol Evol 70: 557–71.

Chiarugi P, Fiaschi T, Taddei ML, Talini D, Giannoni E, et al. (2001)Two vicinal cysteines confer a peculiar redox regulation to low molecular weight protein tyrosine phosphatase in response to platelet-derived growth factor receptor stimulation.J Biol Chem 276: 33478-87.

Chiarugi P, Fiaschi T (2007) Redox signalling in anchorage-dependent cell growth. Cell Signal 19: 672–82.

Chieffi P, Minucci S, Cobellis G, Fasano S, Pierantoni R (1995)Changes in proto-oncogene activity in the testis of the frog, Rana esculenta, during the annual reproductive cycle.Gen Comp Endocrinol 99: 127-36.

Chieffi P, Angelini F, Pierantoni R (1997)Proto-oncogene activity in the testis of the lizard, Podarcis s. sicula, during the annual reproductive cycle.Gen Comp Endocrinol 108: 173–81.

Chignalia AZ, Schuldt EZ, Camargo LL, Montezano AC, Callera GE, et al. (2012)Testosterone induces vascular smooth muscle cell migration by NADPH oxidase and c-Src-dependent pathways.Hypertension 59: 1263-71.

Childs DZ, Rees M, Rose KE, Grubb PJ,Ellner SP (2004) Evolution of size-dependent flowering in a variable environment: construction and analysis of a stochastic integral projection model. Proc R Soc Lond B 271: 425–34.

Childs DZ, Metcalf CJ, Rees M (2010)Evolutionary bet-hedging in the real world: empirical evidence and challenges revealed by plants.Proc Biol Sci 277: 3055-64.

Chimpanzee Sequencing and Analysis Consortium (2005) Initial sequence of the chimpanzee genome and comparison with the human genome. Nature 437: 69–87.

Chinnusamy V, Zhu JK (2009) Epigenetic regulation of stress responses in plants. Curr Opin Plant Biol 12: 133–9.

Chisu V, Manca P, Lepore G, Gadau S, Zedda M, Farina V (2006)Testosterone induces neuroprotection from oxidative stress. Effects on catalase activity and 3-nitro-L-tyrosine incorporation into alpha-tubulin in a mouse neuroblastoma cell line.Arch Ital Biol 144: 63-73.

Chitwood DH, Timmermans MC (2010) Small RNAs are on the move. Nature 467: 415–9.

Chiu SM, Xue LY, Friedman LR, Oleinick NL (1993) Copper ion-mediated sensitization of nuclear matrix attachment sites to ionizing radiation. Biochemistry 32: 6214-9.

Chmurzynska A (2010) Fetal programming: link between early nutrition, DNA methylation, and complex diseases. Nutr Rev 68: 87-98.

Cho SH, Lee CH, Ahn Y, Kim H, Kim H, et al. (2004) Redox regulation of PTEN and protein tyrosine phosphatases in H2O2 mediated cell signaling. FEBS Lett 560: 7–13.

Chodavarapu RK, Feng S, Bernatavichute YV, Chen PY, Stroud H, et al. (2010)Relationship between nucleosome positioning and DNA methylation.Nature 466: 388-92.

Choi CS, Sano H (2007) Abiotic-stress induces demethylation and transcriptional activation of a gene encoding a glycerophosphodiesterase-like protein in tobacco plants. Mol Genet Genomics 277: 589–600.

Choi S-K, Yoon S-R, Calabrese P, Arnheim N (2008) A germ-line-selective advantage rather than an increased mutation rate can explain some unexpectedly common human disease mutations. Proc Natl Acad Sci USA 105: 10143–8.

Choi S-K, Yoon S-R, Calabrese P, Arnheim N (2012) Positive selection for new disease mutations in the human germline: evidence from the heritable cancer syndrome multiple endocrine neoplasia type 2B. PLoS Genet 8: e1002420.

Choi Y, Gehring M, Johnson L, Hannon M, Harada JJ, et al. (2002) DEMETER, a DNA glycosylase domain protein, is required for endosperm gene imprinting and seed viability in Arabidopsis. Cell 110: 33–42.

Choleva L, Apostolou A, Rab P, Janko K (2008)Making it on their own: sperm-dependent hybrid fishes (Cobitis) switch the sexual hosts and expand beyond the ranges of their original sperm donors. Philos Trans R Soc Lond B Biol Sci 363: 2911-9.

Chopra I, O'Neill AJ, Miller K (2003)The role of mutators in the emergence of antibiotic-resistant bacteria.Drug Resist Updat 6: 137-45.

Chou HH, Berthet J, Marx CJ (2009) Fast growth increases the selective advantage of a mutation arising recurrently during evolution under metal limitation. PLoS Genet 5: e1000652.

Chou HH, Chiu HC, Delaney NF, Segrè D, Marx CJ (2011) Diminishing returns epistasis among beneficial mutations decelerates adaptation.Science 332:1190-2.

Chouaib S, Bertoglio J, Blay JY, Marchiol-Fournigault C, Fradelizi D (1988) Generation of lymphokine-activated killer cells: synergy between tumor necrosis factor and interleukin 2. Proc Natl Acad Sci USA 85: 6875-9.

Chouard T (2010)Evolution: Revenge of the hopeful monster.Nature 463: 864-7.

Choudhury S, Chainy GB, Mishro MM (2003) Experimentally induced hypo- and hyper-thyroidism influence on the antioxidant defence system in adult rat testis. Andrologia 35: 131–40.

Choudhury S, Panda P, Sahoo L, Panda SK (2013)Reactive oxygen species signaling in plants under abiotic stress.Plant Signal Behav 8: pii: e23681. [Epub ahead of print]

Choulika A, Perrin A, Dujon B, Nicolas JF (1995) Induction of homologous recombination in mammalian chromosomes by using the I-SceI system of Saccharomyces cerevisiae. Mol Cell Biol 15: 1968–73.

Chowdhury VS, Ubuka T, Tsutsui K (2013)Review: Melatonin stimulates the synthesis and release of gonadotropin-inhibitory hormone in birds.Gen Comp Endocrinol 181: 175-8.

Christensen B (1959) Asexual reproduction in the Enchytraeidae (Olig.). Nature 184: 1159–60.

Christensen ST, Leick V, Rasmussen L, Wheatley DN (1998) Signaling in unicellular eukaryotes. Int Rev Cytol 177: 181-253.

Christenson LK, Stouffer RL (1997) Follicle-stimulating hormone and luteinizing hormone/chorionic gonadotropin stimulation of vascular endothelial growth factor production by macaque granulosa cells from pre- and periovulatory follicles. J Clin Endocrinol Metab 82: 2135–42.

Christenson LK, Strauss I, Jerome F (2001) Steroidogenic acute regulatory protein: An update on its regulation and mechanism of action. Arch Med Res 32: 576–86.

Christian JJ (1961) Phenomena associated with population density. Proc Natl Acad Sci USA 47: 428-49.

Christians E, Michel E, Adenot P, Mezger V, Rallu M, et al. (1997) Evidence for the involvement of mouse heat shock factor 1 in the atypical expression of the HSP70.1 heat shock gene during mouse zygotic genome activation. Mol Cell Biol 17: 778–88.

Christians ES, Zhou Q, Renard J, Benjamin IJ (2003) Heat shock proteins in mammalian development.Semin Cell Dev Biol 14: 283–90.

Christiansen FB, Otto SP, Bergman A, Feldman MW (1998)Waiting with and without recombination: the time to production of a double mutant.Theor Popul Biol 53: 199-215.

Christman JK, Sheikhnejad G, Marasco CJ, Sufrin RJ (1995) 5-Methyl-2-deoxycytidine in single stranded DNA can act in cis to signal de novo DNA methylation. Proc Natl Acad Sci USA 92: 7347–51.

Christophersen OA (2013) Why is there so much DHA in the brain, retina and testis? Possible implications for human reproduction and the survival of our species. In: De Meester F, Watson RR, Zibadi S, eds. Omega-6/3 Fatty Acids: Functions, Sustainability Strategies and Perspectives. New York, NY: Humana Press.pp 209-244.

Chu CG, Tan CT, Yu GT, Zhong S, Xu SS, Yan L (2011) A novel retrotransposon inserted in the Dominant Vrn-B1 allele confers spring growth habit in tetraploid wheat (Triticum turgidum L.). G3 (Bethesda) 1: 637–45.

Chua CC, Hamdy RC, Chua BH (1998) Upregulation of vascular endothelial growth factor by H2O2 in rat heart endothelial cells. Free Radic Biol Med 25: 891-7.

Chuang CS, Pai TW, Hu CH, Tzou WS, Dah-Tsyr Chang M, et al. (2011)Functional pathway mapping analysis for hypoxia-inducible factors.BMC Syst Biol 5 Suppl 1:S3.

Chuma S, Hosokawa M, Kitamura K, Kasai S, Fujioka M, et al. (2006) Tdrd1/Mtr-1, a tudor-related gene, is essential for male germ-cell differentiation and nuage/germinal granule formation in mice. Proc Natl Acad Sci USA 103: 15894–9.

Chuma S, Hosokawa M, Tanaka T, Nakatsuji N (2009) Ultrastructural characterization of spermatogenesis and its evolutionary conservation in the germline: Germinal granules in mammals. Mol Cell Endocrinol 306: 17–23.

Chun YJ, Collyer ML, Moloney KA, Nason JD (2007) Phenotypic plasticity of native vs. invasive purple loosestrife: a two-state multivariate approach. Ecology 88: 1499–512.

Chung CS (1962) Relative genetic loads due to lethal and detrimental genes in irradiated populations of Drosophila melanogaster. Genetics 47: 1489–504.

Chung MH, KasaiH, Nishimura S, Yu BP (1992) Protection of DNA damage by dietary restriction. Free Rad Biol Med 12: 523-5.

Churikov D, Zalenskaya IA, Zalensky AO (2004) Male germline-specific histones in mouse and man. Cytogenet Genome Res 105: 203-14.

Cianciolo JM, Norton RA (2006) The ecological distribution of reproductive mode in oribatid mites as related to biological complexity. Exp Appl Acarol 40:1–25.

Ciborowski KL, Consuegra S, García de Leániz C, Beaumont MA, Wang J, Jordan WC (2007) Rare and fleeting: an example of interspecific recombination in animal mitochondrial DNA. Biol Lett 3: 554–7.

Cicchillitti L, Di Stefano V, Isaia E, Crimaldi L, Fasanaro P, et al. (2012)Hypoxia-inducible factor 1-α induces miR-210 in normoxic differentiating myoblasts.J Biol Chem 287: 44761-71.

Ciechanover A (1994) The ubiquitin-proteasome proteolytic pathway. Cell 79: 13–21.

Cilensek ZM, Yehiely F, Kular RK, Deiss LP (2002) A member of the GAGE family of tumor antigens is an anti-apoptotic gene that confers resistance to Fas/CD95/APO-1, Interferon-gamma, taxol and gamma-irradiation. Cancer Biol Ther 1: 380-7.

Cimino F, Esposito F, Ammendola R, Russo T (1997)Gene regulation by reactive oxygen species.Curr Top Cell Regul 35: 123-48.

Cinquetti R, Dramis L (2003) Histological, histochemical and ultrastructural investigations of the testis of Padogobius martensi between annual breeding seasons. Journal of Fish Biology 63: 1402–28.

Ciofu O, Riis B, Pressler T, Poulsen HE, Hoiby N (2005) Occurrence of hypermutable Pseudomonas aeruginosa in cystic fibrosis patients is associated with the oxidative stress caused by chronic lung inflammation. Antimicrob Agents Chemother 49: 2276–82.

Ciota AT, Jia Y, Payne AF, Jerzak G, Davis LJ, et al. (2009) Experimental passage of St. Louis encephalitis virus in vivo in mosquitoes and chickens reveals evolutionarily significant virus characteristics. PLoS ONE 4: e7876.

Circu ML, Aw TY (2010)Reactive oxygen species, cellular redox systems, and apoptosis.Free Radic Biol Med 48: 749-62.

Cirz RT, Romesberg FE (2007) Controlling mutation: intervening in evolution as a therapeutic strategy. Crit Rev Biochem Mol Biol 42: 341–354.

Civetta A, Singh RS (1995) High divergence of reproductive tract proteins and their association with postzygotic reproductive isolation in Drosophila melanogaster and Drosophila virilis group species. J Mol Evol 41: 1085–95.

Civetta A, Singh RS (1999) Broad-sense sexual selection, sex gene pool evolution, and speciation. Genome 42: 1033–41.

Clanton TL (2007) Hypoxia-induced reactive oxygen species formation in skeletal muscle. J Appl Physiol 102: 2379–88.

Clanton W (1934) An unusual situation in the salamander Ambystoma jeffersonianum. Occas Pap Mus Zool Univ Mich 290: 1-14.

Clark AB, Cook ME, Tran HT, Gordenin DA, Resnick MA, Kunkel TA (1999) Functional analysis of human MutSalpha and MutSbeta complexes in yeast. Nucleic Acids Res 27: 736-42.

Clark AG (1987) Genetic correlations: the quantitative genetics of evolutionary constraints. In: Loeschcke V, ed. Genetic constraints on adaptive evolution. New York, NY: Springer. pp 25–45.

Clark AG, Feldman MW (1981)Density-dependent fertility selection in experimental populations of Drosophila melanogaster.Genetics 98:849-69.

Clark AG, Lyckegaard EM (1988) Natural selection with nuclear and cytoplasmic transmission. III. Joint analysis of segregation and mtDNA in Drosophila melanogaster. Genetics 118: 471-81.

Clark AG, Glanowski S, Nielsen R, Thomas PD, Kejariwal A, et al. (2003)Inferring nonneutral evolution from human-chimp-mouse orthologous gene trios.Science 302: 1960-3.

Clark AG, Eisen MB, Smith DR, Bergman CM, Oliver B, et al. (2007) Evolution of genes and genomes on the Drosophila phylogeny. Nature 450: 203–18.

Clark CJ, Sage EH (2008) A prototypic matricellular protein in the tumor microenvironment–where there’s SPARC, there’s fire. J Cell Biochem 104: 721–32.

Clark CW, Harvell CD (1992) Inducible defenses and the allocation of resources: a minimal model. Am Nat 139: 521–39.

Clark IE, Dodson MW, Jiang C, Cao JH, Huh JR, et al. (2006)Drosophila pink1 is required for mitochondrial function and interacts genetically with parkin.Nature 441: 1162-6.

Clark JG (1900) The origin, development and degeneration of the blood-vessels of the human ovary. Johns Hopkins Hosp Rep 9: 593-676.

Clark JM, Eddy EM (1975)Fine structural observations on the origin and associations of primordial germ cells of the mouse.Dev Biol 47: 136-55.

Clark MS, Denekamp NY, Thorne MA, Reinhardt R, Drungowski M, et al. (2012)Long-term survival of hydrated resting eggs from Brachionus plicatilis.PLoS One 7: e29365.

Clark NL, Swanson WJ (2005) Pervasive adaptive evolution in primate seminal proteins. PLoS Genet 1: e35.

Clark SCA, Sharp NP, Rowe L, Agrawal AF (2012) Relative effectiveness of mating success and sperm competition at eliminating deleterious mutations in Drosophila melanogaster. PLoS ONE 7: e37351.

Clark WR (1996) Sex and the Origins of Death. New York, NY: Oxford University Press.

Clarke A, Gaston KJ (2006) Climate, energy and diversity. Proc R Soc Lond B Biol Sci 273: 2257–66.

Clarke B (1979) Evolution of genetic diversity. Proc R Soc Lond B 205: 453–74.

Clarke D, Duarte E, Moya A, Elena S, Domingo E, Holland JJ (1993) Genetic bottlenecks and population passages cause profound fitness differences in RNA viruses. J Virol 67: 222–8.

Clarke DK, Duarte EA, Elena SF, Moya A, Domingo E, Holland J (1994)The red queen reigns in the kingdom of RNA viruses.Proc Natl Acad Sci USA 91: 4821-4.

Clarke E (2011) Plant individuality and multilevel selection theory. In: Sterelny K, Calcott B, eds. The major transitions in evolution revisited. Cambridge, MA: MIT Press. pp 227–51.

Clarke E (2012) Plant individuality: a solution to the demographer’s dilemma.Biol Philos 27: 321-61.

Clarke JR (1955) Influence of numbers on reproduction and survival in two experimental vole populations. Proc R Soc Lond B 144: 68-85.

Clarke SH, Claflin JL, Rudikoff S (1982) Polymorphism in immunoglobulin heavy chains suggesting gene conversion. Proc Natl Acad Sci USA 79: 3280–4.

ClaverysJP, Prudhomme M, Martin B (2006) Induction of competence regulons as a general response to stress in gram-positive bacteria. Annu Rev Microbiol 60: 451–75.

Claverys JP, Håvarstein LS (2007) Cannibalism and fratricide: mechanisms and raison d’être. Nat Rev Microbiol 5: 219–29.

Cleeter MW, Cooper JM, Darley-Usmar VM, Moncada S, Schapira AH (1994) Reversible inhibition of cytochrome c oxidase, the terminal enzyme of the mitochondrial respiratory chain, by nitric oxide. Implications for neurodegenerative diseases. FEBS Lett 345: 50–4.

Clegg EJ (1963) Studies on artificial cryptorchidism: morphological and quantitative changes in the Sertoli cells of the rat testis. J Endocrinol 26: 567–74.

Clegg JS (2005) Desiccation tolerance in encysted embryos of the animal extremophile, Artemia. Integr Com Biol 45: 715–24.

Clementi E, Brown GC, Foxwell N, Moncada S (1999) On the mechanism by which vascular endothelial cells regulate their oxygen consumption. Proc Natl Acad Sci USA 96: 1559–62.

Clermont Y (1962) Quantitative analysis of spermatogenesis of the rat: a revised model for the renewal of spermatogonia. Am J Anat 111: 111-29.

Clermont Y (1972) Kinetics of spermatogenesis in mammals: Seminiferous epithelium cycle and spermatogonial renewal. Physiol Rev 52: 198-236.

Clune J, Misevic D, Ofria C, Lenski RE, Elena SF, Sanjuán R (2008) Natural selection fails to optimize mutation rates for long-term adaptation on rugged fitness landscapes. PLoS Comput Biol 4: e1000187.

Clutton-Brock TH (2007) Sexual selection in males and females. Science 318: 1882–5.

Clutton-Brock TH, Vincent AC (1991) Sexual selection and the potential reproductive rates of males and females. Nature 351: 58–60.

Coates DJ (1992) Genetic consequences of a bottleneck and spatial genetic structure in the triggerplant, Stylidium coronifore (Stylidiaceae). Heredity 69: 512-20.

Cobb HG, Grefenstette JJ (1993) Genetic algorithms for tracking changing environments. In: Proceedings of the 5th international conference on genetic algorithms. Citeseer. pp 523–530.

Codelia VA, Cisternas P, Moreno RD (2008) Relevance of caspase activity during apoptosis in pubertal rat spermatogenesis. Mol Reprod Dev 75: 881-9.

Codelia VA, Cisterna M, Álvarez AR, Moreno RD (2010) p73 participates in male germ cells apoptosis induced by etoposide. Mol Hum Reprod 16: 734-42.

Coffin JH (1992) Genetic diversity and evolution of retroviruses. Curr Top Microbiol Immunol 176: 143-64.

Coffman CR (2003) Cell migration and programmed cell death of Drosophila germ cells. Ann NY Acad Sci 995: 117–26.

Cohen D (1966) Optimizing reproduction in a randomly varying environment. J Theor Biol 12: 119–29.

Cohen I, Knopf JA, Irihimovitch V, Shapira M (2005)A proposed mechanism for the inhibitory effects of oxidative stress on Rubisco assembly and its subunit expression.Plant Physiol 137: 738-46.

Cohen SE, Walker GC (2010) The transcription elongation factor NusA is required for stress-induced mutagenesis in Escherichia coli. Curr Biol 20: 80-5.

Cohen SM, Purilo DT, Ellwein LB (1991) Pivotal role of increased cell proliferation in human carcinogenesis. Mod Pathol 4: 371-82.

Colbourne JK, Pfrender ME, Gilbert D, Thomas WK, Tucker A, et al. (2011)The ecoresponsive genome of Daphnia pulex.Science 331: 555-61.

Colegrave N (2002) Sex releases the speed limit on evolution. Nature 420: 664–6.

Colegrave N (2012)The evolutionary success of sex.EMBO Rep 13: 774-8.

Colegrave N, Kotiaho JS, Tomkins JL (2002a) Mate choice or polyandry: reconciling genetic compatibility and good genes sexual selection. Evol Ecol Res 4: 911–7.

Colegrave N, Kaltz O, Bell G (2002b) The ecology and genetics of fitness in Chlamydomonas. VIII. The dynamics of adaptation to novel environments after a single episode of sex. Evolution 56: 14–21.

Colegrave N, Collins S (2008)Experimental evolution: experimental evolution and evolvability. Heredity 100: 464–70.

Collier S, Tassabehji M, Sinnott P, Strachan T (1993) A de novo pathological point mutation at the 21-hydroxylase locus: implications for gene conversion in the human genome. Nat Genet 3: 260–5.

Collin O, Bergh A, Damber JE, Widmark A (1993) Control of testicular vasomotion by testosterone and tubular factors in rats. J Reprod Fertil 97: 115–121.

Collin O, Bergh A (1996a) Leydig cells secrete factors which increase vascular permeability and endothelial cell proliferation. Int J Androl 19: 221–8.

Collin O, Bergh A (1996b) Acute stress influences testicular microcirculation in adult rats. Miniposters 9th European Testis Workshop, p F2.

Collin O, Damber JE, Bergh A (1996) Control of testicular vasomotion and blood flow by serotonin and mast cells products in rats. J Reprod Fertil 106: 17-22.

Collin O, Zupp JL, Setchell BP (2000)Testicular vasomotion in different mammals.Asian J Androl 2: 297-300.

Collins AM, Williams V, Evans JD (2004) Sperm storage and antioxidative enzyme expression in the honey bee, Apis mellifera. Insect Mol Biol 13: 141–6.

Collins AR (2004) The comet assay for DNA damage and repair: principles, applications, and limitations. Mol Biotechnol 26: 249–61.

Collu R, Gibb W, Ducharme JR (1984) Effects of stress on the gonadal function. J Endocrinol Invest 7: 529-37.

Coltman DW, Bowen WD, Wright JM (1998) Birth weight and neonatal survival of harbour seal pups are positively correlated with genetic variation measured by microsatellites. Proc R Soc B 265: 803–9.

Coltman DW, Pilkington JG, Smith JA, Pemberton JM (1999) Parasite-mediated selection against inbred Soay sheep in a free-living, island population. Evolution 53: 1259–67.

Comai L (2005)The advantages and disadvantages of being polyploid.Nat Rev Genet 6: 836-46.

Comai L, Tyagi AP, Winter K, Holmes-Davis R, Reynolds SH, et al. (2000) Phenotypic instability and rapid gene silencing in newly formed Arabidopsis allotetraploids. Plant Cell 12: 1551–68.

Comas D, Pääbo S, Bertranpetit J (1995) Heteroplasmy in the control region of human mitochondrial DNA. Genome Res 5: 89–90.

Comitato R, Esposito T, Cerbo G, Angelini F, Varriale B, Cardone A (2006) Impairment of spermatogenesis and enhancement of testicular germ cell apoptosis induced by exogenous all-trans-retinoic acid in adult lizard Podarcis sicula. J Exp Zool A Comp Exp Biol305: 288-98.

Commoner B, Townsend J, Pake GE (1954) Free radicals in biological materials. Nature 174: 689–91.

Congdon JD, Vitt LJ, Hadley NF (1978) Parental investment: comparative reproductive energetics in bisexual and unisexual lizards genus Cnemidophorus. Am Nat 112: 509–21.

Conley AB, Miller WJ, Jordan IK (2008) Human cis natural antisense transcripts initiated by transposable elements. Trends Genet 24: 53-6.

Conley AB, Piriyapongsa J, Jordan IK (2008) Retroviral promoters in the human genome. Bioinformatics 24: 1563–7.

Connallon T (2010) Genic capture, sex linkage, and the heritability of fitness. Am Nat 175, 564–76.

Connallon T, Knowles L (2005) Intergenomic conflict revealed by patterns of sex-biased gene expression. Trends Genet 21: 495-9.

Connallon T, Cox RM, Calsbeek R (2010) Fitness consequences of sex-specific selection. Evolution 64: 1671–82.

Connallon T, Clark AG (2012)A general population genetic framework for antagonistic selection that accounts for demography and recurrent mutation.Genetics 190: 1477-89.

Conrad DF, Keebler JE, DePristo MA, Lindsay SJ, Zhang Y, et al. (2011) Variation in genome-wide mutation rates within and between human families. Nat Genet 43: 712–4.

Constantini D, Coluzza C, Fanfani A, Dell’Omo G (2007) Effects of carotenoid supplementation on colour expression, oxidative stress and body mass in rehabilitated captive adult kestrels (Falco tinnunculus). J Comp Physiol B Biochem Syst Environ Physiol 177: 723–31.

Conte D, Narindrasorasak S, Sarkar B (1996) In vivo and in vitro iron-replaced zinc finger generates free radicals and causes DNA damage. J Biol Chem 271: 5125-30.

Conway de Macario E, Macario AJL (2000) Stressors, stress, and survival: Overview. Front Biosci 5: d780–6.

Cook PJ (1992) Phosphogenesis around the Proterozoic-Phanerozoic transition. J Geol Soc London 149: 615–20.

Cook PJ, Ju BG, Telese F, Wang X, Glass CK, Rosenfeld MG (2009) Tyrosine dephosphorylation of H2AX modulates apoptosis and survival decisions. Nature 458: 591–6.

Cooke BA, Abayasekara DRE, Choi MSK, Dirami G, Phipp LH, West AP (1992) The effect of stress-induced ligands on testosterone formation in Leydig cells. In: Sheppard KE, Boublik JH, Funder JW, eds. Stress and Reproduction. New York, NY: Raven Press. pp 135–144.

Cooke MS, Evans MD, Dizdaroglu M, Lunec J (2003) Oxidative DNA damage: mechanisms, mutation, and disease. FASEB J 17: 1195–214.

Cooke PS, Meisami E (1991) Early hypothyroidism in rats causes increased adult testis and reproductive organ size but does not change testosterone levels. Endocrinology 129: 237–43.

Cooke PS, Hess RA, Porcelli J, Meisami E (1991) Increased sperm production in adult rats after transient neonatal hypothyroidism. Endocrinology 129: 244–8.

Cooke PS, ZhaoYD, Bunick D (1994) Triiodothyronine inhibits proliferation and stimulates differentiation of cultured neonatal Sertoli cells: possible mechanism for increased adult testis weight and sperm production induced by neonatal goitrogen treatment. Biol Reprod 51: 1000–5.

Cooper CE, Mason MG, Nicholls P (2008)A dynamic model of nitric oxide inhibition of mitochondrial cytochrome c oxidase.Biochim Biophys Acta 1777: 867-76.

Cooper DN, Krawczak M (1990) The mutational spectrum of single base-pair substitutions causing human genetic disease: patterns and predictions. Hum Genet 85: 55–74.

Cooper DN, Mort M, Stenson PD, Ball EV, Chuzhanova NA (2010)Methylation-mediated deamination of 5-methylcytosine appears to give rise to mutations causing human inherited disease in CpNpG trinucleotides, as well as in CpG dinucleotides.Hum Genomics 4:406-10.

Cooper TF (2007) Recombination speeds adaptation by reducing competition between beneficial mutations in populations of Escherichia coli. PLoS Biol 5: e225.

Cooper TF, Lenski RE, Elena SF (2005) Parasites and mutational load: an experimental test of a pluralistic theory for the evolution of sex. Proc R Soc B Biol Sci 272: 311–7.

Cooper TF, Morby AP, Gunn A, Schneider D (2006) Effect of random and hub gene disruptions on environmental and mutational robustness in Escherichia coli. BMC Genomics 7: 237.

Cooper VS, Lenski RE (2000) The population genetics of ecological specialization in evolving Escherichia coli populations. Nature 407: 736-9.

Cooper WS (1984) Expected time to extinction and the concept of fundamental fitness. J Theor Biol 107: 603-29.

Cooper WS, Kaplan RH (1982) Adaptive ‘‘coin-flipping’’: a decision-theoretic examination of natural selection for random individual variation. J Theor Biol 94: 135-51.

Coors A, De Meester L (2008) Synergistic, antagonistic and additive effects of multiple stressors: predation threat, parasitism and pesticide exposure in Daphnia magna. J Appl Ecol 45: 1820-8.

Corda S, Laplace C, Vicaut E, Duranteau J (2001)Rapid reactive oxygen species production by mitochondria in endothelial cells exposed to tumor necrosis factor-alpha is mediated by ceramide.Am J Respir Cell Mol Biol 24: 762-8.

Corley LS, Moore AJ (1999) Fitness of alternative modes of reproduction: developmental constraints and the evolutionary maintenance of sex. Proc R Soc B Biol Sci 266: 471–6.

Corn PG (2008) Hypoxic regulation of miR-210: Shrinking targets expand HIF-1’ s influence. Cancer Biol Ther 7: 265–7.

Cornillon S, Foa C, Davoust J, Buonavista N, Gross JD, Golstein P (1994) Programmed cell death in Dictyostelium. J Cell Sci 107: 2691-704.

Corning PA (2005) Holistic Darwinism: Synergy, Cybernetics and the Bioeconomics of Evolution. Chicago, IL: University of Chicago Press.

Cornwallis CK, Birkhead TR (2007) Changes in sperm quality and numbers in response to experimental manipulation of male social status and female attractiveness. Am Nat 170: 758–70.

Coros CJ, Piazza CL, Chalamcharla VR, Smith D, Belfort M (2009) Global regulators orchestrate group II intron retromobility. Mol Cell 34: 250–6.

Corrie AM, Crozier RH, van Heeswijck R,Hoffmann AA (2002) Clonal reproduction and population genetic structure of grape phylloxera, Daktulosphaira vitifoliae, in Australia. Heredity 88: 203–11.

Corriero A, Desantis S, Bridges CR, Kime DE, Megalofonou P, et al. (2007) Germ cell proliferation and apoptosis during different phases of swordfish (Xiphias gladius L.) spermatogenetic cycle. J Fish Biol 70: 83-99.

Corriero A, Medina A, Mylonas CC, Bridges CR, Santamaria N, et al. (2009) Proliferation and apoptosis of male germ cells in captive Atlantic bluefin tuna (Thunnus thynnus L.) treated with gonadotropin-releasing hormone agonist (GnRHa). Anim Reprod Sci 116: 346-57.

Cortellino S, Xu J, Sannai M, Moore R, Caretti E, et al. (2011) Thymine DNA glycosylase is essential for active DNA demethylation by linked deamination-base excision repair. Cell 146: 67-79.

Cortopassi GA, Shibata D, Soong N-W, Arnheim N (1992) A pattern of accumulation of a somatic deletion of mitochondrial DNA in aging human tissues. Proc Natl Acad Sci U S A 89: 7370–4.

Corzett CH, Goodman MF, Finkel SE (2013) Competitive fitness during feast and famine: How SOS DNA polymerases influence physiology and evolution in Escherichia coli. Genetics 194: 409-20.

Cosmides LM, Tooby J (1981) Cytoplasmic inheritance and intragenomic conflict. J Theor Biol 81: 83–129.

Cossins A (1998) Cryptic clues revealed. Nature 396: 309-10.

Costantini D (2008) Oxidative stress in ecology and evolution: lessons from avian studies. Ecol Lett 11: 1238–51.

Costanzo V, Robertson K, Bibikova M, Kim E, Grieco D, et al. (2001) Mre11 protein complex prevents double-strand break accumulation during chromosomal DNA replication. Mol Cell 8: 137–47.

Costur P, Filiz S, Gonca S, Çulha M, Gülecen T, et al. (2012) Expression of inducible nitric oxide synthase (iNOS) in the azoospermic human testis. Andrologia 44 (Suppl. 1): 654–660.

Côté G, Perry G, Blier P, Bernatchez L (2007) The influence of gene-environment interactions on GHR and IGF-1 expression and their association with growth in brook charr, Salvelinus fontinalis (Mitchill). BMC Genetics 8: 87.

Cotgreave IA, Gerdes RG (1998) Recent trends in glutathione biochemistry—glutathione–protein interactions: a molecular link between oxidative stress and cell proliferation? Biochem Biophys Res Commun 242: 1–9.

Cotton S (2009) Condition-dependent mutation rates and sexual selection. J Evol Biol 22: 899–906.

Cotton S, Small J, Pomiankowski A (2006) Sexual selection and condition-dependent mate preferences. Curr Biol 16: R755–R765.

Couch RB, Cate TR, Douglas RG, Jr, Gerone PJ, Knight V (1966) Effect of route of inoculation on experimental respiratory viral disease in volunteers and evidence for airborne transmission. Bacteriol Rev 30: 517–29.

Coufal NG, Garcia-Perez JL, Peng GE, Marchetto MC, Muotri AR, et al. (2011) Ataxia telangiectasia mutated (ATM) modulates long interspersed element-1 (L1) retrotransposition in human neural stem cells. Proc Natl Acad Sci USA 108: 20382-7.

Coulombre JL, Russell ES (1954) Analysis of the pleiotropism at the W locus in the mouse. The effects of W and V substitutions upon post natal development of germ cells. J Exp Zool 126: 277-96.

Coulondre C,Miller JH, Farabaugh PJ, GilbertW (1978) Molecular basis of base substitution hotspots in Escherichia coli. Nature 274: 775–80.

Courcelle J, Khodursky A, Peter B, Brown PO, Hanawalt PC (2001) Comparative gene expression profiles following UV exposure in wild-type and SOS-deficient Escherichia coli. Genetics 158: 41–64.

Coussens M, Maresh JG, Yanagimachi R, Maeda G, Allsopp R (2008) Sirt1 deficiency attenuates spermatogenesis and germ cell function. PLoS ONE 3: e1571.

Coustau C, Chevillon C, ffrench-Constant R (2000) Resistance to xenobiotics and parasites: Can we count the cost? Trends Ecol Evol 15: 378–83.

Covarrubias L, Hernández-García D, Schnabel D, Salas-Vidal E, Castro-Obregón S (2008) Function of reactive oxygen species during animal development: passive or active? Dev Biol 320: 1-11.

Covello KL, Kehler J, Yu H, Gordan JD, Arsham AM (2006) HIF-2alpha regulates Oct-4: effects of hypoxia on stem cell function, embryonic development, and tumor growth. Genes Dev 20: 557–70.

Cowell IG, Aucott R, Mahadevaiah SK, Burgoyne PS, Huskisson N, et al. (2002) Heterochromatin, HP1 and methylation at lysine 9 of histone H3 in animals. Chromosoma 111: 22–36.

Cowen LE, Lindquist SL (2005) Hsp90 potentiates the rapid evolution of new traits: Drug resistance in diverse fungi. Science 309: 2185-9.

Cowles RB (1958) The evolutionary significance of the scrotum. Evolution 12: 417–8.

Cowles RB (1965) Hyperthermia, aspermia, mutation rates and evolution. Q Rev Biol 40: 341–367.

Cowles RB, Burleson GM (1945) The sterilizing effects of high temperature on the male germ plasm of the yucca night lizard, Xantusia vigilis. Am Nat 79: 417-35.

Cowperthwaite MC, Bull JJ, Meyers LA (2006) From bad to good: Fitness reversals and the ascent of deleterious mutations. PLoS Comput Biol 2: e141.

Cowperthwaite MC, Economo EP, Harcombe WR, Miller EL, Meyers LA (2008) The ascent of the abundant: How mutational networks constrain evolution. PLoS Comput Biol 4: e1000110.

Cox DN, Chao A, Lin H (2000) piwi encodes a nucleoplasmic factor whose activity modulates the number and division rate of germline stem cells. Development 127: 503–14.

Cox EC, Gibson TC (1974) Selection for high mutation rates in chemostats. Genetics 77: 169–84.

Cox RT, Spradling AC (2003) A Balbiani body and the fusome mediate mitochondrial inheritance during Drosophila oogenesis. Development 130: 1579–90.

Cox RT, Spradling AC (2006) Milton controls the early acquisition of mitochondria by Drosophila oocytes. Development 133: 3371–7.

Coyle S, Kroll E (2008) Starvation induces genomic rearrangements and starvation-resilient phenotypes in yeast. Mol Biol Evol 25: 310–8.

Crabtree HG (1929) Observations on the carbohydrate metabolism of tumours. Biochem J 23: 536-45.

Craig EA (1985) The heat-shock response. CRC Crit Rev Biochem 18: 239–80.

Craig NL (1997) Target site selection in transposition. Annu Rev Biochem 66: 437-74.

Crean AJ, Marshall DJ (2009) Coping with environmental uncertainty: dynamic bet hedging as a maternal effect. Philos Trans R Soc Lond B Biol Sci 364: 1087-96.

Crean AJ, Dwyer JM, Marshall DJ (2012) Fertilization is not a new beginning: the relationship between sperm longevity and offspring performance. PLoS ONE 7: e49167.

Cree LM, Samuels DC, de Sousa Lopes SC, Rajasimha HK, Wonnapinij P, et al. (2008) A reduction of mitochondrial DNA molecules during embryogenesis explains the rapid segregation of genotypes. Nat Genet 40: 249–54.

Crenshaw JW (1965) Radiation-induced increases in fitness in the flour beetle Tribolium confusum. Science 149:426-7.

Crespi B (2010) The origins and evolution of genetic disease risk in modern humans. Ann NY Acad Sci 1206: 80–109.

Crespo R, Shivaprasad HL (2010) Decrease of fertility in a broiler breeder flock due to testicular atrophy. Avian Dis 54:142-5.

Crews D (1996) Temperature-dependent sex determination: the interplay of steroid hormones and temperature. Zool Sci 13: 1–13.

Crews D (2003) Sex determination: where environment and genetics meet. Evol Dev 5: 50–5.

Crews D, Bergeron JM (1994) Role of reductase and aromatase in sex determination in the red-eared slider (Trachemys scripta), a turtle with temperature-dependent sex determination. J Endocrinol 143: 279–89.

Crews D, Fleming A, Willingham E, Baldwin R, Skipper J (2001) Role of steroidogenic factor I and aromatase in temperature-dependent sex determination in the red-eared slider turtle. J Exp Zool 290: 597–606.

Crews D, Gore AC, Hsu TS, Dangleben NL, Spinetta M, et al. (2007) Transgenerational epigenetic imprints on mate preference. Proc Natl Acad Sci USA 104: 5942–6.

CrickF (1958) On protein synthesis. Symp Soc Exp Biol 12: 138–67.

Crick F (1970) Central dogma of molecular biology. Nature 227: 561-3.

Crighton D, Wilkinson S, O'Prey J, Syed N, Smith P, et al. (2006) DRAM, a p53-induced modulator of autophagy, is critical for apoptosis. Cell 126: 121–134.

Crispo E (2007) The Baldwin effect and genetic assimilation: revisiting two mechanisms of evolutionary change mediated by phenotypic plasticity. Evolution 61: 2469-79.

Critchlow SE, Jackson SP (1998) DNA end-joining: from yeast to man. Trends Biochem Sci 23: 394–8.

Critchlow V, Liebelt A, Bar-sela M, Mountcastle W, Lipscomb HS (1963) Sex differences in resting pituitary-adrenal function in the rat. Am J Physiol 205: 807–15.

Crnokrak P, Roff DA (1999) Inbreeding depression in the wild. Heredity 83: 260–70.

Crombach A, Hogeweg P (2008) Evolution of evolvability in gene regulatory networks. PLoS Comput Biol 4:e1000112.

Cromie GA, Connelly JC, Leach DR (2001) Recombination at double-strand breaks and DNA ends: conserved mechanisms from phage to humans. Mol Cell 8: 1163-74.

Crone EE (2001) Is survivorship a better fitness surrogate than fecundity? Evolution 55: 2611–4.

Cropley JE, Suter CM, Beckman KB, Martin DI (2006) Germ-line epigenetic modification of the murine A(vy) allele by nutritional supplementation. Proc Natl Acad Sci USA 103: 17308-12.

Crosby ME, Kulshreshtha R, Ivan M, Glazer PM (2009) MicroRNA regulation of DNA repair gene expression in hypoxic stress. Cancer Res 69: 1221-9.

Cross BA, Silver IA (1962) Neurovascular control of oxygen tension in the testis and epididymis. J Reprod Fertil 3: 377–95.

Crotty S, Cameron CE, Andino R (2001) RNA virus error catastrophe: direct molecular test by using ribavirin. Proc Natl Acad Sci USA 98: 6895–900.

Crow JF (1957) Genetics of insecticide resistance to chemicals. Annu Rev Entomol 2: 227–46.

Crow JF (1958) Some possibilities for measuring selection intensities in man. Hum Biol 30: 1-13.

Crow JF (1970) Mathematical Topics in Population Genetics (ed. Kojima, K. I.) (Springer, Heidelberg, 1970).

Crow JF (1987) Population genetics history: a personal view. Annu Rev Genet 21:1-22.

Crow JF (1993a) Mutation, mean fitness, and genetic load. Oxford Surv Evol Biol 9: 3–42.

Crow JF (1993b) How much do we know about spontaneous human mutation rates? Environ Mol Mutagen 21: 122–9.

Crow JF (1997a) Molecular evolution—who is in the driver’s seat? Nat Genet 17: 129–30.

Crow JF (1997b) The high spontaneous mutation rate: is it a health risk? Proc Natl Acad Sci USA 94: 8380–6.

Crow JF (1999) The odds of losing at genetic roulette. Nature 397: 293-4.

Crow JF (2000) The origins patterns and implications of human spontaneous mutation. Nat Rev Genet 1: 40–7.

Crow JF (2006) Age and sex effects on human mutation rates: An old problem with new complexities. J Radiat Res (Tokyo) 47: B75–82.

Crow JF, Morton NE (1955) Measurement of gene frequency drift in small populations. Evolution 9: 202–14.

Crow JF, Abrahamson S (1965) Genetic effects of ionizing radiation. In: Etter LE, ed. The science of ionizing radiation. Springfield, IL: Charles C Thomas Publ. pp 263-87.

Crow JF, Kimura M (1965) Evolution in sexual and asexual populations. Am Nat 99: 439–50.

Crow JF, Kimura M (1970) An introduction to population genetics theory. Minneapolis (MN): Burgess Publishing Company.

Crutsinger GM, Collins MD, Fordyce JA, Gompert Z, Nice CC, Sanders NJ (2006) Plant genotypic diversity predicts community structure and governs an ecosystem process. Science 313: 966-8.

Cruzan MB, Barrett SCH (1993) Contribution of cryptic incompatibility to the mating system of Eichhornia paniculata (Pontederiaceae). Evolution 47:925–34.

Csermely P, Kahn CR (1991) The 90 kD heat-shock protein (Hsp-90) possesses an ATP binding site and autophosphorylation activity. J Biol Chem 266: 4943-50.

Csermely P, Kajtár J, Hollósi M, Jalsovszky G, Holly S, et al. (1993) ATP induces a conformational change of the 90-kD heat shock protein (hsp90). J Biol Chem 268: 1901-7.

Csordás G, Hajnóczky G (2009) SR/ER-mitochondrial local communication: calcium and ROS.Biochim Biophys Acta 1787: 1352-62.

Cubas P, Vincent C, Coen E (1999) An epigenetic mutation responsible for natural variation in floral symmetry. Nature 401: 157–61.

Cubo J (2003) Evidence for speciational change in the evolution of ratites (Aves: Palaeognathae). Biol J Linnean Soc 80: 99–106.

Cuellar O (1994) Biogeography of parthenogenetic animals. Biogeographica 70: 1–13.

Cuevas JM, Elena SF, Moya A (2002) Molecular basis of adaptive convergence in experimental populations of RNA viruses. Genetics 162: 533-42.

Cuevas JM, Domingo-Calap P, Pereira-Gómez M, Sanjuán R (2009) Experimental evolution and population genetics of RNA viruses. Open Evol J 3: 9–16.

Cui L, Wall PK, Leebens-Mack JH, Lindsay BG, Soltis DE, et al. (2006) Widespread genome duplications throughout the history of flowering plants. Genome Res 16: 738-49.

Cui W, Li B, Bai Y, Miao X, Chen Q, et al. (2013) Potential role for Nrf2 activation in the therapeutic effect of MG132 on diabetic nephropathy in OVE26 diabetic mice. Am J Physiol Endocrinol Metab 304: E87–E99.

Cui Y, Chen RS, Wong WH (2000) The coevolution of cell senescence and diploid sexual reproduction in unicellular organisms. Proc Natl Acad Sci USA 97: 3330-5.

Cullen BR (2006) Viruses and microRNAs. Nat Genet 38(Suppl): S25–30.

Cullis CA (1987) The generation of somatic and heritable variation in response to stress. Am Nat 130: S62-S73.

Cullis CA (2005) Mechanisms and control of rapid genomic changes in flax. Ann Bot (Lond) 95: 201–6.

Cumming D, Quigley ME, Yen SSC (1983) Acute suppression of circulating testosterone levels by cortisol in men. J Clin Endocrinol Metab 57: 671–3.

Cumming GS (2002) Comparing climate and vegetation as limiting factors for species ranges of African ticks. Ecology 83: 255–68.

Cummins JM, Yanagimachi R (1982) Sperm–egg ratios and the site of the acrosome reaction during in vivo fertilization in the hamster. Gamete Res 5: 239–56.

Cunningham EJA, Russell AF (2000) Egg investment is influenced by male attractiveness in the mallard. Nature 404: 74–7.

Curley JP, Mashoodh R (2010) Parent-of-origin and trans-generational germline influences on behavioral development: the interacting roles of mothers, fathers, and grandparents. Dev Psychobiol 52: 312-30.

Currie DJ (1991) Energy and large-scale patterns of animal and plant-species richness. Am Nat 137: 27–49.

Currie DJ, Mittelbach GG, Cornell HV, Field R, Guégan JF, et al. (2004) Predictions and tests of climate-based hypotheses of broad-scale variation in taxonomic richness. Ecol Lett 7: 1121-34.

Curtis WC (1902) The life history, the normal fission and the reproductive organs of Planaria maculata. Proc Boston Soc Nat Hist 30: 515–59.

Cutforth T, Rubin GM (1994) Mutations in Hsp83 and cdc37 impair signaling by the sevenless receptor tyrosine kinase in Drosophila. Cell 77: 1027–36.

Cutler DJ (2000) Estimating divergence times in the presence of an overdispersed molecular clock. Mol Biol Evol 17: 1647-60.

Cutler RG (1991) Human longevity and aging: possible role of reactive oxygen species. Ann NY Acad Sci 621: 1-28.

Cuzin F, Rassoulzadegan M (2010) Non-Mendelian epigenetic heredity: Gametic RNAs as epigenetic regulators and transgenerational signals. Essays Biochem 48: 101–106.

Cyr AR, Domann FE (2011) The redox basis of epigenetic modifications: from mechanisms to functional consequences. Antioxid Redox Signal 15: 551-89.

Czarna M, Jarmuszkiewicz W (2005) Activation of alternative oxidase and uncoupling protein lowers hydrogen peroxide formation in amoeba Acanthamoeba castellanii mitochondria. FEBS Lett 579: 3136-40.

Czech B, Hannon GJ (2010) Small RNA sorting: matchmaking for Argonautes. Nat Rev Genet 12: 19–31.

Czolowska R (1969) Observations on the origin of the “germinal cytoplasm” in Xenopus laevis. J Embryol Exp Morphol 22: 229–51.

Dabiké M, Preller A (1999) Cytoarchitecture of Caudiverbera caudiverbera stage VI oocytes: a light and electron microscope study. Anat Embryol (Berl) 199: 489-97.

Daboussi MJ, Capy P (2003) Transposable elements in filamentous fungi. Ann Rev Microbiol 57: 275-99.

D’Abramo LR (1980) Ingestion rate decrease as the stimulus for sexuality in populations of Moina macrocopa. LimnolOceanogr 25: 422–9.

D'Abrizio P, Baldini E, Russo PF, Biordi L, Graziano FM, et al. (2004) Ontogenesis and cell specific localization of Fas ligand expression in the rat testis. Int J Androl 27: 304-10.

Dacks J, Roger AJ (1999) The first sexual lineage and the relevance of facultative sex. J Mol Evol 48:779-83.

Dakouane Giudicelli M, Serazin V, Rouillac Le Sciellour C, Albert M, Selva J, Giudicelli Y (2008) Increased achondroplasia mutation frequency with advanced age and evidence for G1138A mosaicism in human testis biopsies. Fertil Steril 89: 1651–1656.

da Cunha AB, Burla H,Dobzhansky T (1950) Adaptive chromosome polymorphism in Drosophila willistoni. Evolution 4: 212–35.

Dadhich RK, Real FM, Zurita F, Barrionuevo FJ, Burgos M, Jiménez R (2010) Role of apoptosis and cell proliferation in the testicular dynamics of seasonal breeding mammals: a study in the Iberian mole, Talpa occidentalis. Biol Reprod 83: 83-91.

Dagda RK, Chu CT (2009) Mitochondrial quality control: insights on how Parkinson's disease related genes PINK1, parkin, and Omi/HtrA2 interact to maintain mitochondrial homeostasis. J Bioenerg Biomembr 41: 473-9.

Dagsgaard C, Taylor LE, O’brien KM, Poyton RO (2001) Effects of anoxia and the mitochondrion on expression of aerobic nuclear COX genes in yeast: evidence for a signaling pathway from the mitochondrial genome to the nucleus. J Biol Chem 276: 7593–601.

Dahl E (1951) On the relation between summer temperature and the distribution of alpine vascular plants in the lowlands of Fennoscandia. Oikos 3: 22–52.

Dahlgren BT (1979) The effects of population density on fecundity and fertility in the guppy, Poecilia reticulata (Peters). J Fish Biol 15:71-91.

Dai J, Xie W, Brady TL, Gao J, Voytas DF (2007) Phosphorylation regulates integration of the yeast Ty5 retrotransposon into heterochromatin. Mol Cell 27: 289–99.

Dalakouras A, Wassenegger M (2013) Revisiting RNA-directed DNA methylation. RNA Biol 10: 453-5.

Dalbiès-Tran R, Mermillod P (2003) Use of heterologous complementary DNA array screening to analyze bovine oocyte transcriptome and its evolution during in vitro maturation. Biol Reprod 68: 252–61.

D’Alessandro EK, Sponaugle S, Cowen RK (2013) Selective mortality during the larval and juvenile stages of snappers (Lutjanidae) and great barracuda Sphyraena barracuda. Mar Ecol Prog Ser 474: 227-42.

D'Alessio A, Riccioli A, Lauretti P, Padula F, Muciaccia B, et al. (2001) Testicular FasL is expressed by sperm cells. Proc Natl Acad Sci USA 98: 3316–21.

Dallinger WD (1887) The president's address. J R Microsc Soc 7: 184-99.

Dalton TP, Shertzer HG, Puga A (1999) Regulation of gene expression by reactive oxygen. Annu Rev Pharmacol Toxicol 39: 67-101.

Daly MJ, Gaidamakova EK, Matrosova VY, Vasilenko A, Zhai M, et al. (2007) Protein oxidation implicated as the primary determinant of bacterial radioresistance. PLoS Biol 5: e92.

Damber J, Lindahl O, Selstam G, Tenland T (1982) Testicular blood flow measured with a laser Doppler flowmeter: acute effects of catecholamines. Acta Physiol Scand 115: 209–15.

Damber JE, Lindahl O, Selstam G, Tenland T (1983) Rhythmical oscillations in rat testicular microcirculation as recorded by laser Doppler flowmetry. Acta Physiol Scand 118:117–123.

Damber JE, Bergh A, Widmark KA (1987) Testicular blood flow and microcirculation in rats after treatment with ethane dimethyl sulfonate. Biol Reprod 37: 1291–6.

Damber JE, Bergh A, Widmark A (1990) Age-related differences in testicular microcirculation. Int J Androl 13: 197–206.

Damber JE, Bergh A (1992) Testicular microcirculation—a forgotten essential in andrology? Int J Androl 15: 285–92.

Damber JE, Maddocks S, Widmark A, Bergh A (1992) Testicular blood flow and vasomotion can be maintained by testosterone in Leydig cell-depleted rats. Int J Androl 15: 385–93.

Damuth J (1981) Population density and body size in mammals. Nature 290: 699–700.

Damuth J (1987) Interspecific allometry of population density in mammals and other animals: the independence of body mass and population energy-use. Biol J Linn Soc 31: 193–246.

Danchin É, Charmantier A, Champagne FA, Mesoudi A, Pujol B, Blanchet S (2011a) Beyond DNA: integrating inclusive inheritance into an extended theory of evolution. Nat Rev Genet 12: 475-86.

Danchin EGJ, Flot JF, Perfus-Barbeoch L, Van Doninck K (2011b) Genomic perspectives on the long-term absence of sexual reproduction in animals. In: Pontarotti P, ed. Evolutionary Biology: Concepts, Biodiversity, Macroevolution and Genome Evolution. Berlin, Germany: Springer. pp 223-242.

Dandona P, Mohanty P, Ghanim H, Aljada A, Browne R, et al. (2001) The suppressive effect of dietary restriction and weight loss in the obese on the generation of reactive oxygen species by leukocytes, lipid peroxidation, and protein carbonylation. J Clin Endocrinol Metab 86: 355-62.

Danforth BN (1999) Emergence dynamics and bet hedging in a desert bee, Perdita portalis. Proc Natl Acad Sci USA 266: 1985–94.

Dang CV (2007) The interplay between MYC and HIF in the Warburg effect. Ernst Schering Found Symp Proc 4: 35-53.

Dang CV, O'Donnell KA, Zeller KI, Nguyen T, Osthus RC, Li F (2006) The c-Myc target gene network. Semin Cancer Biol 16: 253-64.

Dang CV, Kim JW, Gao P, Yustein J (2008) The interplay between MYC and HIF in cancer. Nat Rev Cancer 8: 51-6.

Daniels SB, Peterson KR, Strausbaugh LD, Kidwell MG, Chovnick A (1990) Evidence for horizontal transmission of the P transposable element between Drosophila species. Genetics 124: 339–55.

Danjoh I, Fujiyama A (1999) Ras-mediated signaling pathway regulates the expression of a low-molecular-weight heat-shock protein in fission yeast. Gene 236: 347-52.

Danner DB, Deich RA, Sisco KL, Smith HO (1980) An eleven-base-pair sequence determines the specificity of DNA uptake in Haemophilus transformation. Gene 11: 311-8.

Danner DB, Smith HO, Narang SA (1982) Construction of DNA recognition sites active in Haemophilus transformation. Proc Natl Acad Sci USA 79: 2393-7.

Dantas APV, Franco MP, Silva-Antonialli MM, Tostes RCA, Fortes D, et al. (2004) Gender differences in superoxide generation in microvessels of hypertensive rats: role of NAD(P)H-oxidase. Cardiovasc Res 61: 22–9.

D'Apice MR, Tenconi R, Mammi I, van den Ende J, Novelli G (2004) Paternal origin of LMNA mutations in Hutchinson-Gilford progeria. Clin Genet 65: 52–4.

Darboux I, Charles JF, Pauchet Y, Warot S, Pauron D (2007) Transposon-mediated resistance to Bacillus sphaericus in a field-evolved population of Culex pipiens (Diptera: Culicidae). Cell Microbiol 9: 2022–9.

Darevsky IS, Kupriyanova LA, Uzzell TM (1985) Parthenogenesis in reptiles. In: Gans C, Billett F (eds) Biology of the Reptilia. John Wiley and Sons Inc, New York, pp. 412–526

Darwin CR (1859, 6th edn. 1872) On the origin of species. London, UK: John Murray.

Darwin C (1871) The descent of man, and selection in relation to sex. London, UK: John Murray.

Darwin F, Seward AC, eds. (1903) More Letters of Charles Darwin. Volume I. London, UK: John Murray.

Darzynkiewicz Z (2010) Another “Janus paradox” of p53: induction of cell senescence versus quiescence. Aging (Albany NY) 2: 329-30.

Das J (2006) The role of mitochondrial respiration in physiological and evolutionary adaptation. Bioessays 28: 890-901.

Das K, Chainy GB (2004) Thyroid hormone influences antioxidant defense system in adult rat brain. Neurochem Res 29: 1755–66.

Daskalos A, Nikolaidis G, Xinarianos G, Savvari P, Cassidy A, et al. (2009) Hypomethylation of retrotransposable elements correlates with genomic instability in non-small cell lung cancer. Int J Cancer 124: 81-7.

Da Sylva TR, Gordon CS, Wu GE (2009) A genetic approach to quantifying human in vivo mutation frequency uncovers transcription level effects. Mutat Res 670: 68–73.

Datta A, Jinks-Robertson S (1995) Association of increased spontaneous mutation rates with high levels of transcription in yeast. Science 268: 1616–9.

David JR, Arens MF, Cohet Y (1971) Stérilité mâle à haute température chez Drosophila melanogaster: nature, progressivité et réversabilité des effets de la chaleur. CR Acad Sci Paris 272: 1007-10.

David JR, Araripe LO, Chakir M, Legout H, Lemos B, et al. (2005) Male sterility at extreme temperatures: a significant but neglected phenomenon for understanding Drosophila climatic adaptations. J Evol Biol 18: 838–46.

Davidsen T, Rødland EA, Lagesen K, Seeberg E, Rognes T, Tønjum T (2004) Biased distribution of DNA uptake sequences towards genome maintenance genes. Nucleic Acids Res 32: 1050-8.

Davidsen T, Amundsen EK, Rødland EA, Tønjum T (2007) DNA repair profiles of disease-associated isolates of Neisseria meningitidis. FEMS Immunol Med Microbiol 49: 243-51.

Davidson CJ, Surette MG (2008) Individuality in bacteria. Annu Rev Genet 42: 253-68.

Davie JR, Spencer VA (1999) Control of histone modifications. J Cell Biochem Suppl 32–33: 141–8.

Davies CM, Guilak F, Weinberg JB, Fermor B (2008) Reactive nitrogen and oxygen species in interleukin-1-mediated DNA damage associated with osteoarthritis. Osteoarthritis Cartilage 16: 624-30.

Davies E, Peters AD, Keightley PD (1999) High frequency of cryptic deleterious mutations in Caenorhabitis elegans. Science 285: 1748–51.

Davies KJ (1995) Oxidative stress: the paradox of aerobic life. Biochem Soc Symp 61: 1-31.

Davies TJ, Savolainen V, Chase MW, Moat J, Barraclough TG (2004) Environmental energy and evolutionary rates in flowering plants. Proc Biol Sci 271: 2195–200.

Davis BD (1989) Transcriptional bias: a non-Lamarckian mechanism for substrate-induced mutations. Proc Natl Acad Sci USA 86: 5005-9.

Davis CT, Beasley DWC, Guzman H, Raj P, D'Anton M, et al. (2003) Genetic variation among temporally and geographically distinct West Nile virus isolates, United States, 2001, 2002. Emerg Infect Dis 9: 1423–9.

Davis CT, Ebel GD, Lanciotti RS, Brault AC, Guzman H, et al. (2005) Phylogenetic analysis of North American West Nile virus isolates, 2001–2004: Evidence for the emergence of a dominant genotype. Virology 342: 252–65.

Davis JC, Brandman O, Petrov DA (2005) Protein evolution in the context of Drosophila development. J Mol Evol 60: 774-85.

Davis L, Smith GR (2001) Meiotic recombination and chromosome segregation in Schizosaccharomyces pombe. Proc Natl Acad Sci USA 98: 8395–402.

Davis LE, Schreck CB (1997) The energetic response to handling stress in juvenile coho salmon. T Am Fish Soc 126: 248–58.

Davis TL, Trasler JM, Moss SB, Yang GJ, Bartolomei MS (1999) Acquisition of the H19 methylation imprint occurs differentially on the parental alleles during spermatogenesis. Genomics 58: 18–28.

Davis TL, Yang GJ, McCarrey JR, Bartolomei MS (2000) The H19 methylation imprint is erased and re-established differentially on the parental alleles during male germ cell development. Hum Mol Genet 9: 2885–94.

Dawkins R (1976) The selfish gene. Oxford, UK: Oxford University Press.

Dawkins R (1982) The extended phenotype. Oxford, UK: Oxford University Press.

Dawley RM, Schultz RJ, Goddard KA (1987) Clonal reproduction and polyploidy in unisexual hybrids of Phoxinus eos and Phoxinus neogaeus (Pisces: Cyprinidae). Copeia 1987: 275–83.

Dawley RM (1989) An introduction to unisexual vertebrates. In: Dawley RM, Bogart JP, eds. Evolution and ecology of unisexual vertebrates. Albany, NY: New York State Museum Bulletin. pp 1-18.

Dawley RM, Bogart JP, eds. (1989) Evolution and ecology of unisexual vertebrates. New York, NY: New York State Museum.

Dawood MM, Strickberger MW (1969) Effect of larval interaction on viability in Drosophila melanogaster. III. Effects of biotic residues. Genetics 63: 213–20.

Dawson A (1998) Comparative reproductive physiology of nonmammalian species. Pure Appl Chem 70: 1657-69.

Dawson A, King VM, Bentley GE, Ball GF (2001) Photoperiodic control of seasonality in birds. J Biol Rhythms 16: 365–80.

Dawson KJ (1998) Evolutionarily stable mutation rates. J Theor Biol 194: 143–57.

Daxinger L, Whitelaw E (2010) Transgenerational epigenetic inheritance: more questions than answers. Genome Res 20: 1623-8.

Daxinger L, Whitelaw E (2012)Understanding transgenerational epigenetic inheritance via the gametes in mammals.Nat Rev Genet 13: 153-62.

Day JR, Laping NJ, Lampert-Etchells M, Brown SA, O'Callaghan JP, et al. (1993) Gonadal steroids regulate the expression of glial fibrillary acidic protein in the adult male rat hippocampus. Neuroscience 55: 435-43.

Day T, Pritchard J, Schluter D (1994) A comparison of two sticklebacks. Evolution 48: 1723–34.

Day T, Bonduriansky R (2011) A unified approach to the evolutionary consequences of genetic and nongenetic inheritance. Am Nat 178: E18–E36.

Dayhoff M, Schwartz R, Orcutt B (1978) A model of evolutionary change in proteins. In: Dayhoff M, ed. Atlas of Protein Sequence and Structure. Silver Spring, MD: National Biomedical Research Foundation. pp 345–52.

De SK, Chen HL, Pace JL, Hunt JS, Terranova PF, Enders GC (1993) Expression of tumor necrosis factor-α in mouse spermatogenic cells. Endocrinology 133: 389–96.

Deaton AM, Webb S, Kerr AR, Illingworth RS, Guy J, et al. (2011) Cell type-specific DNA methylation at intragenic CpG islands in the immune system. Genome Res 21: 1074-86.

Deaton AM, Bird A (2011) CpG islands and the regulation of transcription. Genes Dev 25: 1010-22.

Debat V, Milton CC, Rutherford S, Klingenberg CP, Hoffmann AA (2006) Hsp90 and the quantitative variation of wing shape in Drosophila melanogaster. Evolution 60: 2529–38.

de Beco S, Ziosi M, Johnston LA (2012) New frontiers in cell competition. Dev Dyn 241: 831-41.

de Boer JG, Ripley LS (1984) Demonstration of the production of frameshift and base-substitution mutations by quasipalindromic DNA sequences. Proc Natl Acad Sci USA 81: 5528–31.

de Boer P, Ramos L, de Vries M, Gochhait S (2010) Memoirs of an insult: sperm as a possible source of transgenerational epimutations and genetic instability. Mol Hum Reprod 16: 48-56.

De Boer RJ, Freitas AA, Perelson AS (2001) Resource competition determines selection of B cell repertoires. J Theor Biol 212: 333-43.

De Bont R, van Larebeke N (2004) Endogenous DNA damage in humans: a review of quantitative data. Mutagenesis 19: 169–85.

Debora BN, Vidales LE, Ramírez R, Ramírez M, Robleto EA, et al. (2011) Mismatch repair modulation of MutY activity drives Bacillus subtilis stationary-phase mutagenesis. J Bacteriol 193: 236-45.

De Bosscher K, Vanden Berghe W, Haegeman G (2006) Cross-talk between nuclear receptors and nuclear factor kappaB. Oncogene 25: 6868-86.

Decaestecker E, Gaba S, Raeymaekers JA, Stoks R, Van Kerckhoven L, et al. (2007) Host-parasite 'Red Queen' dynamics archived in pond sediment. Nature 450: 870-3.

De Cesaris P, Starace D, Starace G, Filippini A, Stefanini M, Ziparo E (1999) Activation of Jun N-terminal kinase/stress-activated protein kinase pathway by tumor necrosis factor α leads to intercellular adhesion molecule-1 expression. J Biol Chem 274: 28978–82.

Deckwerth TL, Johnson EM Jr (1993) Temporal analysis of events associated with programmed cell death (apoptosis) of sympathetic neurons deprived of nerve growth factor. J Cell Biol 123: 1207–22.

Dedryver CA, Le Gallic JF, Gauthier JP, Simon JC (1998) Life cycle of the cereal aphid Sitobion avenae F.: polymorphism and comparison of life history traits associated with sexuality. Ecol Entomol 23: 123–32.

Dedryver CA, Hullé M, Le Gallic JF, Caillaud MC, Simon JC (2001) Coexistence in space and time of sexual and asexual populations of the cereal aphid Sitobion avenae. Oecologia 128: 379–88.

Deem A, Keszthelyi A, Blackgrove T, Vayl A, Coffey B, et al. (2011) Break-induced replication is highly inaccurate. PLoS Biol 9: e1000594.

Dees ND, Bahar S (2010) Mutation size optimizes speciation in an evolutionary model. PLoS ONE 5: e11952.

De Ferrari L, Aitken S (2006) Mining housekeeping genes with a Naive Bayes classifier. BMC Genomics 7: 277.

Defoort EN, Kim PM, Winn LM (2006) Valproic acid increases conservative homologous recombination frequency and reactive oxygen species formation: a potential mechanism for valproic acid-induced neural tube defects. Mol Pharmacol 69: 1304–10.

De França LR, Hess RA, Cooke PS, Russell LD (1995) Neonatal hypothyroidism causes delayed Sertoli cell maturation in rats treated with propylthiouracil: evidence that the Sertoli cell controls testis growth. Anat Rec 242: 57-69.

De Gendt K, Swinnen JV, Saunders PT, Schoonjans L, Dewerchin M, et al. (2004) A Sertoli cell-selective knockout of the androgen receptor causes spermatogenic arrest in meiosis. Proc Natl Acad Sci USA 101: 1327–32.

de Grey ADNJ (1999) The Mitochondrial Free Radical Theory of Aging. Georgetown, TX: Landes.

Degtyarev SV, Grishanin AK, Belyakin SN, Rubtsov NB, Zhimulev IF, Akifyev AP (2002) DNA sequences eliminated during chromatin diminution from somatic cell chromosomes of Cyclops kolensis. Dokl Biochem Biophys 384: 148–51.

de Haan JB, Wolvetang EJ, Cristiano F, Iannello R, Bladier C, et al. (1997) Reactive oxygen species and their contribution to pathology in Down syndrome. Adv Pharmacol 38: 379-402.

de Haan JB, Crack PJ, Flentjar N, Iannello RC, Hertzog PJ, Kola I (2003) An imbalance in antioxidant defense affects cellular function: the pathophysiological consequences of a reduction in antioxidant defense in the glutathione peroxidase-1 (Gpx1) knockout mouse. Redox Rep 8: 69-79.

Deininger PL, Moran JV, Batzer MA, Kazazian HH Jr (2003) Mobile elements and mammalian genome evolution. Curr Opin Genet Dev 13: 651–8.

De Iuliis GN, Wingate JK, Koppers AJ, McLaughlin EA, Aitken RJ (2006) Definitive evidence for the nonmitochondrial production of superoxide anion by human spermatozoa. J Clin Endocrinol Metab 91: 1968–75.

De Iuliis GN, Thomson LK, Mitchell LA, Finnie JM, Koppers AJ, et al. (2009) DNA damage in human spermatozoa is highly correlated with the efficiency of chromatin remodeling and the formation of 8-hydroxy-2’-deoxyguanosine, a marker of oxidative stress. Biol Reprod 81: 517-24.

de Jong G (1990) Quantitative genetics of reaction norms. J Evol Biol 3: 447–68.

de Jong G (1995) Phenotypic plasticity as a product of selection in a variable environment. Am Nat 145: 493–512.

de Jong WW (1998) Molecules remodel the mammalian tree. Trends Ecol Evol 13: 270–5.

Dekanty A, Romero NM, Bertolin AP, Thomas MG, Leishman CC, et al. (2010) Drosophila genome-wide RNAi screen identifies multiple regulators of HIF–dependent transcription in hypoxia. PLoS Genet 6: e1000994.

De Keulenauer GW, Alexander RW, Ushio-Kukai M, Ishizaka N, Griendling KK (1998) Tumor necrosis factor α activates a p22phox based NADH oxidase in vascular smooth muscle. Biochem J 329: 653–7.

Dekker R, Beukema JJ (1993) Dynamics and growth of a bivalve, Abra tenuis, at the northern edge of its distribution. J Mar Biol Assoc UK 73: 497–511.

Edwards MS, Hernández-Carmona G (2005) Delayed recovery of giant kelp near its southern range limit in the North Pacific following El Niño. Mar Biol 147: 273–9.

de Kovel CGF, Jong G (1999) Responses of sexual and apomictic genotypes of Taraxacum officinale to variation in light. Plant Biol 1: 541–6.

de Kretser DM, Kerr JB (1988) The cytology of the testis. In: Knobil E, Neill J, eds. The Physiology of reproduction. New York, NY: Raven Press.pp 837-932.

de Kretser DM, Kerr JB (1994) The cytology of the testis. In: Knobil E, Neill JD, eds. Physiology and Reproduction. New York, NY: Raven Press. pp 1177-1290.

de la Cova C, Abril M, Bellosta P, Gallart P, Johnston LA (2004) Drosophila myc regulates organ size by inducing cell competition. Cell 117: 107–16.

de la Iglesia F, Elena SF (2007) Fitness declines in Tobacco etch virus upon serial bottleneck transfers. J Virol 81: 4941-7.

de la Peña M, Elena SF, Moya A (2000) Effect of deleterious mutation accumulation of the fitness of RNA bacteriophage MS2. Evolution 54: 686–91.

Delaporte M, Soudant P, Lambert C, Jegaden M, Moal J, et al. (2007) Characterisation of physiological and immunological differences between Pacific oysters (Crassostrea gigas) genetically selected for high or low survival to summer mortalities and fed different rations under controlled conditions. J Exp Mar Biol Ecol 353: 45–57.

de la Vega E, Hall MR, Wilson KJ, Reverter A, Woods RG, et al. (2007) Stress-induced gene expression profiling in the black tiger shrimp Penaeus monodon. Physiol Genomics 31: 126–38.

del Campo R, Morosini MI, de la Pedrosa EG, Fenoll A, Munoz-Almagro C, et al. (2005) Population structure, antimicrobial resistance, and mutation frequencies of Streptococcus pneumoniae isolates from cystic fibrosis patients. J Clin Microbiol 43: 2207–14.

del Carmen Seleme M, Vetter MR, Cordaux R, Bastone L, Batzer MA, Kazazian Jr HH (2006) Extensive individual variation in L1 retrotransposition capability contributes to human genetic diversity. Proc Natl Acad Sci USA 103: 6611–6616.

De Leeuw F, Zhang T, Wauquier C, Huez G, Kruys V, Gueydan C (2007) The cold-inducible RNA-binding protein migrates from the nucleus to cytoplasmic stress granules by a methylation-dependent mechanism and acts as a translational repressor. Exp Cell Res 313: 4130–44.

Delfino F, Walker WH (1998) Stage-specific nuclear expression of NF-kappaB in mammalian testis. Mol Endocrinol 12: 1696–707.

Delfino FJ, Walker WH (1999a) NF-kappaB induces cAMP-response element binding protein gene transcription in Sertoli cells. J Biol Chem 274: 35607–13.

Delfino FJ, Walker WH (1999b) Hormonal regulation of the NFkappaB signaling pathway. Mol Cell Endocrinol 157: 1–9.

Delfino FJ, Boustead JN, Fix C, Walker WH (2003) NF-kappaB and TNF-alpha stimulate androgen receptor expression in Sertoli cells. Mol Cell Endocrinol 201: 1–12.

Delmotte F, Leterme N, Gauthier JP, Rispe C, Simon JC (2002) Genetic architecture of sexual and asexual populations of the aphid Rhopalosiphum padi based on allozyme and microsatellite markers. Mol Ecol 11: 711–23.

Delmotte F, Sabater-Muñoz B, Prunier-Leterme N, Latorre A, Sunnucks P, et al. (2003) Phylogenetic evidence for hybrid origins of asexual lineages in an aphid species. Evolution 57: 1291-303.

De Loof A (2011) Longevity and aging in insects: Is reproduction costly; cheap; beneficial or irrelevant? A critical evaluation of the "trade-off" concept. J Insect Physiol 57: 1-11.

De Lope F, Møller AP (1993) Female reproductive effort depends on the degree of ornamentation of their mates. Evolution 47: 1152–60.

de los Ríos A, Valea S, Ascaso C, Davila A, Kastovsky J, et al. (2010) Comparative analysis of the microbial communities inhabiting halite evaporites of the Atacama Desert. Int Microbiol 13:79-89.

Delport W, Scheffler K, Seoighe C (2009) Models of coding sequence evolution. Brief Bioinform 10: 97-109.

Del Punta K, Charreau EH, Pignataro OP (1996) Nitric oxide inhibits Leydig cell steroidogenesis. Endocrinology 137: 5337-43.

Delsuc F, Stanhope M, Douzery E (2003) Molecular systematics of armadillos (Xenarthra, Dasypodidae): contribution of maximum likelihood and Bayesian analyses of mitochondrial and nuclear genes. Mol Phylogenet Evol 28: 261–75.

DeLuca SZ, O’Farrell PH (2012) Barriers to male transmission of mitochondrial DNA in sperm development. Dev Cell 22: 660–8.

de Magalhães JP, Church GM (2005) Genomes optimize reproduction: aging as a consequence of the developmental program. Physiology (Bethesda) 20: 252–9.

de Magalhães JP, Church GM (2006) Cells discover fire: employing reactive oxygen species in development and consequences for aging. Exp Gerontol 41:1-10.

De Meester L, Weider LJ, Tollrian R (1995) Alternative antipredator defenses and genetic-polymorphism in a pelagic predator–prey system. Nature 378: 483–5.

Demerec M (1937) Frequency of spontaneous mutations in certain stocks of Drosophila melanogaster. Genetics 22: 469-78.

Demetrius L (2005) Of mice and men. EMBO Rep 6: S39–S44.

Demple B, Harrison L (1994) Repair of oxidative damage to DNA: enzymology and biology. Annu Rev Biochem 63: 915–48.

Dempster ER (1955) Maintenance of genetic heterogeneity. Cold Spring Harb Symp Quant Biol 20: 25-32.

Denamur E, Matic I (2006) Evolution of mutation rates in bacteria. Mol Microbiol 60: 820–7.

Denamur E, Bonacorsi S, Giraud A, Duriez P, Hilali F, et al. (2002) High frequency of mutator strains among human uropathogenic Escherichia coli isolates. J Bacteriol 184: 605–9.

den Boer PJ (1968) Spreading of risk and stabilization of animal numbers. Acta Biotheor 18: 165-94.

de Nettancourt D (1997) Self-incompatability in angiosperms. Sex Plant Reprod 10:185–99.

Denekamp NY, Thorne MAS, Clark MS, Kube M, Reinhardt R, Lubzens E (2009) Discovering genes associated with dormancy in the monogonont rotifer Brachionus plicatilis. BMC Genomics 10: 108.

Denekamp NY, Suga T, Hagiwara A, Reinhardt R, Lubzens E (2010) A role for molecular studies in unveiling the pathways for formation of rotifer resting eggs and their survival during dormancy. In: Lubzens E, Cerdà J, Clark MS, eds. Dormancy and Resistance in Harsh Environments. Berlin, Germany: Springer-Verlag. pp 109–132.

Deng HW (1996) Environmental and genetic-control of sexual reproduction in Daphnia. Heredity 76: 449–58.

Deng W, Lin H (2002) miwi, a murine homolog of piwi, encodes a cytoplasmic protein essential for spermatogenesis. Dev Cell 2: 819-30.

Denissenko MF, Chen JX, Tang MS, Pfeifer GP (1997) Cytosine methylation determines hot spots of DNA damage in the human P53 gene. Proc Natl Acad Sci USA 94: 3893-8.

Denk AG, Holzmann A, Peters A, Vermeirssen ELM, Kempenaers B (2005) Paternity in mallards: effects of sperm quality and female sperm selection for inbreeding avoidance. Behav Ecol 16: 825–33.

Denko NC, Giaccia AJ, Stringer JR, Stambrook PJ (1994) The human Ha-ras oncogene induces genomic instability in murine fibroblasts within one cell cycle. Proc Natl Acad Sci USA 91: 5124-8.

Dennett DC (1995) Darwin's dangerous idea. London, UK: Allen Lane, Penguin.

Densmore LD, Moritz CC, Wright JW, Brown WM (1989) Mitochondrial DNA analyses and the origin and relative age of parthenogenetic lizards (genus Cnemidophorus). Nine sexlineatus–group parthenogens. Evolution 43: 969–83.

Denver DR, Morris K, Lynch M, Vassilieva LL, Thomas WK (2000) High direct estimate of the mutation rate in the mitochondrial genome of Caenorhabditis elegans. Science 289: 2342–4.

Denver DR, Morris K, Lynch M, Thomas WK (2004) High mutation rate and predominance of insertions in the Caenorhabditis elegans nuclear genome. Nature 430: 679–82.

Denver DR, Dolan PC, Wilhelm LJ, Sung W, Lucas-Lledo JI, et al. (2009) A genome-wide view of Caenorhabditis elegans base-substitution mutation processes. Proc Natl Acad Sci USA 106: 16310–4.

Depping R, Hägele S, Wagner KF, Wiesner RJ, Camenisch G, et al. (2004) A dominant-negative isoform of hypoxia-inducible factor-1 alpha specifically expressed in human testis. Biol Reprod 71:331-9.

Dercole F, Ferriere R, Rinaldi S (2010) Chaotic Red Queen coevolution in three-species food chains. Proc Biol Sci 277: 2321-30.

de Reviers M, Hochereau-de Reviers MT, Blanc MR, Brillard JP, Courot M, Pelletier J (1980) Control of Sertoli and germ cell populations in the cock and sheep testes. Reprod Nutr Dev 20: 241–9.

de Reviers M, Williams JB (1984) Testis development and production of spermatozoa in the cockerel (Gallus domesticus). In: Cunningham FJ, Lake PE, Hewitt D, eds. Reproductive Biology of Poultry. British Poultry Science Ltd. pp 183–202.

de Rooij DG, Lok D (1987) Regulation of the density of spermatogonia in the seminiferous epithelium of the Chinese hamster: II. Differentiating spermatogonia. Anat Rec 217: 131-6.

de Rooij DG, Russell LD (2000) All you wanted to know about spermatogonia but were afraid to ask. J Androl 21: 776–98.

DeRose CM, Claycamp HG (1991) Oxidative stress effects on conjugational recombination and mutation in catalase-deficient Escherichia coli. Mutat Res 255: 193-200.

Derry LA, Brasier MD, Corfield RM, Rozanov AY, Zhuravlev AY (1994) Sr and C isotopes in Lower Cambrian carbonates from the Siberian Craton; a paleoenvironmental record during the “Cambrian explosion”. Earth Planet Sci Lett 128: 671–81.

Derry WB, Putzke AP, Rothman JH (2001) Caenorhabditis elegans p53: Role in apoptosis, meiosis, and stress resistance. Science 294: 591–5.

Desagher S, Martinou JC (2000) Mitochondria as the central control point of apoptosis. Trends Cell Biol 10: 369-77.

Desai MM, Fisher DS (2007) Beneficial mutation-selection balance and the effect of linkage on positive selection. Genetics 176: 1759–98.

Desai MM, FisherDS, Murray AW (2007) The speed of evolution and maintenance of variation in asexual populations. Curr Biol 17: 385–94.

Desai MM, Fisher DS (2011) The balance between mutators and nonmutators in asexual populations. Genetics 188: 997-1014.

Desai N, Sabanegh E Jr., Kim T, Agarwal A (2009) Free radical theory of aging: implications in male infertility. Urology 75: 14–9.

Desalvo MK, Rvoolstra C, Sunagawa S, Schwarz JA, Stillman JH, et al. (2008) Differential gene expression during thermal stress and bleaching in the Caribbean coral Montastraea faveolata. Mol Ecol 17: 3952–71.

Descamps S, Boutin S, Berteaux D, McAdam AG, Gaillard JM (2008) Cohort effects in red squirrels: the influence of density, food abundance and temperature on future survival and reproductive success. J Anim Ecol 77: 305-14.

Desjardins GC, Beaudet A, Meaney MJ, Brawer JR (1995) Estrogen-induced hypothalamic ß-endorphin neuron loss: a possible model of hypothalamic aging. Exp Gerontol 30: 253-67.

Desjarlais JR, Berg JM (1992) Toward rules relating zinc finger protein sequences and DNA binding site preferences. Proc Natl Acad Sci USA 89: 7345–9.

De Smet C, Lurquin C, Lethe B, Martelange V, Boon T (1999) DNA methylation is the primary silencing mechanism for a set of germ line- and tumor-specific genes with a CpG-rich promoter. Mol Cell Biol 19: 7327–35.

de Smedt V, Szöllösi D, Kloc M (2000) The balbiani body: asymmetry in the mammalian oocyte. Genesis 26: 208–12.

Desset S, Meignin C, Dastugue B, Vaury C (2003) COM, a heterochromatic locus governing the control of independent endogenous retroviruses from Drosophila melanogaster. Genetics 164: 501–509.

Desset S, Buchon N, Meignin C, Coiffet M, Vaury C (2008) In Drosophila melanogaster the COM locus directs the somatic silencing of two retrotransposons through both Piwi-dependent and –independent pathways. PLoS ONE 3: e1526.

de Stordeur E (1997) Nonrandom partition of mitochondria in heteroplasmic Drosophila. Heredity 79: 615-23.

de Stordeur E, Solignac M, MonnerotM, Mounolou JC (1989) The generation of transplasmic Drosophila simulans by cytoplasmic injection: effects of segregation and selection on the perpetuation of mitochondrial DNA heteroplasmy. Mol Gen Genet 220: 127-32.

Deutsch GB, Zielonka EM, Coutandin D, Weber TA, Schäfer B, et al. (2011a) DNA damage in oocytes induces a switch of the quality control factor TAp63α from dimer to tetramer. Cell 144: 566-76.

Deutsch GB, Zielonka EM, Coutandin D, Dötsch V (2011b) Quality control in oocytes: domain-domain interactions regulate the activity of p63. Cell Cycle 10: 1884-5.

Devine SE, Boeke JD (1996) Integration of the yeast retrotransposon Ty1 is targeted to regions upstream of genes transcribed by RNA polymerase III. Genes Dev 10: 620-33.

de Visser JAGM (2002) The fate of microbial mutators. Microbiology 148: 1247–52.

de Visser JAGM, Hoekstra RF, van den Ende H (1996) The effect of sex and deleterious mutations on fitness in Chlamydomonas. Proc R Soc Lond Ser B Biol Sci 263: 193–200.

de Visser JAGM, Hoekstra RF, van den Ende H (1997a) An experimental test for synergistic epistasis and its application in Chlamydomonas. Genetics 145: 815–9.

de Visser JAGM, Hoekstra RF, van den Ende H (1997b) Test of interaction between genetic markers that affect fitness in Aspergillus niger. Evolution 51: 1499–1505.

de Visser JAGM, Zeyl CW, Gerrish PJ, Blanchard JL, Lenski RE (1999) Diminishing returns from mutation supply rate in asexual populations. Science 283: 404-6.

de Visser JA, Hermisson J, Wagner GP, Ancel Meyers L, Bagheri-Chaichian H, et al. (2003) Perspective: Evolution and detection of genetic robustness. Evolution 57: 1959–72.

de Visser JAGM, Rozen DE (2005) Limits to adaptation in asexual populations. J Evol Biol 18: 779-88.

de Visser JAGM, Rozen DE (2006) Clonal interference and the periodic selection of new beneficial mutations in Escherichia coli. Genetics 172: 2093–100.

de Visser JAGM, ElenaSF (2007) The evolution of sex: empirical insights into the roles of epistasis and drift. Nat Rev Genet 8: 139–49.

de Visser JAGM, Park S-C, Krug J (2009) Exploring the effect of sex on empirical fitness landscapes. Am Nat 174:S15–S30.

de Visser JAGM, Cooper TF, Elena SF (2011) The causes of epistasis. Proc R Soc Lond B 278: 3617–24.

Devlin C, Greco S, Martelli F, Ivan M (2011) miR-210: More than a silent player in hypoxia. IUBMB Life 63: 94–100.

Devlin RH, Nagahama Y (2002) Sex determination and sex differentiation in fish: an overview of genetic, physiological, and environmental influences. Aquaculture 208: 191–364.

De Vos K, Goossens V, Boone E, Vercammen D, Vancompernolle K, et al. (1998). The 55-kDa tumor necrosis factor receptor induces clustering of mitochondria through its membrane-proximal region. J Biol Chem 273: 9673-80.

Dewees AA (1982) Premeiotic male recombination in Drosophila melanogaster. J Hered 73: 330-4.

De Waal LC (2001) A viability study of Fallopia japonica stem tissue. Weed Res 41: 447–60.

DeWeese TL, Shipman JM, Larrier NA, Buckley NM, Kidd LR, et al. (1998) Mouse embryonic stem cells carrying one or two defective Msh2 alleles respond abnormally to oxidative stress inflicted by low-level radiation. Proc Natl Acad Sci USA 95: 11915-20.

DeWitt TJ, Sih A, Wilson DS (1998) Costs and limits of phenotypic plasticity. Trends Ecol Evol 13: 77-81.

DeWitt TJ, Scheiner SM, eds. (2004) Phenotypic Plasticity – Functional and Conceptual Approaches. Oxford, UK: Oxford University Press.

Dews M, Homayouni A, Yu D, Murphy D, Sevignani C, et al. (2006) Augmentation of tumor angiogenesis by a myc-activated microRNA cluster. Nat Genet 38: 1060-5.

Dewsbury DA (1982) Ejaculate cost and male choice. Am Nat 119: 601-10.

Dhanabalan S, Mathur PP (2009) Low dose of 2,3,7,8 tetrachlorodibenzo-p-dioxin induces testicular oxidative stress in adult rats under the influence of corticosterone. Exp Toxicol Pathol 61: 415–23.

D'Herde K, Callebaut M, Roels F, De Prest B, van Nassauw L (1995) Homology between mitochondriogenesis in the avian and amphibian oocyte. Reprod Nutr Dev 35: 305-11.

Di X, Bright AT, Bellott R, Gaskins E, Robert J, et al. (2008) A chemotherapy-associated senescence bystander effect in breast cancer cells. Cancer Biol Ther 7: 864-72.

Dianov GL, Souza-Pinto N, Nyaga SG, Thybo T, Stevnsner T, Bohr VA (2001) Base excision repair in nuclear and mitochondrial DNA. Prog Nucleic Acid Res Mol Biol 68: 285–97.

Díaz B, Moreno E (2005) The competitive nature of cells. Exp Cell Res 306: 317–322.

Dickins TE, Rahman Q (2012) The extended evolutionary synthesis and the role of soft inheritance in evolution. Proc R Soc B 279: 2913–21.

Dickinson HG (1994) Self-pollination: simply a social disease. Nature 367:517–8.

Dickinson WJ, Seger J (1999) Cause and effect in evolution. Nature 399: 30.

Dieckmann U, Law R (1996) The dynamical theory of coevolution: a derivation from stochastic ecological processes. J Math Biol 34:579-612.

Dieckmann U, Doebeli M (1999) On the origin of species by sympatric speciation. Nature 400: 354–7.

Diehn M, Cho RW, Lobo NA, Kalisky T, Dorie MJ, et al. (2009) Association of reactive oxygen species levels and radioresistance in cancer stem cells. Nature 458: 780–3.

Diemer T, Allen JA, Hales KH, Hales DB (2003a) Reactive oxygen disrupts mitochondria in MA-10 tumor Leydig cells and inhibits steroidogenic acute regulatory (StAR) protein and steroidogenesis. Endocrinology 144: 2882-91.

Diemer T, Hales DB, Weidner W (2003b) Immune-endocrine interactions and Leydig cell function: the role of cytokines. Andrologia 35: 55-63.

Dierks A, Baumann B, Fischer K (2012) Response to selection on cold tolerance is constrained by inbreeding. Evolution 66: 2384-98.

Di Giacomo M, Barchi M, Baudat F, Edelmann W, Keeney S, Jasin M (2005) Distinct DNA-damage-dependent and -independent responses drive the loss of oocytes in recombination-defective mouse mutants. Proc Natl Acad Sci USA 102: 737–42.

Dijkstra C, Bult C, Bijlsma S, Daan S, Meijer T, Zijlstra M (1990) Brood size manipulations in the kestrel (Falco tinnunculus) – effects on offspring and parent survival. J Anim Ecol 59: 269–85.

Dikalov S, Griendling KK, Harrison DG (2007) Measurement of reactive oxygen species in cardiovascular studies. Hypertension 49: 717–27.

Di Micco R, Fumagalli M, Cicalese A, Piccinin S, Gasparini P, et al. (2006) Oncogene-induced senescence is a DNA damage response triggered by DNA hyper-replication. Nature 444: 638–42.

Dinarello CA (1985) An update on human interleukin-1: from molecular biology to clinical relevance. J Clin Immunol 5: 287-97.

Dinerstein E, McCracken GF (1990) Endangered greater one-horned rhinoceros carry high levels of genetic variation. Conserv Biol 4: 417-22.

Ding D, Lipshitz HD (1993) A molecular screen for polar-localized maternal RNAs in the early embryo of Drosophila. Zygote 1: 257–71.

Dinger ME, Mercer TR, Mattick JS (2008) RNAs as extracellular signaling molecules. J Mol Endocrinol 40: 151–9.

Diniz-Filho JAF, Fuchs S, Arias MC (1999) Phylogeographical autocorrelation of phenotypic evolution in honey bees (Apis mellifera L.). Heredity 83: 671–80.

d'Istria M, Palmiero C, Serino I, Izzo G, Minucci S (2003) Inhibition of the basal and oestradiol-stimulated mitotic activity of primary spermatogonia by melatonin in the testis of the frog, Rana esculenta, in vivo and in vitro. Reproduction 126: 83-90.

d'Istria M, Serino I, Izzo G, Ferrara D, De Rienzo G, Minucci S (2004) Effects of melatonin treatment on Leydig cell activity in the testis of the frog Rana esculenta. Zygote 12: 293-9.

Ditch S, Sammarco MC, Banerjee A, Grabczyk E (2009) Progressive GAA.TTC repeat expansion in human cell lines. PLoS Genet 5: 1000704.

Dix DJ (1997) Hsp70 expression and function during gametogenesis. Cell Stress Chaperones 2: 73-77.

Dix DJ, Allen JW, Collins BW, Mori C, Nakamura N, Poorman-Allen P, Goulding EH, Eddy EM (1996) Targeted gene disruption of Hsp70-2 results in failed meiosis, germ cell apoptosis, and male infertility. Proc Natl Acad Sci USA 93: 3264-8.

Dixon AFG (1989) Parthenogenetic reproduction and the rate of increase in aphids. In: Minks AK, Harrewijn P, eds. Aphids – their biology, natural enemies and control, Volume A. Amsterdam, The Netherlands: Elsevier. pp 269–287.

Dixon AFG (1998) Aphid ecology (2nd edition). London, UK: Chapman and Hall.

Dizdaroglu M (1992) Oxidative damage to DNA in mammalian chromatin. Mutat Res 275: 331–42.

Dizdaroglu M (1993) Quantitative determination of oxidative base damage in DNA by stable isotope-dilution mass spectrometry. Fed Eur Biochem Soc Lett 315: 1–6.

Dizdaroglu M, Jaruga P, Birincioglu M, Rodriguez H (2002) Free radical-induced damage to DNA: mechanisms and measurement. Free Radic Biol Med 32: 1102-15.

Djahanbakhch O, Ezzati M, Zosmer A (2007) Reproductive ageing in women. J Pathol 211: 219–31.

Dmitrieva NI, Bulavin DV, Burg MB (2003) High NaCl causes Mre11 to leave the nucleus, disrupting DNA damage signaling and repair. Am J Physiol Renal Physiol 285: F266-74.

Dobrzynska MM, Baumgartner A, Anderson D (2004) Antioxidants modulate thyroid hormone- and noradrenaline-induced DNA damage in human sperm. Mutagenesis 19: 325-30.

Dobson H, Ghuman S, Prabhakar S, Smith R (2003) A conceptual model of the influence of stress on female reproduction. Reproduction 125: 151-63.

Dobzhansky T (1937) Genetics and the origin of species. New York, NY: Columbia University Press.

Dobzhansky T (1947) Genetics of natural populations. XIV. A response of certain gene arrangements in the third chromosome of Drosophila pseudoobscura to natural selection. Genetics 32: 142-60.

Dobzhansky T (1950) Evolution in the tropics. Am Sci 38: 209–21.

Dobzhansky T (1970) Genetics of the evolutionary process. New York, NY: Columbia University Press.

D'Odorico P, Laio F, Ridolfi L, Lerdau MT (2008) Biodiversity enhancement induced by environmental noise. J Theor Biol 255: 332-7.

Doebeli M (1996) A quantitative genetic model for sympatric speciation. J Evol Biol 9: 893–909.

Doebeli M, Knowlton N (1998) The evolution of interspecific mutualisms. Proc Natl Acad Sci USA 95: 8676–80.

Doebeli M, Dieckmann U (2000) Evolutionary branching and sympatric speciation caused by different types of ecological interactions. Am Nat 156 (Suppl.): S77–S101.

Doebeli M, Dieckmann U (2003) Speciation along environmental gradients. Nature 421: 259–64.

Doerfler W (1983) DNA methylation and gene activity. Annu Rev Biochem 52: 93–124.

Doeringsfeld MR, Schlosser IJ, Elder JF, Evenson DP (2004) Phenotypic consequences of genetic variation in a gynogenetic complex of Phoxinus eos-neogaeus clonal fish (Pisces:Cyprinidae) inhabiting a heterogeneous environment. Evolution Int J Org Evolution 58: 1261–73.

Doerksen T, Benoit G, Trasler JM (2000) Deoxyribonucleic acid hypomethylation of male germ cells by mitotic and meiotic exposure to 5-azacytidine is associated with altered testicular histology. Endocrinology 141: 3235–44.

Doi A, Suzuki H, Matsuura ET (1999) Genetic analysis of temperature-dependent transmission of mitochondrial DNA in Drosophila. Heredity 82: 555-60.

Dolci S, Pellegrini M, Di Agostino S, Geremia R, Rossi P (2001) Signaling through extracellular signal-regulated kinase is required for spermatogonial proliferative response to stem cell factor. J Biol Chem 276: 40225–33.

Dolinoy D, Weinhouse C, Jones T, Rozek L, Jirtle R (2010) Variable histone modifications at the A (vy) metastable epiallele. Epigenetics 5: 637-44.

Dombroski BA, Mathias SL, Nanthakumar E, Scott AF, Kazazian HH Jr (1991) Isolation of an active human transposable element. Science 254: 1805–8.

Domes K, Scheu S, Maraun M (2007) Resources and sex: soil re-colonization by sexual and parthenogenetic oribatid mites. Pedobiologia 51: 1–11.

Domingo E (2007) Virus evolution. In: Knipe DM, Howley PM, eds. Fields virology, 5th edn. Philadelphia, PA: Lippincott Williams & Wilkins.pp 389–421.

Domingo E, Sabo D, Taniguchi T, Weissman C (1978) Nucleotide sequence heterogeneity of an RNA phage population. Cell 13: 735–44.

Domingo E, Biebricher C, Eigen M, Holland JJ (2001) Quasispecies and RNA Virus Evolution: principles and consequences. Austin, TX: Landes Bioscience.

Domingo E, Escarmis C, Martinez MA, Martinez-Salas E, Mateu MG (1992) Foot and mouth disease virus populations are quasispecies. Curr Topics Microbiol Immunol 176: 33–47.

Domingo E, Escarmís E, Sevilla N, Moya A, Elena SF, et al. (1996) Basic concepts in RNA virus evolution. FASEB J 10: 859-64.

Domingo E, Holland JJ (1997) RNA virus mutations and fitness for survival. Annu Rev Microbiol 51: 151–78.

Domingo E, Baranowski E, Ruiz-Jarabo CM, Martin-Hernandez AM, Saiz JC, Escarmis C (1998) Quasispecies structure and persistence of RNA viruses. Emerg Infect Dis 4: 521–7.

Domingo E, Sheldon J, Perales C (2012) Viral quasispecies evolution. Microbiol Mol Biol Rev 76: 159-216.

Dominguez JM 2nd, Davis RT 3rd, McCullough DJ, Stabley JN, Behnke BJ (2011) Aging and exercise training reduce testes microvascular PO2 and alter vasoconstrictor responsiveness in testicular arterioles. Am J Physiol Regul Integr Comp Physiol 301: R801-10.

Donaghue AM, Sonstegard TS, King LM, Smith EJ, Burt DW (1999) Turkey sperm mobility influences paternity in the context of competitive fertilizations. Biol Reprod 61: 422–7.

Donald SP, Sun XY, Hu CA, Yu J, Mei JM, et al. (2001) Proline oxidase, encoded by p53-induced gene-6, catalyzes the generation of proline-dependent reactive oxygen species. Cancer Res 61: 1810–1815.

Doncaster CP, Pound GE, Cox SJ (2000) The ecological cost of sex. Nature 404: 281–5.

Doncaster CP, Pound GE, Cox SJ (2003) Dynamics of regional coexistence for more or less equal competitors. J Anim Ecol 72: 116–26.

Dong ZY, Wang YM, Zhang ZJ, Shen Y, Lin XY, et al. (2006) Extent and pattern of DNA methylation alteration in rice lines derived from introgressive hybridization of rice and Zizania latifolia Griseb. Theor Appl Genet 113: 196–205.

Donker RB, Mouillet JF, Nelson DM, Sadovsky Y (2007) The expression of Argonaute2 and related microRNA biogenesis proteins in normal and hypoxic trophoblasts. Mol Hum Reprod 13: 273–9.

Donnelly CE, Walker GC (1989) groE mutants of Escherichia coli are defective in umuDC-dependent UV mutagenesis. J Bacteriol 171: 6117-25

Donnelly CE, Walker GC (1992) Coexpression of UmuD‘ with UmuC suppresses the UV mutagenesis deficiency of groE mutants. J Bacteriol 174: 3133-9.

Donohue K, Pyle EH, Messiqua D, Heschel MS, Schmitt J (2001) Adaptive divergence in plasticity in natural populations of Impatiens capensis and its consequences for performance in novel habitats. Evolution 55: 692–702.

Doolittle WF, Sapienza C (1980) Selfish genes, the phenotype paradigm and genome evolution. Nature 284: 601–3.

Doreswamy K, Muralidhara (2005) Genotoxic consequences associated with oxidative damage in testis of mice subjected to iron intoxication. Toxicology 206: 169-78.

Dorken ME, Eckert CG (2001) Severely reduced sexual reproduction in northern populations of a clonal plant, Decodonverticillatus (Lythraceae). J Ecol 89: 339–50.

Dorken ME, Neville KJ, Eckert CG (2004) Evolutionary vestigialization of sex in a clonal plant: selection versus neutral mutation in geographically peripheral populations. Proc Biol Sci 271: 2375-80.

Doroszuk A, Wojewodzic MW, Kammenga JE (2006) Rapid adaptive divergence of life-history traits in response to abiotic stress within a natural population of a parthenogenetic nematode. Proc R Soc Biol Sci Ser B 273: 2611-8.

Doroszuk A, te Brake E, Crespo-Gonzalez D, Kammenga JE (2007) Response of secondary production and its components multiple stressors in nematode field populations. J Appl Ecol 44: 446–55.

dos Reis M, Wernisch L (2009) Estimating translational selection in eukaryotic genomes. Mol Biol Evol 26: 451–61.

Dotto GP, Gilman MZ, Maruyama M, Weinberg RA (1986) c-Myc and c-fos expression in differentiating mouse primary keratinocytes. EMBO J 5: 2853–7.

Dowling DK, Simmons LW (2009) Reactive oxygen species as universal constraints in life-history evolution. Proc Biol Sci 276: 1737–45.

Downs CA, Dillon JRT, Fauth JE, Woodley CM (2001) A molecular biomarker system for assessing the health of gastropods (Ilyanassa obsoleta) exposed to natural and anthropogenic stressors. J Exp Mar Biol Ecol 259: 189–214.

Doyle JJ, Flagel LE, Paterson AH, Rapp RA, Soltis DE, et al. (2008) Evolutionary genetics of genome merger and doubling in plants. Annu Rev Genet 42: 443–61.

Doyle SM, Diamond M, McCabe PF (2010) Chloroplast and reactive oxygen species involvement in apoptotic-like programmed cell death in Arabidopsis suspension cultures. J Exp Bot 61: 473–82.

Dózsa-Farkas K (1995) Enchytraeus dudichisp. n., a new fragmenting Enchytraeusspecies from Iran (Oligochaeta, Enchytraeidae). Opusc Zool Budapest 27–28: 41–4.

Dózsa-Farkas K (1996) An interesting reproduction type in Enchytraeids (Oligochaeta). Act Zool Hung 42: 3–10.

Draghi JA, Parsons TL, Wagner GP, Plotkin JB (2010) Mutational robustness can facilitate adaptation. Nature 463:353-5.

Drake AJ, Walker BR, Seckle JR (2005) Intergenerational consequences of fetal programming by in utero exposure to glucocorticoids in rats. Am J Physiol 288: R34-R38.

Drake JW (1991) A constant rate of spontaneous mutation in DNA-based microbes. Proc Natl Acad Sci USA 88: 7160–4.

Drake JW (1993) Rates of spontaneous mutation among RNA viruses. Proc Natl Acad Sci USA 90: 4171–5.

Drake JW (1999) The distribution of rates of spontaneous mutation over viruses, prokaryotes, and eukaryotes. Ann NY Acad Sci 870: 100-7.

Drake JW, Allen EF, Forsberg SA, Preparata R-M, Greening EO (1969) Genetic control of mutation rates in bacteriophage T4. Nature 221: 1128–32.

Drake JW, Charlesworth B, Charlesworth D, Crow JF (1998) Rates of spontaneous mutation. Genetics 148: 1667–86.

Dreher D, Junod AF (1996) Role of oxygen free radicals in cancer development. Eur J Cancer 32A: 30-8.

Dreiss AN, Silva N, Richard M, Moyen F, Thery M, et al. (2008) Condition-dependent genetic benefits of extrapair fertilization in female blue tits Cyanistes caeruleus. J Evol Biol 21: 1814–22.

Dröge W (2002) Free radicals in the physiological control of cell function. Physiol Rev 82: 47–95.

Drossel B, McKane A (2000) Competitive speciation in quantitative genetic models. J Theor Biol 204: 467-78.

Drost JB, Lee WR (1995) Biological basis of germline mutation: comparisons of spontaneous germline mutation rates among drosophila, mouse and human. Environ Mol Mutagen 25 Suppl 26: 48-64.

Drost JB, Lee WR (1998) The developmental basis for germline mosaicism in mouse and Drosophila melanogaster. Genetica 102–103: 421–43.

Drouin G (2006) Chromatin diminution in the copepod Mesocyclops edax: Diminution of tandemly repeated DNA families from somatic cells. Genome 49: 657–65.

Drummond AJ, Pybus OG, Rambaut A, Forsberg R, Rodrigo AG (2003) Measurably evolving populations. Trends Ecol Evol 18: 481–8.

Drummond DA, Bloom JD, Adami C, Wilke CO, Arnold FH (2005) Why highly expressed proteins evolve slowly. Proc Natl Acad Sci USA 102: 14338–43.

Drummond DA, Raval A, Wilke CO (2006) A single determinant dominates the rate of yeast protein evolution. Mol Biol Evol 23: 327–37.

Drummond DA, Wilke CO (2008) Mistranslation-induced protein misfolding as a dominant constraint on coding-sequence evolution. Cell 134: 341-52.

Drummond H, González E, Osorno JL (1986) Parent-offspring cooperation in the blue-footed booby (Sula nebouxii): social roles in infanticidal brood reduction. Behav Ecol Sociobiol 19: 365–72.

Drury SS, Theall K, Gleason MM, Smyke AT, De Vivo I, et al. (2011) Telomere length and early severe social deprivation: linking early adversity and cellular aging. Mol Psychiatry: 1-9.

Dryja TP, Mukai S, Petersen R, Rapaport JM, Walton D, Yandell DW (1989) Parental origin of mutations of the retinoblastoma gene. Nature 339: 556-8.

Dryja TP, Morrow JF, Rapaport JM (1997) Quantification of the paternal allele bias for new germline mutations in the retinoblastoma gene. Hum Genet 100: 446-9.

D’Souza TG, Storhas M, Michiels NK (2005) The effect of ploidy level on fitness in parthenogenetic flatworms. Biol J Linn Soc 85: 191–8.

D’Souza TG, Schulte RD, Schulenberg H, Michiels NK (2006) Paternal inheritance in parthenogenetic forms of the planarian Schmidtea polychroa. Heredity 97: 97–101.

D’Souza TG, Michiels NK (2009) Sex in parthenogenetic planarians: phylogenetic relic or evolutionary resurrection? In: Schön I, Martens K, Van Dijk P, eds. Lost sex. Berlin, Germany: Springer Publications. pp 377-397.

D'Souza TG, Michiels NK (2010) The costs and benefits of occasional sex: theoretical predictions and a case study. J Hered 101 Suppl 1: S34-41.

Du J, Wang Y, Hunter R, Wei Y, Blumenthal R, et al. (2009) Dynamic regulation of mitochondrial function by glucocorticoids. Proc Natl Acad Sci USA 106: 3543-8.

Duarte EA, Clarke DK, Moya A, Domingo E, Holland JJ (1992) Rapid fitness losses in mammalian RNA virus clones due to Muller’s ratchet. Proc Natl Acad Sci USA 89: 6015–9.

Duarte EA, Clarke DK, Moya A, Elena SF, Domingo E, Holland J (1993) Many-trillionfold amplification of single RNA virus particles fails to overcome the Muller's ratchet effect. J Virol 67: 3620-3.

Dubnau D (1982) Genetic transformation in Bacillus subtilis. In: Dubnau D, ed. The Molecular Biology of the Bacilli, Vol. 1. New York, NY: Academic Press. pp 148-175.

Dubnau D (1991) Genetic competence in Bacillus subtilis. Microbiol Rev 55: 395-424.

Dubnau, D (1999) DNA uptake in bacteria. Annu Rev Microbiol 53: 217–44.

Dubnau D, Losick R (2006) Bistability in bacteria. Mol Microbiol 61: 564–72.

Dubois PM, Morel G, Forest MG, Dubois MP (1978) Localization of luteinizing hormone (LH) and testosterone or dihydrotestosterone (DHT) in the gonadotropic cells of the anterior pituitary by using ultracryomicrotomy and immunocytochemistry. Horm Metab Res 10: 250-2.

Duchen MR (1999) Contributions of mitochondria to animal physiology: from homeostatic sensor to calcium signalling and cell death. J Physiol 516: 1-17.

Duclos M, Gouarne C, Martin C, Rocher C, Mormède P, Letellier T (2004) Effects of corticosterone on muscle mitochondria identifying different sensitivity to glucocorticoids in Lewis and Fischer rats. Am J Physiol Endocrinol Metab 286: E159–E167.

Ducsay CA, Myers DA (2011) eNOS activation and NO function: differential control of steroidogenesis by nitric oxide and its adaptation with hypoxia. J Endocrinol 210: 259-69.

Dudley JW (1977) Seventy-six generations of selection for oil and protein percentage in maize. In: Pollack E, Kempthorne O, Bailey TB, eds. Proceedings of the International Conference on Quantitative Genetics. Ames, IA: Iowa State University Press. pp 459-473.

Dudycha JL (2004) Mortality dynamics of Daphnia in contrasting habitats and their role in ecological divergence. Freshwater Biol 49: 505–14.

Dufau ML, Tinajero JC, Fabbri A (1993) Corticotropin-releasing factor: an antireproductive hormone of the testis. FASEB J 7: 299–307.

Duffin PM, Seifert HS (2010) DNA uptake sequence-mediated enhancement of transformation in Neisseria gonorrhoeae is strain dependent. J Bacteriol 192: 4436-44.

Duffy MA, Brassil CE, Hall SR, Tessier AJ, Caceres CE, Conner JK (2008) Parasite-mediated disruptive selection in a natural Daphnia population. BMC Evol Biol 8: 80.

Duffy S, Shackelton LA, Holmes EC (2008) Rates of evolutionary change in viruses: patterns and determinants. Nat Rev Genet 9: 267–76.

Dufour S, Weltzien F-A, Sébert M-E, Le Belle N, Vidal B, et al. (2005) Dopaminergic inhibition of reproduction in teleost fishes: ecophysiological and evolutionary implications. Ann NY Acad Sci 1040: 9–22.

Dufty AM, Clobert J, Møller AP (2002) Hormones, developmental plasticity and adaption. Trends Ecol Evol 17: 190–6.

Duigou S, Ehrlich SD, Noirot P, Noirot-Gros MF (2004) Distinctive genetic features exhibited by the Y-family DNA polymerases in Bacillus subtilis. Mol Microbiol 54: 439–51.

Duigou S, Ehrlich SD, Noirot P, Noirot-Gros MF (2005) DNA polymerase I acts in translesion synthesis mediated by the Y-polymerases in Bacillus subtilis. Mol Microbiol 57: 678–90.

Dukas R (1998) Evolutionary ecology of learning. In: Dukas R, ed. Cognitive ecology. Chicago, IL: University of Chicago Press. pp 129–174.

Dulic V (2011) Be quiet and you’ll keep young: does mTOR underlie p53 action in protecting against senescence by favoring quiescence? Aging (Albany NY) 3: 3-4.

Dumollard R, Marangos P, Fitzharris G, Swann K, Duchen M, Carroll J (2004) Sperm-triggered [Ca2+] oscillations and Ca2+ homeostasis in the mouse egg have an absolute requirement for mitochondrial ATP production. Development 131: 3057-67.

Dumollard R, Duchen M, Sardet C (2006) Calcium signals and mitochondria at fertilisation. Semin Cell Dev Biol 17: 314–23.

Dumollard R, Ward Z, Carroll J, Duchen MR (2007a) Regulation of redox metabolism in the mouse oocyte and embryo. Development 134: 455–65.

Dumollard R, Duchen M, Carroll J (2007b) The role of mitochondrial function in the oocyte and embryo. Curr Top Dev Biol 77: 21-49.

Dunbrack RL, Coffin C, Howe R (1995) The cost of males and the paradox of sex: Experimental investigation of the short-term competitive advantages of evolution in sexual populations. Proc R Soc Lond B Biol Sci 262: 45-9.

Duncan AB, Mitchell SE, Little TJ (2006) Parasite-mediated selection and the role of sex and diapause in Daphnia. J Evol Biol 19: 1183-9.

Duncan AB, Little TJ (2007) Parasite-driven genetic change in a natural population of Daphnia. Evolution 61: 796–803.

Duncan BK, Miller JH (1980) Mutagenic deamination of cytosine residues in DNA. Nature 287: 560–1.

Dunkel L, Hirvonen V, Erkkilä K (1997) Clinical aspects of male germ cell apoptosis during testis development and spermatogenesis. Cell Death Differ 4: 171-9.

Dunn GA, Bale TL (2011) Maternal high-fat diet effects on third-generation female body size via the paternal lineage. Endocrinology 152: 2228-36.

Dunnington EA, Siegel PB (1996) Long-term divergent selection for eight-week body weight in white Plymouth rock chickens. Poultry Sci 75: 1168–79.

Dunny GM, Leonard BAB, Hedberg PJ (1995) Pheromone-inducible conjugation in Enterococcus faecalis: interbacterial and host-parasite chemical communication. J Bacteriol 177: 871-6.

Dupressoir A, Heidmann T (1996) Germ line-specific expression of intracisternal A-particle retrotransposons in transgenic mice. Mol Cell Biol 16: 4495–503.

Duran EH, Morshedi M, Taylor S, Oehninger S (2002) Sperm DNA quality predicts intrauterine insemination outcome: a prospective cohort study. Hum Reprod 17:3122-8.

Durazzi JT, Stehli FG (1972) Average generic age, the planetary temperature gradient, and pole location. Syst Zool 21: 384–9.

Duret L (2002) Evolution of synonymous codon usage in metazoans. Curr Opin Genet Dev 12: 640-9.

Duret L, Mouchiroud D (1999) Expression pattern, and surprisingly, gene length shape codon usage in Caenorhabditis, Drosophila, and Arabidopsis. Proc Natl Acad Sci USA 96: 4482–7.

Duret L, Mouchiroud D (2000) Determinants of substitution rates in mammalian genes: expression pattern affects selection intensity but not mutation rate. Mol Biol Evol 17: 68–74.

Duret L, Cohen J, Jubin C, Dessen P, Goût JF, et al. (2008) Analysis of sequence variability in the macronuclear DNA of Paramecium tetraurelia: A somatic view of the germline. Genome Res 18: 585–96.

Duret L, Galtier N (2009) Biased gene conversion and the evolution of mammalian genomic landscapes. Annu Rev Genomics Hum Genet 10: 285–311.

Dutta K, Datta G, Verma NC (1996) Effect of nitrogen starvation on DNA repair in Saccharomyces cerevisiae. J Gen Appl Microbiol 42: 27-37.

Duxbury MS, Ashley SW, Whang EE (2005) RNA interference: a mammalian SID-1 homologue enhances siRNA uptake and gene silencing efficacy in human cells. Biochem Biophys Res Commun 331: 459-63.

Dwyer G, Elkinton JS, Buonaccorsi JP (1997) Host heterogeneity in susceptibility and disease dynamics: tests of a mathematical model. Am Nat 150: 685–707.

Dybas LK, Dybas HS (1981) Coadaptation and taxonomic differentiation of sperm and spermathecae in featherwing beetles. Evolution 35: 168–74.

Dybdahl MF, Lively CM (1998) Host-parasite coevolution: evidence for rare advantage and time-lagged selection in a natural population. Evolution 52: 1057–66.

Dybdhal MF, Kane SL (2005) Adaptation vs. phenotypic plasticity in the success of a clonal invader. Ecology 86: 1592–601.

Dycaico MJ, Provost GS, Kretz PL, Ransom SL, Moores JC, Short JM (1994) The use of shuttle vectors for mutation analysis in transgenic mice and rats. Mutat Res 307: 461–78.

Dykhuizen D, Davies M (1980) An experimental model: bacterial specialists and generalists competing in chemostats. Ecology 61: 1213–27.

Dym M (1994)Spermatogonial stem cells of the testis. Proc Natl Acad Sci USA 91: 11287–9.

Dym M, Fawcett DW (1970) The blood-testis barrier in the rat and the physiological compartmentation of the seminiferous epithelium. Biol Reprod 3: 308–26.

Dym M, Raj HGM (1977) Response of adult rat Sertoli cells and Leydig cells to depletion of luteinizing hormone and testosterone. Biol Reprod 17: 676–96.

Dynowski M, Schaaf G, Loque D, Moran O, Ludewig U (2008) Plant plasma membrane water channels conduct the signaling molecule H2O2. Biochem J 414: 53–61.

Dzeja PP, Bortolon R, Perez-Terzic C, Holmuhamedov EL, Terzic A (2002) Energetic communication between mitochondria and nucleus directed by catalyzed phosphotransfer. Proc Natl Acad Sci USA 99: 10156-61.

Dziuk PJ (1996) Factors that influence the proportion of offspring sired by a male following heterospermic insemination. Anim Reprod Sci 43: 65–88.

Eads BD, Colbourne JK, Bohuski E, Andrews J (2007) Profiling sex-biased gene expression during parthenogenetic reproduction in Daphnia pulex. BMC Genomics 8: 464.

Eads BD, Andrews J, Colbourne JK (2008) Ecological genomics in Daphnia: stress responses and environmental sex determination. Heredity 100: 184–90.

Earl DJ, Deem MW (2004) Evolvability is a selectable trait. Proc Natl Acad Sci USA 101:11531-6.

Eastgate JA, Symons JA, Wood NC, Grinlinton FM, di Giovine FS, Duff GW (1988) Correlation of plasma interleukin 1 levels with disease activity in rheumatoid arthritis. Lancet 2: 706-9.

Eastwood MD, Cheung SW, Lee KY, Moffat J, Meneghini MD (2012) Developmentally programmed nuclear destruction during yeast gametogenesis. Dev Cell 23: 35-44.

Ebel GD, Dupuis AP 2nd, Ngo K, Nicholas D, Kauffman E, et al. (2001) Partial genetic characterization of West Nile virus strains, New York State, 2000. Emerg Infect Dis 7: 650–3.

Ebel GD, Carricaburu J, Young D, Bernard KA, Kramer LD (2004) Genetic and phenotypic variation of West Nile virus in New York, 2000–2003. Am J Trop Med Hyg 71: 493–500.

Eberhard WG (1980) Evolutionary consequences of intracellular organelle competition. Q Rev Biol 55: 231-49.

Eberhard WG (1996) Female control: Sexual selection by cryptic female choice. Princeton, NJ: Princeton University Press.

Eberhard WG (2004) Rapid divergent evolution of sexual morphology: comparative tests of antagonistic coevolution and traditional female choice. Evolution 58: 1947-70.

Ebert BL, Firth JD, Ratcliffe PJ (1995) Hypoxia and mitochondrial inhibitors regulate expression of glucose transporter-1 via distinct cis-acting sequences. J Biol Chem 270: 29083–9.

Ebert D (2008) Host-parasite coevolution: Insights from the Daphnia-parasite model system. Curr Opin Microbiol 11: 290-301.

Echeverria PC, Picard D (2010) Molecular chaperones, essential partners of steroid hormone receptors for activity and mobility. Biochim Biophys Acta 1803: 641–9.

Echols H (1981) SOS functions, cancer and inducible mutation. Cell 25: 1–2.

Echtay KS (2007) Mitochondrial uncoupling proteins--what is their physiological role? Free Radic Biol Med 43: 1351-71.

Echtay KS, Roussel D, St-Pierre J, Jekabsons MB, Cadenas S, et al. (2002a) Superoxide activates mitochondrial uncoupling proteins. Nature 415: 96–9.

Echtay KS, Murphy MP, Smith RA, Talbot DA, Brand MD (2002b) Superoxide activates mitochondrial uncoupling protein 2 from the matrix side. Studies using targeted antioxidants. J Biol Chem 277: 47129–35.

Echtay KS, Esteves TC, Pakay JL, Jekabsons MB, Lambert AJ, et al. (2003) A signaling role for 4-hydroxy-2-nonenal in regulation of mitochondrial uncoupling. EMBO J 22: 4103-10.

Echtay KS, Pakay JL, Esteves TC, Brand MD (2005) Hydroxynonenal and uncoupling proteins: a model for protection against oxidative damage. Biofactors 24: 119-30.

Eckelbarger KJ, Tyler PA, Langton RW (1998) Gonadal morphology and gametogenesis in the sea pen Pennatula aculeate (Anthozoa: Pennatulacea) from the Gulf of Maine. Mar Biol 132: 677-90.

Eckelbarger KJ, Young CM (2002) Spermiogenesis and modified sperm morphology in the "seepworm" Methanoaricia dendrobranchiata (Polychaeta: Orbiniidae) from a methane seep environment in the Gulf of Mexico: implications for fertilization biology. Biol Bull 203: 134-43.

Eckert CG (2002) The loss of sex in clonal plants. Evol Ecol15: 501–20.

Eckert CG, Allen M (1997) Cryptic self-incompatibility in tristylous Decodon verticillatus (Lythraceae). Am J Bot 84:1391–7.

Eckert CG, Dorken ME, Mitchell SA (1999) Loss of sex in clonal populations of a flowering plant, Decodon verticillatus (Lythraceae). Evolution 53: 1079-92.

Eckert CG, Samis KE, Lougheed SC (2008) Genetic variation across species' geographical ranges: the central-marginal hypothesis and beyond. Mol Ecol 17: 1170-88.

Edashige K, Sakamoto M, Kasai M (2000) Expression of mRNAs of the aquaporin family in mouse oocytes and embryos. Cryobiology 40: 171-5.

Eddy EM (1974) Fine structural observations on the form and distribution of nuage in germ cells of the rat. Anat Rec 178: 731–57.

Eddy EM (1975) Germ plasm and the differentiation of the germ cell line. Int Rev Cytol 43: 229–80.

Eddy EM (1999) Role of heat shock protein HSP70-2 in spermatogenesis. Rev Reprod 4: 23-30.

Edelman GM (1987) Neural Darwinism. The Theory of Neuronal Group Selection. New York, NY: Basic Books.

Edelman GM, Mountcastle VB (1978) The Mindful Brain: Cortical Organization and the Group Selective Theory of Higher Brain Function. Cambridge, MA: MIT Press.

Edelman GM, Gally JA (2001) Degeneracy and complexity in biological systems. Proc Natl Acad Sci USA 98: 13763-8.

Eden A, Gaudet F, Waghmare A, Jaenisch R (2003) Chromosomal instability and tumors promoted by DNA hypomethylation. Science 300: 455.

Eden M (1967) Inadequacies of neo-Darwinian evolution as a scientific theory. In: Moorhead PS, Kaplan MM, eds. Mathematical challenges to the neo-Darwinian interpretation of evolution. Philadelphia, PA: Wistar Institute Press. pp 109-111.

Editorial (1989) Thrifty genotype rendered detrimental by progress? Lancet 2: 839-40.

Edmunds PJ, Elahi R (2007) The demographics of a 15-year decline in cover of the Caribbean reef coral Montastraea annularis. Ecol Monogr 77:3–18.

Edwards CTT, Holmes EC, Pybus OG, Wilson DJ, Viscidi RP, et al. (2006) Evolution of the HIV-1 envelope gene is dominated by purifying selection. Genetics 174: 1441-53.

Edwards TC, Collopy MW (1983) Obligate and facultative brood reduction in eagles; an examination of factors that influence fratricide. Auk 100: 630-5.

Edwards RA, Witherspoon M, Wang K, Afrasiabi K, Pham T, et al. (2009) Epigenetic repression of DNA mismatch repair by inflammation and hypoxia in inflammatory bowel disease–associated colorectal cancer. Cancer Res 69: 6423–9.

Egel R (1995) The synaptonemal complex and the distribution of meiotic recombination events. Trends Genet 11: 206-8.

Ehlers A, Worm B, Reusch TBH (2008) Importance of genetic diversity in eelgrass Zostera marina for its resilience to global warming. Mar Ecol Prog Ser 355: 1–7.

Ehrendorfer F (1980) Polyploidy and distribution. In: Lewis WH, ed. Polyploidy – Biological Relevance. New York, NY: Plenum Press. pp 45–60.

Ehrlén J, Lehtilä K (2002) How perennial are perennial plants? Oikos 98: 308–22.

Ehrlich M (2009) DNA hypomethylation in cancer cells. Epigenomics 1: 239-59.

Ehrlich M, Gama-Sosa MA, Huang LH, Midgett RM, Kuo KC, et al. (1982) Amount and distribution of 5-methylcytosine in human DNA from different types of tissues of cells. Nucleic Acids Res 10: 2709–21.

Ehrlich PR, Raven PH (1964) Butterflies and plants: a study in coevolution. Evolution 18: 586-608.

Eiben AE, Smith JE (2008) Introduction to Evolutionary Computing. Berlin, Germany: Springer.

Eichler EE, Flint J, Gibson G, Kong A, Leal SM, Moore JH, Nadeau JH (2010) Missing heritability and strategies for finding the underlying causes of complex disease. Nat Rev Genet 11: 446–50.

Eickbush TH, Furano AV (2002) Fruit flies and humans respond differently to retrotransposons. Curr Opin Genet Dev 12: 669–74.

Eigen M (1971) Self-organization of matter and the evolution of biological macromolecules. Naturwissenschaften 58: 465–523.

Eigen M (1987) New concepts for dealing with the evolution of nucleic acids. Cold Spring Harb Symp Quant Biol 52: 307–19.

Eigen M (1992) Steps towards life: a perspective on evolution. Oxford, UK: Oxford University Press.

Eigen M (1993) The origin of genetic information: viruses as models. Gene 135:37–47.

Eigen M (1996) On the nature of viral quasispecies. Trends Microbiol 4: 216–8.

Eigen M, Schuster P (1977) A principle of natural selforganization. Naturwissenschaften 64: 541–65.

Eigen M, Schuster P (1979) The Hypercycle: A Principle of Natural Self-Organization. Berlin, Germany: Springer.

Eigen M, McCaskill J, Schuster P (1988) Molecular quasispecies. J Phys Chem 92: 6881–91.

Eigen M, McCaskill J, Schuster P (1989) The molecular quasi-species. Adv Chem Phys 75: 149-263.

Eigner JE, Boldker H, Michaels G (1961) The thermal degradation of nucleic acids. Biochim Biophys Acta 51: 165–8.

Eilers M, Eisenman RN (2008) Myc’s broad reach. Genes Dev 22: 2755–66.

Einum S, Fleming IA (2004) Environmental unpredictability and offspring size: conservative versus diversified bet-hedging. Evol Ecol Res 6: 443–55.

Eisenberg E, Levanon EY (2003) Human housekeeping genes are compact. Trends Genet 19: 362-5.

Eissenstat DM, Yanai RD (1997) The ecology of root lifespan. Adv Ecol Res 27: 1–60.

Eitam A, Blaustein L, Mangel M (2005) Density and intercohort priority effects on larval Salamandra salamandra in temporary pools. Oecologia 146: 36–42.

Eitan Y, Soller M (2004) Selection induced genetic variation: A new model to explain direct and indirect effects of sixty years of commercial selection for juvenile growth rate in broiler chickens, with implications for episodes of rapid evolutionary change. In: Wasser SP, ed. Evolutionary Theory and Processes: Modern Horizons. Papers in Honor of Eviatar Nevo. Dordrecht, The Netherlands: Kluwer Academic. pp 153–176.

Eizaguirre C, Yeates SE, Lenz TL, Kalbe M, Milinski M (2009) MHC-based mate choice combines good genes and maintenance of MHC polymorphism. Mol Ecol 18: 3316-29.

Ekram MB, Kang K, Kim H, Kim J (2012) Retrotransposons as a major source of epigenetic variations in the mammalian genome. Epigenetics 7: 370-82.

El-Awady RA, Dikomey E, Dahm-Daphi J (2001) Heat effects on DNA repair after ionizing radiation: hyperthermia commonly increases the number of nonrepaired double-strand breaks and structural rearrangements. Nucleic Acids Res 29: 1960-6.

Elder JF, Schlosser IJ (1995) Extreme clonal uniformity of Phoxinus eos/neogaeus gynogens (Pisces: Cyprinidae) among variable habitats in northern Minesota beaver ponds. Proc Natl Acad Sci USA 92: 5001–5.

Eldredge N, Gould SJ (1972) Punctuated equilibrium: an alternative to phyletic gradualism. In: Schopf TJM, ed. Models in paleobiology. San Francisco, CA: Freeman, Cooper. pp 82–115.

Eldredge N, Thompson JN, Brakefield PM, Gavrilets S, Jablonski D, Jackson JBC, et al. (2005) The dynamics of evolutionary stasis. Paleobiology 31: 133–45.

Elena SF (1999) Little evidence for synergism among deleterious mutations in a nonsegmented RNA virus. J Mol Evol 49: 703–7.

Elena SF, Lenski RE (1997) Test of synergistic interactions among deleterious mutations in bacteria. Nature 390: 395–8.

Elena SF, Moya A (1999) Rate of deleterious mutation and the distribution of its effects on fitness in vesicular stomatitis virus. J Evol Biol 12: 1078–88.

Elena SF, Lenski RE (2003) Evolution experiments with microorganisms: the dynamics and genetic bases of adaptation. Nat Rev Genet 4: 457–69.

Elena SF, de Visser JAGM (2003) Environmental stress and the effects of mutation. J Biol 2: 12.

Elena SF, Sanjuán R (2008) The effect of genetic robustness on evolvability in digital organisms. BMC Evol Biol 8: 284.

Elgar M, Crespi J (1992) Cannibalism: ecology and evolution among diverse taxa. Oxford, UK: Oxford University Press.

El-Gohary M, Awara WM, Nassar S, Hawas S (1999) Deltamethrin-induced testicular apoptosis in rats: the protective effect of nitric oxide synthase inhibitor. Toxicology 132: 1-8.

Elkins C, Thomas CE, Seifert HS, Sparling PF (1991) Species-specific uptake of DNA by gonococci is mediated by a 10-base-pair sequence. J Bacteriol 173: 3911-3.

Elkjær ML, Nejsum LN, Gresz V, Kwon TH, Jensen UB, et al. (2001) Immunolocalization of aquaporin-8 in rat kidney, gastrointestinal tract, testis, and airways. Am J Physiol Renal Physiol 281: F1047–F1057.

Ellegren H (2000) Microsatellite mutations in the germline: implications for evolutionary inference. Trends Genet 16: 551–8.

Ellegren H (2004) Microsatellites: simple sequences with complex evolution. Nat Rev Genet 5: 435-45.

Ellegren H (2007) Characteristics, causes and evolutionary consequences of male-biased mutation. Proc R Soc Lond B 274: 1–10.

Ellegren H (2009) A selection model of molecular evolution incorporating the effective population size. Evolution 63: 301-5.

Ellegren H, Fridolfsson AK (1997) Male-driven evolution of DNA sequences in birds. Nat Genet 17: 182-4.

Ellegren H, Fridolfsson AK (2003) Sex-specific mutation rates in salmonid fish. J Mol Evol. 56: 458–63.

Ellegren H, Smith NGC, Webster MT (2003) Mutation rate variation in the mammalian genome. Curr Opin GenetDev 13: 562–8.

Ellegren H, Parsch J (2007) The evolution of sex-biased genes and sex-biased gene expression. Nat Rev Genet 8: 689-98.

Ellegren H, Sheldon BC (2008) Genetic basis of fitness differences in natural populations. Nature 452: 169-75.

Ellgaard L, Helenius A (2003) Quality control in the endoplasmic reticulum. Nat Rev Mol Cell Biol 4: 181-91.

Ellish NJ, Saboda K, O'Connor J, Nasca PC, Stanek EJ, Boyle C (1996) A prospective study of early pregnancy loss. Hum Reprod 11: 406-12.

Ellner S (1985) ESS germination strategies in randomly varying environments. I. Logistic-type models. Theor Popul Biol 28: 50-79.

Ellner S (1986) Germination dimorphisms and parent-offspring conflict in seed germination. J Theor Biol 123: 173-85.

Ellner SP (1997) You bet your life: Life-history strategies in fluctuating environments. In: Othmer HG, Adler HG, Lewis MA, Dallon JC, eds. Case Studies in Mathematical Modeling: Ecology, Physiology and Cell Biology. Upper Saddle River, NJ: Prentice Hall. pp 3–24.

Ellner SP, Hairston NG Jr, Kearns CM, Babai D (1999) The roles of fluctuating selection and long-term diapause in microevolution of diapause timing in a freshwater copepod. Evolution 53: 111–22.

Ellstrand NC, Roose ML (1987) Patterns of genotypic diversity in clonal plant species. Am J Bot 74: 123-31.

El-Maarri O, Olek A, Balaban B, Montag M, van der Ven H, et al. (1998) Methylation levels at selected CpG sites in the factor VIII and FGFR3 genes, in mature female and male germ cells: implications for male-driven evolution. Am J Hum Genet 63: 1001–8.

Elman I, Breier A (1997) Effects of acute metabolic stress on plasma progesterone and testosterone in male subjects: relationship to pituitary-adrenocortical axis activation. Life Sci 61: 1705-12.

El Mouatassim S, Guérin P, Ménézo Y (1999) Expression of genes encoding antioxidant enzymes in human and mouse oocytes during the final stages of maturation. Mol Hum Reprod 5: 720–5.

El Naby A (2012) Expression analysis of regulatory microRNA in bovine cumulus oocyte complex and preimplantation embryos. Thesis. Bonn, Germany: University of Bonn.

Elowitz MB, Levine AJ, Siggia ED, Swain PS (2002) Stochastic gene expression in a single cell. Science 297: 1183–6.

Elroy-Stein O, Bernstein Y, Groner Y (1986) Overproduction of human Cu/Zn-superoxide dismutase in transfected cells: extenuation of paraquat-mediated cytotoxicity and enhancement of lipid peroxidation. EMBO J 5: 615-22.

Elroy-Stein O, Groner Y (1988) Impaired neurotransmitter uptake in PC12 cells overexpressing human Cu/Zn-superoxide dismutase--implication for gene dosage effects in Down syndrome. Cell 52: 259-67.

Else PL, Hulbert AJ (2003) Membranes as metabolic pacemakers. Clin Exp Pharmacol Physiol 30: 559-64.

Else PL, Turner N, Hulbert AJ (2004) The evolution of endothermy: role for membranes and molecular activity. Physiol Biochem Zool 77: 950-8.

Elsey RH, Joanen T, McNease L, Lance V (1990) Stress and plasma corticosterone levels in the American alligator-relationships with stocking density and nesting success. Comp Biochem Physiol 95: 55-63.

Elsey RM, Lance VA, Joanen T, McNease L (1991) Acute stress suppresses plasma estradiol levels in female alligators (Alligator mississipiensts). Comp Biochem Physiol A Comp Physiol 100: 649-52.

Elzinga JA, Chevasco V, Mappes J, Grapputo A (2012) Low parasitism rates in parthenogenetic bagworm moths do not support the parasitoid hypothesis for sex. J Evol Biol 25: 2547-58.

Ema M, Hirota K, Mimura J, Abe H, Yodoi J, et al. (1999) Molecular mechanisms of transcription activation by HLF and HIF1alpha in response to hypoxia: their stabilization and redox signal-induced interaction with CBP/p300. EMBO J 18: 1905–14.

Emmerich R, Giez I, Lange OL, Proksch P (1993) Toxicity and antifeedant activity of lichen compounds against the polyphagous herbivorous insect Spodoptera littoralis. Phytochem 33: 1389–94.

Emara MM, Fujimura K, Sciaranghella D, Ivanova V, Ivanov P, Anderson P (2012) Hydrogen peroxide induces stress granule formation independent of eIF2α phosphorylation. Biochem Biophys Res Commun 423: 763-9.

Emlen JM (1966) The role of time and energy in food preference. Am Nat 100: 611-17.

Enattah NS, Jensen TG, Nielsen M, Lewinski R, Kuokkanen M, et al. (2008) Independent introduction of two lactase-persistence alleles into human populations reflects different history of adaptation to milk culture. Am J Hum Genet 82:57–72.

Enders LS, Nunney L (2010) Sex-specific effects of inbreeding in wild-caught Drosophila melanogaster under benign and stressful conditions. J Evol Biol 23: 2309–23.

Endler JA (1986) Natural Selection in the Wild. Princeton, NJ: Princeton University Press.

Endo T, Aten RF, Leykin L, Behrman HR (1993) Hydrogen peroxide evokes antisteroidogenic and antigonadotropic actions in human granulosa luteal cells. J Clin Endocrinol Metab 76: 337–42.

Enfield FD (1980) Long term effects of selection: the limits to response. In: Robertson A, ed. Selection Experiments in Laboratory and Domestic Animals. Slough, UK: Commonwealth Agricultural Bureau. pp 69-86.

Engels BM, Hutvagner G (2006) Principles and effects of microRNA-mediated post-transcriptional gene regulation. Oncogene 25: 6163–9.

Engels WR, Johnson-Schlitz D, Flores C, White L, Preston CR (2007) A third link connecting aging with double strand break repair. Cell Cycle 6: 131-5.

Engelward BP, Boosalis MS, Chen BJ, Deng Z, Siciliano MJ, Samson LD (1993) Cloning and characterization of a mouse 3- methyladenine/7-methyl-guanine/3-methylguanine DNA glycosylase cDNA whose gene maps to chromosome 11. Carcinogenesis 14: 175–81.

Enghoff H (1976) Parthenogenesis and bisexuality in the millipede, Nemasoma varicorne C. L. Koch, 1847 (Diplopoda: Blaniulidae). Morphological, ecological and biogeographical aspects. Vidensk Meddr Dansk Naturhistorisk Forening 139: 21–59.

Enright AJ, John B, Gaul U, Tuschl T, Sander C, et al. (2003) MicroRNA targets in Drosophila. Genome Biol 5: R1.

Entringer S, Buss C, Wadhwa PD (2010) Prenatal stress and developmental programming of human health and disease risk: Concepts and integration of empirical findings. Curr Opin Endocrinol Diabetes Obes 17: 507–16.

Entringer S, Epel ES, Kumsta R, Lin J, Hellhammer DH, et al. (2011) Stress exposure in intrauterine life is associated with shorter telomere length in young adulthood. Proc Natl Acad Sci USA 108: E513-8.

Eöry L, Halligan DL, Keightley PD (2010) Distributions of selectively constrained sites and deleterious mutation rates in the hominid and murid genomes. Mol Biol Evol 27: 177–92.

Epel E (2009) Telomeres in a life-span perspective: A new “psychobiomarker”? Curr Dir Psychol Sci 18: 6–10.

Epel ES, Blackburn EH, Lin J, Dhabhar FS, Adler NE, et al. (2004) Accelerated telomere shortening in response to life stress. Proc Natl Acad Sci USA 101: 17312–5.

Epp GT (1996) Clonal variation in the survival and reproduction of Daphnia pulicaria under low-food stress. Freshwater Biol 35: 1–10.

Eppig JJ (1996) Coordination of nuclear and cytoplasmic oocyte maturation in eutherian mammals. Reprod Fertil Dev 8: 485–9.

Eppig JJ, Wigglesworth K (1995) Factors affecting the developmental competence of mouse oocytes grown in vitro: oxygen concentration. Mol Reprod Dev 42: 447-56.

Epplen C, Melmer G, Siedlaczck I, Schwaiger FW, Mäueler W, Epplen JT (1993) On the essence of “meaningless” simple repetitive DNA in eukaryote genomes. In: Pena SDJ, Chakraborty R, Epplen JT, Jeffreys AJ, eds. DNA Fingerprinting: State of the Science. Basel, Switzerland: Birkhäuser Verlag. pp 29-45.

Eppley SM, Rosenstiel TN, Graves CB, Llaneza García E (2011) Limits to sexual reproduction in geothermal bryophytes. Int J Plant Sci 172: 870-8.

Erdmann G, Floren A, Lisenmair KE, Scheu S, Maraun M (2006) Little effect of forest age on oribatid mites on the bark of trees. Pedobiologia 50: 433-41.

Ergün S, Kiliç N, Fiedler W, Mukhopadhyay AK (1997) Vascular endothelial growth factor and its receptors in normal human testicular tissue. Mol Cell Endocrinol 131: 9–20.

Ergun S, Buschmann C, Heukeshoven J, Dammann K, Schnieders F, et al. (2004) Cell type-specific expression of LINE-1 open reading frames 1 and 2 in fetal and adult human tissues. J Biol Chem 279: 27753–63.

Erhardt S, Su IH, Schneider R, Barton S, Bannister AJ, et al. (2003) Consequences of the depletion of zygotic and embryonic enhancer of zeste 2 during preimplantation mouse development. Development 130: 4235–48.

Erickson RP (1990) Postmeiotic gene-expression. Trends Genet 6: 264–9.

Eriksen MS, Torjesen HA, Bakken PA (2003) Prenatal exposure to corticosterone impairs embryonic development and increases fluctuating assymetry in chickens (Gallus gallus domesticus). Br Poult Sci 44: 690–7.

Eriksen MS, Espmark A, Braastad BO, Salte R, Bakken M (2007) Long-term effects of maternal cortisol exposure and mild hyperthermia during embryogeny on survival, growth and morphological anomalies in farmed Atlantic salmon Salmo salar offspring. J Fish Biol 70: 462–73.

Eriksson M, Brown WT, Gordon LB, Glynn MW, Singer J, et al. (2003) Recurrent de novo point mutations in laminA cause Hutchinson-Gilford progeria syndrome. Nature 423: 293–8.

Eriksson O (1996) Regional dynamics of plants: A review of evidence for remnant, source–sink and metapopulations. Oikos 77: 248–58.

Eriksson O, Ehrlén J (2001) Landscape fragmentation and the viability of plant populations. In: Silvertown J, Antonovics J, eds. Integrating ecology and evolution in a spatial context. Oxford, UK: Blackwell Publishing.  pp 157–175.

Erill I, Campoy S, Barbe J (2007) Aeons of distress: an evolutionary perspective on the bacterial SOS response. FEMS Microbiol Rev 31: 637-56.

Erkkilä K, Henriksen K, Hirvonen V, Rannikko S, Salo J, et al. (1997) Testosterone regulates apoptosis in adult human seminiferous tubules in vitro. J Clin Endocrinol Metab 82: 2314–21.

Erkkilä K, Hirvonen V, Wuokko E, Parvinen M, Dunkel L (1998) N-acetyl-L-cysteine inhibits apoptosis in human male germ cells in vitro. J Clin Endocrinol Metab 83: 2523-31.

Erkkilä K, Pentikäinen V, Wikström M, Parvinen M, Dunkel L (1999) Partial oxygen pressure and mitochondrial permeability transition affect germ cell apoptosis in the human testis. J Clin Endocrinol Metab 84: 4253-9.

Erkkilä K, Aito H, Aalto K, Pentikäinen V, Dunkel L (2002) Lactate inhibits germ cell apoptosis in the human testis. Mol Hum Reprod 8: 109-17.

Erkkila K, Kyttanen S, Wikstrom M, Taari K, Sinha Hikim AP, et al. (2006) Regulation of human male germ cell death by modulators of ATP production. Am J Physiol Endocrinol Metab 290: E1145–E1154.

Erlandsson R, Wilson JF, Paabo S (2000) Sex chromosomal transposable element accumulation and male-driven substitutional evolution in humans. Mol Biol Evol 17: 804-12.

Erwin DH (2005) Development, ecology, and environment in the Cambrian metazoan radiation. Proc Calif Acad Sci 56 (Suppl. 1): 24-31.

Erwin D (2009) Early origin of the bilaterian developmental toolkit. Philos Trans R Soc B Biol Sci 364: 2253-61.

Escarceller M, Hicks J, Gudmundsson G, Trump G, Touati D, et al. (1994) Involvement of Escherichia coli DNA polymerase II in response to oxidative damage and adaptive mutation. J Bacteriol 176: 6221-8.

Escarmís C, Dávila M, Charpentier N, Bracho A, Moya A, Domingo E (1996) Genetic lesions associated with Muller’s ratchet in an RNA virus. J Mol Biol 264: 255–67.

Escarmís C, Gómez-Mariano G, Dávila M, Lázaro E, Domingo E (2002) Resistance to extinction of low fitness virus subjected to plaque-to-plaque transfers: diversification by mutation clustering. J Mol Biol 315: 647–61.

Escarmís C, Lázaro E, Manrubia SC (2006) Population bottlenecks in quasispecies dynamics. Curr Top Microbiol Immunol 299: 141–70.

Escarmís C, Perales C, Domingo E (2009) Biological effect of Muller’s ratchet: distant capsid site can affect picornavirus protein processing. J Virol 83: 6748–56.

Eschel I (1973a) Clone selection and the evolution of modifying features. Theor Popul Biol 4: 196–208.

Eschel I (1973b) Clone-selection and optimal rates of mutation. J Appl Probability 10: 728–38.

Eshel I, Feldman MW (1970) On the evolutionary effect of recombination. Theor Popul Biol 1: 88–100.

Eshelman JR, Markowitz SD (1996) Mismatch repair defects in human carcinogenesis. Hum Mol Gen 5: 1489-94.

ESHRE Capri Workshop Group (2005) Fertility and ageing. Hum Reprod Update 11: 261–76.

Espino J, Bejarano I, Ortiz A, Lozano GM, García JF, et al. (2010) Melatonin as a potential tool against oxidative damage and apoptosis in ejaculated human spermatozoa. Fertil Steril 94: 1915–7.

Essajee SM, Pollack H, Rochford G, Oransky I, Krasinski K, Borkowsky W (2000) Early changes in quasispecies repertoire in HIV-infected infants: Correlation with disease progression. Aids Res Hum Retrov 16: 1949–57.

Esser C, Alberti S, Höhfeld J (2004) Cooperation of molecular chaperones with the ubiquitin/proteasome system. Biochim Biophys Acta 1695: 171-88.

Estes S, Lynch M (2003) Rapid fitness recovery in mutationally degraded lines of Caenorhabditis elegans. Evolution 57: 1022–30.

Estes S, Arnold SJ (2007) Resolving the paradox of stasis: models with stabilizing selection explain evolutionary divergence on all timescales. Am Nat 169: 227–44.

Estes S, Teotónio H (2009) The experimental study of reverse evolution. In: Garland T, Rose MR, eds. Experimental Evolution: Concepts, Methods, and Applications of Selection Experiments. Berkeley, CA: University of California Press. pp 135–155.

Evan G, Harrington E, Fanidi A, Land H, Amati B, Bennett M (1994) Integrated control of cell proliferation and cell death by the c-myc oncogene. Philos Trans R Soc Lond B Biol Sci 345: 269-75.

Evan GI, Vousden KH (2001) Proliferation, cell cycle and apoptosis in cancer. Nature 411: 342–8.

Evans AR, Limp-Foster M, Kelley MR (2000) Going APE over Ref-1. Mutat Res 461: 83–108.

Evans DA, Paddock EF (1979) Mitotic crossing-over in plants. In: Sharp WR, Larsen PO, Paddock EF, Raghavan V, eds. Plant Cell and Tissue Culture, Principles and Applications. Columbus, OH: Ohio State University Press. pp 315–351.

Evans JP, Box TM, Brooshooft P, Tatler JR, Fitzpatrick JL (2010) Females increase egg deposition in favor of large males in the rainbowfish, Melanotaenia australis. Behav Ecol 21: 465–9.

Evans MD, Cooke MS (2004) Factors contributing to the outcome of oxidative damage to nucleic acids. Bioessays 26: 533–42.

Evans ME, Dennehy JJ (2005)Germ banking: bet-hedging and variable release from egg and seed dormancy.Q Rev Biol 80: 431-51.

Evans ME, Ferrière R, Kane MJ, Venable DL (2007) Bet hedging via seed banking in desert evening primroses (Oenothera, Onagraceae): demographic evidence from natural populations. Am Nat 169: 184-94.

Evans MS, Stewart JA (1977) Epibenthic and benthic microcrustaceans (copepods, cladocerans, ostracods) from a nearshore area in Southeastern Lake Michigan. Limnol Oceanogr 22: 1059–66.

Evans PJ, Whiteman M, Tredger JM, Halliwell B (1997) Antioxidant properties of S-adenosyl-L-methionine: a proposed addition to organ storage fluids. Free Radic Biol Med 23: 1002-8.

Evenson DP, Larson KL, Jost LK (2002) Sperm chromatin structure assay: its clinical use for detecting sperm DNA fragmentation in male infertility and comparisons with other techniques. J Androl 23: 25–43.

Evgenev MB, Zelentsova H, Shostak N, Kozitsina M, Barskyi V, et al. (1997) Penelope, a new family of transposable elements and its possible role in hybrid dysgenesis in Drosophila virilis. Proc Natl Acad Sci USA 94: 196–201.

Ewen-Kampen B, Schwager EE, Extavour CGM (2010) The molecular machinery of germ line specification. Mol Reprod Dev 77: 3-18.

Ewing JF, Maines MD (1995) Distribution of constitutive (HO-2) and heat-inducible (HO-1) heme oxygenase isozymes in rat testes: HO-2 displays stage-specific expression in germ cells. Endocrinology 136: 2294-302.

Excoffier L (1990) Evolution of human mitochondrial DNA: evidence for departure from a pure neutral model of populations at equilibrium. J Mol Evol 30: 125-39.

Extavour CG (2007) Evolution of the bilaterian germ line: lineage origin and modulation of specification mechanisms. Integr Comp Biol 47: 770-85.

Extavour C, García-Bellido A (2001) Germ cell selection in genetic mosaics in Drosophila melanogaster. Proc Natl Acad Sci USA 98: 11341–6.

Extavour CG, Akam M (2003) Mechanisms of germ cell specification across the metazoans: epigenesis and preformation. Development 130: 5869–84.

Eyre-Walker A (2002) Changing effective population size and the McDonald–Kreitman test. Genetics 162: 2017–24.

Eyre-Walker A (2006) The genomic rate of adaptive evolution. Trends Ecol Evol 21: 569–75.

Eyre-Walker A, Keightley PD (1999) High genomic deleterious mutation rates in hominids. Nature 397: 344–7.

Eyre-Walker A, Keightley PD, Smith NGC, Gaffney D (2002) Quantifying the slightly deleterious mutation model of molecular evolution. Mol Biol Evol 19: 2142–9.

Eyre-Walker A, Keightley PD (2007) The distribution of fitness effects of new mutations. Nat Rev Genet 8: 610–8.

Ezard TH, Aze T, Pearson PN, Purvis A (2011) Interplay between changing climate and species' ecology drives macroevolutionary dynamics. Science 332: 349-51.

Ezashi T, Das P, Roberts RM (2005) Low O2 tensions and the prevention of differentiation of hES cells. Proc Natl Acad Sci USA 102: 4783–8.

Fabbri A, Tinajero JC, Dufau ML (1990) Corticotropin-releasing factor is produced by rat Leydig cells and has a major local antireproductive role in the testis. Endocrinology 127: 1541–3.

Fabbri M, Garzon R, Cimmino A, Liu Z, Zanesi N, et al. (2007) MicroRNA-29 family reverts aberrant methylation in lung cancer by targeting DNA methyltransferases 3A and 3B. Proc Natl Acad Sci USA 104: 15805–10.

Faber BM, Mercan R, Hamacher P, Muasher SJ, Toner JP (1997)The impact of an egg donor's age and her prior fertility on recipient pregnancy outcome.Fertil Steril 68: 370-2.

Fabisiewicz A, Janion C (1998) DNA mutagenesis and repair in UV-irradiated E-coli K-12 under condition of mutation frequency decline. Mutat Res 402: 59-66.

Fabre F, Boulet A, Roman H (1984) Gene conversion at different points in the mitotic cycle of Saccharomyces cerevisiae.Mol Gen Genet 195: 139–43.

Facundo HTF, Carreira RS, de Paula JG, Santos CCX, Ferranti R, et al. (2006) Ischemic preconditioning requires increases in reactive oxygen release independent of mitochondrial K+ channel activity. Free Radic Biol Med 40: 469–79.

Faddy MJ, Gosden RG, Gougeon A, Richardson SJ, Nelson JF (1992) Accelerated disappearance of ovarian follicles in mid-life: Implications for forecasting menopause. Hum Reprod 7: 1342–6.

Faddy MJ, Gosden RG, Oktay K, Nelson JF (1999) Factoring in complexity and oocyte memory - can transformations and cyperpathology distort reality? Fertil Steril 71: 1170–2.

Faes MR, Caldas-Bussiere MC, Viana KS, Dias BL, Costa FR, Escocard RM (2009) Nitric oxide regulates steroid synthesis by bovine antral granulosa cells in a chemically defined medium. Anim Reprod Sci 110: 222–36.

Fagerström T, Briscoe DA, Sunnucks P (1998) Evolution of mitotic cell-lineages in multicellular organisms. Trends Ecol Evol 13: 117–20.

Fahim MS, Messiha FS, Girgis SM (1980) Effect of acute and chronic simulated high altitude on male reproduction and testosterone level. Arch Androl 4: 217–9.

Falconer DS (1989) Introduction to Quantitative Genetics. New York, NY: Longmans.

Falconer DS, Mackay TFC (1996) Introduction to Quantitative Genetics, 4th edn. Essex, UK: Addison Wesley Longman.

Falk DA, Holsinger KE (1991) Genetics and Conservation of Rare Plants. New York, NY: Oxford University Press.

Falkowski PG (2006) Evolution: tracing oxygen’s imprint on earth’s metabolic evolution. Science 311: 1724–5.

Falkowski PG, Katz ME, Milligan AJ, Fennel K, Cramer BS, et al. (2005) The rise of oxygen over the past 205 million years and the evolution of large placental mammals. Science 309: 2202–4.

Falster DS, Westoby M (2003) Plant height and evolutionary games. Trends Ecol Evol 18: 337-43.

Fan T, Yan Q, Huang J, Austin S, Cho E, et al. (2003) Lsh-deficient murine embryonal fibroblasts show reduced proliferation with signs of abnormal mitosis. Cancer Res 63: 4677–83.

Fan W, Waymire KG, Narula N, Li P, Rocher C, et al. (2008) A mouse model of mitochondrial disease reveals germline selection against severe mtDNA mutations. Science 319:958-62.

Fankhauser S (2010) The costs of adaptation. Wiley Interdiscipl Rev Climate Change 1: 23–30.

Farci P, Purcell RH (2000) Clinical significance of hepatitis C virus genotypes and quasispecies. Semin Liver Dis 20: 103–26.

Farci P, Strazzera R, Alter HJ, Farci S, Degioannis D, et al. (2002) Early changes in hepatitis C viral quasispecies during interferon therapy predict the therapeutic outcome. Proc Natl Acad Sci USA 99: 3081–6.

Farías JG, Bustos-Obregón E, Reyes JG (2005) Increase in testicular temperature and vascularization induced by hypobaric hypoxia in rats. J Androl 26: 693-7.

Farkash EA, Kao GD, Horman SR, Prak ETL (2006) Gamma radiation increases endonuclease-dependent L1 retrotransposition in a cultured cell assay. Nucleic Acids Res 34: 1196–204.

Farrelly FW, Finkelstein DB (1984) Complete sequence of the heat shock-inducible HSP90 gene of Saccharomyces cerevisiae. J Biol Chem 359: 5745-51.

Fasanaro P, D’Alessandra Y, Di Stefano V, Melchionna R, Romani S, et al. (2008) MicroRNA-210 modulates endothelial cell response to hypoxia and inhibits the receptor tyrosine-kinase ligand Ephrin-A3. J Biol Chem 283: 15878–83.

Fasanaro P, Greco S, Lorenzi M, Pescatori M, Brioschi M (2009) An integrated approach for experimental target identification of hypoxia-induced miR-210. J Biol Chem 284: 35134–43.

Fatehi AN, Bevers MM, Schoevers E, Roelen BA, Colenbrander B, Gadella BM (2006) DNA damage in bovine sperm does not block fertilization and early embryonic development but induces apoptosis after the first cleavages. J Androl 27: 176–88.

Faulk C, Barks A, Dolinoy DC (2013) Phylogenetic and DNA methylation analysis reveal novel regions of variable methylation in the mouse IAP class of transposons. BMC Genomics 14: 48.

Faulk EA, McCully JD, Tsukube T, Hadlow NC, Krukenkamp IB, Levitsky S (1995) Myocardial mitochondrial calcium accumulation modulates nuclear calcium accumulation and DNA fragmentation. Ann Thor Surg 60: 338-44.

Faulkner GJ, Kimura Y, Daub CO,Wani S, Plessy C, et al. (2009) The regulated retrotransposon transcriptome of mammalian cells. Nat Genet 41: 563–71.

Favaro E, Ramachandran A, McCormick R, Gee H, Blancher C, et al. (2010) MicroRNA-210 regulates mitochondrial free radical response to hypoxia and Krebs cycle in cancer cells by targeting iron sulfur cluster protein ISCU. PLoS One 5: e10345.

Favre M, Sornette D (2012) Strong gender differences in reproductive success variance, and the times to the most recent common ancestors. J Theor Biol 310: 43-54.

Fawcett DW, Eddy EM, Phillips DM (1970) Observations on the fine structure and relationships of the chromatoid body in mammalian spermatogenesis. Biol Reprod 2: 129–53.

Fawcett DW, Neaves WB, Flores MN (1973) Comparative observations on interstitial lymphatics and organization of the interstitial tissue of the mammalian testis. Biol Reprod 9: 500-32.

Fawcett JA, Van de Peer Y (2010) Angiosperm polyploids and their road to evolutionary success. Trends Evol Biol 2: e3.

Fay JC, Wyckoff GJ, Wu C-I (2001) Positive and negative selection on the human genome. Genetics 158: 1227–34.

Fay JC, Wyckoff GJ, Wu C-I (2002) Testing the neutral theory of molecular evolution with genomic data from Drosophila. Nature 415: 1024–6.

Feder ME, Hofmann GE (1999) Heat-shock proteins, molecular chaperones, and the stress response: Evolutionary and ecological physiology. Annu Rev Physiol 61: 243–82.

Fedoroff N (2006) Redox regulatory mechanisms in cellular stress responses. Ann Bot 98: 289-300.

Fedorov Y, Anderson EM, Birmingham A, Reynolds A, Karpilow J, et al. (2006) Off-target effects by siRNA can induce toxic phenotype. RNA 12: 1188-96.

Fehér T, Bogos B, Méhi O, Fekete G, Csörgo B, et al. (2012) Competition between transposable elements and mutator genes in bacteria. Mol Biol Evol 29: 3153-9.

Fehrer C, Brunauer R, Laschober G, Unterluggauer H, Reitinger S, et al. (2007) Reduced oxygen tension attenuates differentiation capacity of human mesenchymal stem cells and prolongs their lifespan. Aging Cell 6: 745–57.

Feig DI, Reid TM, Loeb LA (1994) Reactive oxygen species in tumorigenesis. Cancer Res 54(7 Suppl): 1890s-1894s.

Feig LA, Bellve AR, Erickson NH, Klagsbrun M (1980) Sertoli cells contain a mitogenic polypeptide. Proc Natl Acad Sci USA 77: 4774–8.

Feigl H (1970) The 'orthodox' view of theories: Remarks in defense as well as critique. In: Radner M, Winokur S, eds. Minnesota Studies in the Philosophy of Science, Volume IV. Minneapolis: University of Minnesota Press.

Feil EJ, Maiden MC, Achtman M, Spratt BG (1999) The relative contributions of recombination and mutation to the divergence of clones of Neisseria meningitidis. Mol Biol Evol 16: 1496-502.

Feinberg AP (2007) Phenotypic plasticity and the epigenetics of human disease. Nature 447: 433–40.

Feinberg AP (2008) Epigenetics at the epicenter of modern medicine. JAMA 299: 1345–50.

Feinberg AP, Irizarry RA (2010) Stochastic epigenetic variation as a driving force of development, evolutionary adaptation, and disease. Proc Natl Acad Sci USA 107: 1757–64.

Feinberg EH, Hunter CP (2003) Transport of dsRNA into cells by the transmembrane protein SID-1. Science 301: 1545–7.

Feissner RF, Skalska J, Gaum WE, Sheu SS (2009) Crosstalk signaling between mitochondrial Ca2+ and ROS. Front Biosci 14: 1197-218.

Feitsma H, Leal MC, Moens PB, Cuppen E, Schulz RW (2007) Mlh1 deficiency in zebrafish results in male sterility and aneuploid as well as triploid progeny in females. Genetics 175: 1561–9.

Feldman MW (1972) Selection for linkage modification. I. Random mating populations. Theor Popul Biol 21: 430–9.

Feldman MW, Lewontin RC (1975) The heritability hang-up. Science 190: 1163-8.

Feldman MW, Christiansen FB, Brooks LD (1980) Evolution of recombination in a constant environment. Proc Natl Acad Sci USA 77: 4838–41.

Feldman MW, Liberman U (1986) An evolutionary reduction principle for genetic modifiers. Proc Natl Acad Sci USA 83: 4824–7.

Feldman MW, Otto SP, Christiansen FB (1997) Population genetic perspectives on the evolution of recombination. Annu Rev Genet 30: 261–95.

Feldser DM, Kostova KK, Winslow MM, Taylor SE, Cashman C, et al. (2010) Stage-specific sensitivity to p53 restoration during lung cancer progression. Nature 468: 572-5.

Felippes FF, Weigel D (2009) Triggering the formation of tasiRNAs in Arabidopsis thaliana: the role of microRNA miR173. EMBO Rep 10: 264-70.

Félix M-A, Wagner A (2008) Robustness and evolution: concepts, insights and challenges from a developmental model system. Heredity 100: 132–40.

Felsenstein J (1974) The evolutionary advantage of recombination. Genetics 78: 737-56.

Felsher DW, Bishop JM (1999) Reversible tumorigenesis by MYC in hematopoietic lineages. Mol Cell 4: 199–207.

Felty Q, Xiong WC, Sun D, Sarkar S, Singh KP, et al. (2005) Estrogen-induced mitochondrial reactive oxygen species as signal-transducing messengers. Biochemistry 44: 6900-9.

Feng G, Tsui HC, Winkler ME (1996) Depletion of the cellular amounts of the MutS and MutH methyl-directed mismatch repair proteins in stationary-phase Escherichia coli K-12 cells. J Bacteriol 178: 2388–96.

Feng G, Leem YE, Levin HL (2013) Transposon integration enhances expression of stress response genes. Nucleic Acids Res 41: 775-89.

Feng L, Xia Y, Garcia GE, Hwang D, Wilson CB (1995) Involvement of reactive oxygen intermediates in cyclooxygenase-2 expression induced by interleukin-1, tumor necrosis factor-alpha, and lipopolysaccharide. J Clin Invest 95: 1669-75.

Feng S, Jacobsen SE, Reik W (2010) Epigenetic reprogramming in plant and animal development. Science 330: 622–7.

Feng Z, Levine AJ (2010) The regulation of energy metabolism and the IGF-1/mTOR pathways by the p53 protein. Trends Cell Biol 20: 427-34.

Feng Z, Lin M, Wu R (2011) The regulation of aging and longevity: a new and complex role of p53. Genes Cancer 2: 443-52.

Ferguson-Smith AC, Patti ME (2011) You are what your dad ate. Cell Metab 13: 115-7.

Fernández J, Plastino A, Diambra L, Mostaccio C (1998) Dynamics of coevolutive processes. Phys Rev E 57: 5897–903.

Fernandez V, Barrientos X, Kipreos K, Valenzuela A, Videla LA (1985) Superoxide radical generation, NADPH oxidase activity, and cytochrome P-450 content of rat liver microsomal fractions in an experimental hyperthyroid state: relation to lipid peroxidation. Endocrinology 117: 496–501.

Fernandez V, Tapia G, Varela P, Romanque P, Cartier-Ugarte D, Videla LA (2005) Thyroid hormone-induced oxidative stress in rodents and humans: A comparative view and relation to redox regulation of gene expression. Comp Biochem Physiol C Toxicol Pharmacol 142: 231-9.

Fernàndez-Busquets X, Körnig A, Bucior I, Burger MM, Anselmetti D (2009) Self-recognition and Ca2+-dependent carbohydrate-carbohydrate cell adhesion provide clues to the Cambrian explosion. Mol Biol Evol 26: 2551-61.

Fernandez-Capetillo O, Lee A, Nussenzweig M, Nussenzweig A (2004) H2AX: the histone guardian of the genome. DNA Repair 3: 959–67.

Fernández-Sánchez A, Madrigal-Santillán E, Bautista M, Esquivel-Soto J, Morales-González A, et al. (2011) Inflammation, oxidative stress, and obesity. Int J Mol Sci 12: 3117-32.

Ferrante A, Nandoskar M, Bates EJ, Goh DH, Beard LJ (1988) Tumour necrosis factor beta (lymphotoxin) inhibits locomotion and stimulates the respiratory burst and degranulation of neutrophils. Immunology 63: 507-12.

Ferrari J, Müller CB, Kraaijeveld AR, Godfray HCJ (2001) Clonal variation and covariation in aphid resistance to parasitoids and a pathogen. Evolution 55: 1805–14.

Ferrarini M, Heltai S, Zocchi MR, Rugarli C (1992) Unusual expression and localization of heat-shock proteins in human tumor cells. Int J Cancer 51: 613–9.

Ferreira C (2002) Mutation, transposition, and recombination: an analysis of the evolutionary dynamics. In: Caulfield HJ, Chen S-H, Cheng H-D, et al., eds. Proceedings of the 6th Joint Conference on Information Sciences, 4th International Workshop on Frontiers in Evolutionary Algorithms. Research Triangle Park, NC, USA. pp 614-617.

Ferrell JE Jr (2002) Self-perpetuating states in signal transduction: positive feedback, double-negative feedback and bistability. Curr Opin Cell Biol 14: 140–8.

Ferri D, Mazzone A, Liquori GE, Cassano G, Svelto M, Calamita G (2003) Ontogeny, distribution, and possible functional implications of an unusual aquaporin, AQP8, in mouse liver. Hepatology 38: 947–57.

Ferri KF, Kroemer G (2001) Organelle-specific initiation of cell death pathways. Nat Cell Biol 3: E255–E263.

Ferris PJ, Pavlovic C, Fabry S, Goodenough UW (1997) Rapid evolution of sex-related genes in Chlamydomonas. Proc Natl Acad Sci USA 94: 8634–9.

Feschotte C (2008) Transposable elements and the evolution of regulatory networks. Nat Rev Genet 9: 397-405.

Feschotte C, Pritham EJ (2007) DNA transposons and the evolution of eukaryotic genomes. Annu Rev Genet 41: 331–68.

Festa-Bianchet M, Jorgenson JT (1998) Selfish mothers: reproductive expenditure and resource availability in bighorn ewes. Behav Ecol 9: 144-50.

Ficz G, Branco MR, Seisenberger S, Santos F, Krueger F, et al. (2011) Dynamic regulation of 5-hydroxymethylcytosine in mouse ES cells and during differentiation. Nature 473: 398-402.

Fieldes MA, Amyot LM (1999) Epigenetic control of early flowering in flax lines induced by 5-azacytidine applied to germinating seed. J Hered 90: 199–206.

Fiers W (1991) Tumor necrosis factor. Characterization at the molecular, cellular and in vivo level. FEBS Lett 285: 199-212.

Fierst JL (2011) Sexual dimorphism increases evolvability in a genetic regulatory network. Evol Biol 38: 52–67.

Fijalkowska IJ, Dunn RL, Schaaper RM (1993) Mutants of Escherichia coli with increased fidelity of DNA replication. Genetics 134: 1023–30.

Fike DA, Grotzinger JP, Pratt LM, Summons RE (2006) Oxidation of the Ediacaran ocean. Nature 444: 744–7.

Filatov DA, Charlesworth D (2002) Substitution rates in the X-linked and Y-linked genes of the plants, Silene latifolia and S. dioica. Mol Biol Evol 19: 898–907.

Filkowski J, Yeoman A, Kovalchuk O, Kovalchuk I (2004) Systemic plant signal triggers genome instability. Plant J 38: 1-11.

Filkowski JN, Ilnytskyy Y, Tamminga J, Koturbash I, Golubov A, et al. (2010) Hypomethylation and genome instability in the germline of exposed parents and their progeny is associated with altered miRNA expression. Carcinogenesis 31: 1110-5.

Finch CE, Holmes DJ (2010) Ovarian aging in developmental and evolutionary contexts. Ann NY Acad Sci 1204: 82-94.

Findlay S, Rowe G (1990) Computer experiments on the evolution of sex: the haploid case. J Theor Biol 146: 379-93.

Finkel SE, Kolter R (1999) Evolution of microbial diversity during prolonged starvation. Proc Natl Acad Sci USA 96: 4023–7.

Finkel SE, Kolter R (2001) DNA as a nutrient: novel role for bacterial competence gene homologs. J Bacteriol 183: 6288-93.

Finkel T (1998) Oxygen radicals and signaling. Curr Opin Cell Biol 10: 248-53.

Finkel T, Holbrook NJ (2000) Oxidants, oxidative stress and the biology of ageing. Nature 408: 239–47.

Finlay BJ, Fenchel T (2004) Cosmopolitan metapopulations of free-living microbial eukaryotes. Protist 155: 237–44.

Finnegan EJ (2002) Epialleles - a source of random variation in times of stress. Curr Opin Plant Biol 5: 101-6.

Fiorenza MT, Farkas T, Dissing M, Kolding D, Zimarino V (1995) Complex expression of murine heat shock transcription factors. Nucleic Acids Res 23: 467–74.

Fire A, Xu S, Montgomery MK, Kostas SA, Driver SE, Mello CC (1998) Potent and specific genetic interference by double-stranded RNA in Caenorhabditis elegans. Nature 391: 806–11.

Firman RC, Simmons LW (2010) Experimental evolution of sperm quality via postcopulatory sexual selection in house mice. Evolution 64: 1245–56.

Firth JD, Ebert BL, Pugh CW, Ratcliffe PJ (1994) Oxygen-regulated control elements in the phosphoglycerate kinase 1 and lactate dehydrogenase A genes: similarities with the erythropoietin 3’ enhancer. Proc Natl Acad Sci USA 91: 6496–500.

Firth JD, Ebert BL, Ratcliffe PJ (1995) Hypoxic regulation of lactate dehydrogenase A: interaction between hypoxia inducible factor 1 and cAMP response elements. J Biol Chem 270: 21021–7.

Fischer AG (1960) Latitudinal variations in organic diversity. Evolution 14: 64-81.

Fischer B, Kunzel W, Kleinstein J, Gips H (1992) Oxygen tension in follicular fluid falls with follicle maturation. Eur J Obstet Gynecol Reprod Biol 43: 39-43.

Fischer BM, Schatz H, Maraun M (2010) Community structure, trophic position and reproductive mode of soil and bark-living oribatid mites in an alpine grassland ecosystem. Exp Appl Acarol 52: 221-37.

Fischer O, Schmid-Hempel P (2005) Selection by parasites may increase host recombination frequency. Biol Lett 1: 193–5.

Fisher CR, Graves KH, Parlow A, Simpson ER (1998) Characterization of mice deficient in aromatase (ArKO) because of targeted disruption of the cyp 19 gene. Proc Natl Acad Sci USA 95: 6965–70.

Fisher DS (2011) Leading the dog of selection by its mutational nose. Proc Natl Acad Sci USA 108: 2633-4.

Fisher DO, Double MC, Blomberg SP, Jennions MD, Cockburn A (2006) Postmating sexual selection increases lifetime fitness of polyandrous females in the wild. Nature 444: 89–92.

Fisher E (2007) Assessing the health of coral reef ecosystems in the Florida Keys at community, individual and cellular scales. Thesis. University of South Florida.

Fisher HM, Aitken RJ (1997) Comparative analysis of the ability of precursor germ cells and epididymal spermatozoa to generate reactive oxygen metabolites. J Exp Zool 277: 390-400.

Fisher HS, Hoekstra HE (2010) Competition drives cooperation among closely related sperm of deer mice. Nature 463: 801-3.

Fisher RA (1930) The genetical theory of natural selection. Oxford, UK: Clarendon Press.

Fisher RA (1958) The genetical theory of natural selection. 2nd edn. Oxford, UK: Oxford University Press.

Fisk D, Latta L, Knapp R, Pfrender M (2007) Rapid evolution in response to introduced predators I: rates and patterns of morphological and life-history trait divergence. BMC Evol Biol 7: 22.

Fitch WM, Leiter JME, Li XQ, Palese P (1991) Positive Darwinian evolution in human influenza A viruses. Proc Natl Acad Sci USA 88: 4270-4.

Fitter AH (1986) Acquisition and utilization of resources. In: Crawley MJ, ed. Plant ecology. Oxford, UK: Blackwell. pp 376-406.

Fitzpatrick DA (2012) Horizontal gene transfer in fungi. FEMS Microbiol Lett 329: 1–8.

FitzPatrick DR, Ramsay J, McGill NI, Shade M, Carothers AD, Hastie ND (2002) Transcriptome analysis of human autosomal trisomy. Hum Mol Genet 11: 3249–56.

Fitzpatrick JL, Evans JP (2009) Reduced heterozygosity impairs sperm quality in endangered mammals. Biol Lett 5: 320–3.

Fitzpatrick JL, Montgomerie R, Desjardins JK, Stiver KA, Kolm N, et al. (2009) Female promiscuity promotes the evolution of faster sperm in cichlid fishes. Proc Natl Acad Sci USA 106: 1128–32.

Fitzsimmons JM, Innes DJ (2006) Inter-genotype variation in reproductive response to crowding among Daphnia pulex. Hydrobiologia 568: 187-205.

Fix DF, Glickman BW (1987) Asymmetric cytosine deamination revealed by spontaneous mutational specificity in an Ung strain of Escherichia coli. Mol Gen Genet 209: 78–82.

Fix D, Canugovi C, Bhagwat AS (2008) Transcription increases methylmethane sulfonate-induced mutations in alkB strains of Escherichia coli. DNA Repair (Amst) 7: 1289–97.

Flatscher R, Frajman B, Schönswetter P, Paun O (2012) Environmental heterogeneity and phenotypic divergence: can heritable epigenetic variation aid speciation? Genet Res Int 2012: 698421.

Flatt T (2005) The evolutionary genetics of canalization. Q Rev Biol 80: 287–316.

Fleming I, Michaelis UR, Bredenkotter D, Fisslthaler B, Dehghani F, et al. (2001) Endothelium-derived hyperpolarizing factor synthase (Cytochrome P450 2C9) is a functionally significant source of reactive oxygen species in coronary arteries. Circ Res 88: 44–51.

Fleming TP, Javed Q, Hay M (1992) Epithelial differentiation and intercellular junction formation in the mouse early embryo. Development (Suppl) 17: 105–12.

Flessa KW, Jablonski D (1996) The geography of evolutionary turnover: a global analysis of extant bivalves. In: Jablonski D, Erwin DH, Lipps JH, eds. Evolutionary paleobiology. Chicago, IL: University of Chicago Press. pp 376–397.

Flodman P, Hodge SE (2003) Sex-specific mutation rates for X-linked disorders: estimation and application. Hum Hered 55: 51–5.

Flohr C, Burkle A, Radicella JP, Epe B (2003) Poly(ADP-ribosyl)ation accelerates DNA repair in a pathway dependent on Cockayne syndrome B protein. Nucleic Acids Res 31: 5332–7.

Flores I, Murphy DJ, Swigart LB, Knies U, Evan GI (2004) Defining the temporal requirements for Myc in the progression and maintenance of skin neoplasia. Oncogene 23: 5923–30.

Flowers JM, Li SI, Stathos A, Saxer G, Ostrowski EA, et al. (2010) Variation, sex, and social cooperation: molecular population genetics of the social amoeba Dictyostelium discoideum. PLoS Genet 6: e1001013.

Flynn KM, Cooper TF, Moore FB, Cooper VS (2013) The environment affects epistatic interactions to alter the topology of an empirical fitness landscape. PLoS Genet 9: e1003426.

Foekens JA, Sieuwerts AM, Smid M, Look MP, de Weerd V, et al. (2008) Four miRNAs associated with aggressiveness of lymph node-negative, estrogen receptor-positive human breast cancer. Proc Natl Acad Sci USA 105: 13021-6.

Foellmer MW (2008) Broken genitals function as mating plugs and affect sex ratios in the orb-web spider Argiope aurantia. Evol Ecol Res 10: 449–62.

Fogle CA, Nagle JL, Desai MM (2008) Clonal interference, multiple mutations and adaptation in large asexual populations. Genetics 180:2163-73.

Foksinski M, Rozalski R, Guz J, Ruszkowska B, Sztukowska P, et al. (2004) Urinary excretion of DNA repair products correlates with metabolic rates as well as with maximum life spans of different mammalian species. Free Radic Biol Med 37:1449-54.

Folt CL, Chen CY, Moore MV, Burnaford J (1999) Synergism and antagonism among multiple stressors. Limnol Oceanogr 44: 864–77.

Fondon JW 3rd, Garner HR (2004) Molecular origins of rapid and continuous morphological evolution. Proc Natl Acad Sci USA 101: 18058–63.

Fondon JW 3rd, Hammock EA, Hannan AJ, King DG (2008) Simple sequence repeats: genetic modulators of brain function and behavior. Trends Neurosci 31: 328-34.

Fong YW, Inouye C, Yamaguchi T, Cattoglio C, Grubisic I, Tjian R (2011) A DNA repair complex functions as an Oct4/Sox2 coactivator in embryonic stem cells. Cell 147: 120–31.

Fontaine KM, Cooley JR, Simon C (2007) Evidence for paternal leakage in hybrid periodical cicadas (Hemiptera: Magicicada spp.). PLoS ONE 2: e892.

Fontana W, Schuster P (1998) Continuity in evolution: On the nature of transitions. Science 280: 1451–5.

Fontanari JF, Colato A, Howard RS (2003) Mutation accumulation in growing asexual lineages. Phys Rev Lett 91: 218101.

Fontaneto D, Herniou EA, Boschetti C, Caprioli M, Melone G, et al. (2007) Independently evolving species in asexual bdelloid rotifers. PLoS Biol 5: e87.

Fontaneto D, Boschetti C, Ricci C (2008) Cryptic diversification in ancient asexuals: evidence from the bdelloid rotifer Philodina flaviceps. J Evol Biol 21: 580–7.

Fontaneto D, Kaya M, Herniou EA, Barraclough TG (2009) Extreme levels of hidden diversity in microscopic animals (Rotifera) revealed by DNA taxonomy. Mol Phylogenet Evol 53: 182-9.

Fontecave M (2006) Iron-sulfur clusters: ever-expanding roles. Nat Chem Biol 2: 171–4.

Fonville NC, Ward RM, Mittelman D (2011) Stress-induced modulators of repeat instability and genome evolution. J Mol Microbiol Biotechnol 21: 36-44.

Forabosco A, Sforza C, De Pol A, Vizzotto L, Marzona L, Ferrario VF (1991) Morphometric study of the human neonatal ovary. Anat Rec 231: 201–8.

Forand A, Dutrillaux B, Bernardino-Sgherri J (2004) Gamma-H2AX expression pattern in non-irradiated neonatal mouse germ cells and after low-dose gamma-radiation: relationships between chromatid breaks and DNA double-strand breaks. Biol Reprod 71: 643-9.

Forbes LS (1997) The evolutionary biology of spontaneous abortion in humans. Trends Ecol Evol 12: 446–50.

Forche A, Alby K, Schaefer D, Johnson AD, Berman J, Bennett RJ (2008) The parasexual cycle in Candida albicans provides an alternative pathway to meiosis for the formation of recombinant strains. PLoS Biol 6: e110.

Forche A, Abbey D, Pisithkula T, Weinzierl MA, Ringstrom T, et al. (2011) Stress alters rates and types of loss of heterozygosity in Candida albicans. mBio 2: e00129-11.

Ford JJ, McCoard SA, Wise TH, Lunstra DD, Rohrer GA (2006) Genetic variation in sperm production. Soc Reprod Fertil Suppl 62: 99-112.

Ford WC (2004) Regulation of sperm function by reactive oxygen species. Hum Reprod Update 10: 387-99.

Forman HJ, Fukuto JM, Torres M (2004) Redox signaling: thiol chemistry defines which reactive oxygen and nitrogen species can act as second messengers. Am J Physiol Cell Physiol 287: C246-56.

Forman HJ, Maiorino M, Ursini F (2010) Signaling functions of reactive oxygen species. Biochemistry 49: 835-42.

Forsberg LA, Dannewitz J, Petersson E, Grahn M (2007) Influence of genetic dissimilarity in the reproductive success and mate choice of brown trout - females fishing for optimal MHC dissimilarity. J Evol Biol 20: 1859-69.

Fos M, Dominguez MA, Latorre A, Moya A (1990) Mitochondrial DNA evolution in experimental populations of Drosophila subobscura. Proc Natl Acad Sci USA 87: 4198-201.

Fossøy F, Johnsen A, Lifjeld JT (2008) Multiple genetic benefits of female promiscuity in a socially monogamous passerine. Evolution 62: 145–56.

Foster KR (2004) Diminishing returns in social evolution: the not-so-tragic commons. J Evol Biol 17: 1058-72.

Foster KR, Shaulsky G, Strassmann JE, Queller DC, Thompson CRL (2004) Pleiotropy as a mechanism to stabilize cooperation. Nature 431: 693-6.

Foster KR, Wenseleers T (2006) A general model for the evolution of mutualisms. J Evol Biol 19: 1283-93.

Foster PL (1997) Nonadaptive mutations occur on the F' episome during adaptive mutation conditions in Escherichia coli. J Bacteriol 179: 1550-4.

Foster PL (2005) Stress responses and genetic variation in bacteria. Mutat Res 569: 3–11.

Foster PL (2007) Stress-induced mutagenesis in bacteria. Crit Rev Biochem Molec Biol 42: 373-97.

Fourie AM, Peterson PA, Yang Y (2001) Characterization and regulation of the major histocompatibility complex-encoded proteins Hsp70-Hom and Hsp70-1/2. Cell Stress Chaperones 6: 282–95.

Fowler K, Partridge L (1989) A cost of mating in female fruit flies. Nature 338: 760–1.

Fox CW, McLennan L A, Mousseau TA (1995) Male body size affects female lifetime reproductive success in a seed beetle. Anim Behav 50: 281–4.

Fox CW, Rauter CM (2003) Bet-hedging and the evolution of multiple mating. Evol Ecol Res 5: 273–86.

Fox CW, Wagner JD, Cline S, Thomas FA, Messina FJ (2011) Rapid evolution of lifespan in a novel environment: sex-specific responses and underlying genetic architecture. Evol Biol 38: 182–96.

Fox JA, Dybdahl MF, Jokela J, Lively CM (1996) Genetic structure of coexisting sexual and clonal subpopulations in a freshwater snail (Potamopyrgus antipodarum). Evolution 50: 1541–8.

Fox PM, Vought VE, Hanazawa M, Lee MH, Maine EM, Schedl T (2011) Cyclin E and CDK-2 regulate proliferative cell fate and cell cycle progression in the C. elegans germline. Development 138: 2223–34.

Foy BD, Myles KM, Pierro DJ, Sanchez-Vargas I, Uhlírová M, et al. (2004) Development of a new Sindbis virus transducing system and its characterization in three Culicine mosquitoes and two Lepidopteran species. Insect Mol Biol 13: 89–100.

Foyer CH, Noctor G (2005) Oxidant and antioxidant signalling in plants: a re-evaluation of the concept of oxidative stress in a physiological context. Plant Cell Environ 28: 1056–71.

Frada M, Probert I, Allen MJ, Wilson WH, de Vargas C (2008) The ‘Cheshire Cat’ escape strategy of the coccolithophore Emiliania huxleyi in response to viral infection. Proc Natl Acad Sci USA 105: 15944–9.

Fraga CG, Shigenaga MK, Park JW, Degan P, Ames BN (1990) Oxidative damage to DNA during aging: 8-hydroxy-20-deoxyguanosine in rat organ DNA and urine. Proc Natl Acad Sci USA 87: 4533–7.

Frago LM, Cañón S, de la Rosa EJ, León Y, Varela-Nieto I (2003) Programmed cell death in the developing inner ear is balanced by nerve growth factor and insulin-like growth factor I. J Cell Sci 116: 475-86.

Fraile B, Martin R, De Miguel MP, Arenas MI, Bethencourt FR, et al. (1996) Light and electron microscopic immunohistochemical localization of protein gene product 9.5 and ubiquitin immunoreactivities in the human epididymis and vas deferens. Biol Reprod 55: 291–7.

Framson PE, Sage EH (2004) SPARC and tumor growth: where the seed meets the soil? J Cell Biochem 92: 679–90.

Francavilla S, Cordeschi G, Properzi G, Di Cicco L, Jannini EA, et al. (1991) Effect of thyroid hormone on the pre- and postnatal development of the rat testis. J Endocrinol 129: 35-42.

Francavilla S, D'Abrizio P, Rucci N, Silvano G, Properzi G, et al. (2000) Fas and Fas ligand expression in fetal and adult human testis with normal or deranged spermatogenesis. J Clin Endocrinol Metab 85: 2692-700.

Francavilla S, D'Abrizio P, Cordeschi G, Pelliccione F, Necozione S, et al. (2002) Fas expression correlates with human germ cell degeneration in meiotic and post-meiotic arrest of spermatogenesis. Mol Hum Reprod 8: 213–20.

Franchi LL, Mandl AM, Zuckerman S (1962) The development of the ovary and the process of oogenesis. In: Zuckerman S, ed. The Ovary. London, UK: Academic Press. pp 1-88.

Franchini LF, López-Leal R, Nasif S, Beati P, Gelman DM, et al. (2011) Convergent evolution of two mammalian neuronal enhancers by sequential exaptation of unrelated retroposons. Proc Natl Acad Sci USA 108: 15270–5.

Francia S, Michelini F, Saxena A, Tang D, de Hoon M, et al. (2012) Site-specific DICER and DROSHA RNA products control the DNA-damage response. Nature 488: 231-5.

Francino MP, Ochman H (2001) Deamination as the basis of strand-asymmetric evolution in transcribed Escherichia coli sequences. Mol Biol Evol 18: 1147–50.

Francis AP, Currie DJ (2003) A globally consistent richness–climate relationship for angiosperms. Am Nat 161: 523–36.

Francis D (1998) High frequency recombination during the sexual cycle of Dictyostelium discoideum. Genetics 148: 1829–32.

Francis DD, Meaney MJ (1999) Maternal care and the development of stress responses. Curr Opin Neurobiol 9: 128–34.

Francke U, Felsenstein J, Gartler SM, Migeon BR, Dancis J, et al. (1976) The occurrence of new mutants in the X-linked recessive Lesch-Nyhan disease. Am J Hum Genet 28: 123-37.

Franco R, Schoneveld O, Georgakilas AG, Panayiotidis MI (2008) Oxidative stress, DNA methylation and carcinogenesis. Cancer Lett 266: 6-11.

Franco Mdo C, Dantas AP, Akamine EH, Kawamoto EM, Fortes ZB, et al. (2002) Enhanced oxidative stress as a potential mechanism underlying the programming of hypertension in utero. J Cardiovasc Pharmacol 40: 501–9.

Frank SA (1995) Mutual policing and repression of competition in the evolution of cooperative groups. Nature 377: 520-2.

Frank SA (1996a) Models of parasite virulence. Q Rev Biol 71: 37-78.

Frank SA (1996b) Policing and group cohesion when resources vary. Anim Behav 52: 1163-9.

Frank SA (1998) Foundations of social evolution. Princeton, NJ: Princeton University Press.

Frank SA (2003a) Somatic mutation: early steps in cancer depend on tissue architecture. Curr Biol 13: R261–R263.

Frank SA (2003b) Perspective: repression of competition and the evolution of cooperation. Evolution 57: 693-705.

Frank SA (2009) Evolutionary foundations of cooperation and group cohesion. In: Levin SA, ed. Games, Groups, and the Global Good. Berlin, Germany: Springer-Verlag. pp 3–40.

Frank SA (2010) The trade-off between rate and yield in the design of microbial metabolism. J Evol Biol 23: 609–13.

Frank SA (2011) Natural selection. II. Developmental variability and evolutionary rate. J Evol Biol 24: 2310-20.

Frank SA, Swingland IR (1988) Sex ratio under conditional sex expression. J Theor Biol 135: 415-8.

Frank SA, Slatkin M (1990) Evolution in a variable environment. Am Nat 136: 244-60.

Frank SA, Hurst LD (1996) Mitochondria and male disease. Nature 383: 224.

Frank SA, Nowak MA (2004) Problems of somatic mutation and cancer. BioEssays 26: 291-9.

Frankenhuis MT, Wensing CJ (1979) Induction of spermatogenesis in the naturally cryptorchid pig. Fertil Steril 31: 428–33.

Frankham R (1980) Origin of genetic variation in selection lines. In: Robertson A, ed. Selection Experiments in Laboratory and Domestic Animals. Slough, UK: Commonwealth Agricultural Bureau. pp 56-68.

Frankham R (1995) Effective population size/adult population size ratios in wildlife: a review. Genet Res 66: 95–107.

Frankham R (1996) Relationship of genetic variation to population size in wildlife. Conserv Biol 10: 1500–8.

Frankham R (1997) Do island populations have less genetic variation than mainland populations? Heredity (Edinb) 78: 311-27.

Franklin IR, Frankham R (1998) How large must populations be to retain evolutionary potential? Anim Conserv 1: 69-73.

Franklin TB, Mansuy IM (2010) Epigenetic inheritance in mammals: Evidence for the impact of adverse environmental effects. Neurobiol Dis 39: 61–5.

Fraser AG, Kamath RS, Zipperlen P, Martinez-Campos M, Sohrmann M, Ahringer J (2000) Functional genomic analysis of C. elegans chromosome I by systematic RNA interference. Nature 408: 325–30.

Fraser D, Kaern M (2009) A chance at survival: gene expression noise and phenotypic diversification strategies. Mol Microbiol 71: 1333-40.

Fraser HB, Hirsh AE, Giaever G, Kumm J, Eisen MB (2004) Noise minimization in eukaryotic gene expression. PLoS Biol 2: e137.

Fraser HM, Wilson H, Rudge JS, Wiegand SJ (2005) Single injections of vascular endothelial growth factor trap block ovulation in the macaque and produce a prolonged, dose-related suppression of ovarian function. J Clin Endocrinol Metab 90: 1114–22.

Fraser LR, Quinn PJ (1981) A glycolytic product is obligatory for initiation of the sperm acrosome reaction and whiplash motility required for fertilization in the mouse. J Reprod Fertil 61: 25–35.

Fraterrigo JM, Rusak JA (2008) Disturbance-driven changes in the variability of ecological patterns and processes. Ecol Lett 11: 756-70.

Frederico LA, Kunkel TA, Shaw BR (1990) A sensitive genetic assay for the detection of cytosine deamination: determination of rate constants and the activation energy. Biochemistry 29: 2532–7.

Free M (1977) Blood supply to the testis and its role in local exchange and transport of hormones. In: Johnson A, Gomes W, eds. The Testis, vol. 4. New York, NY: Academic Press. pp 39–90.

Free MJ, Schluntz GA, Jaffe RA (1976) Respiratory gas tensions in tissues and fluids of the male rat reproductive tract. Biol Reprod 14:481-8.

Freed NE, Silander OK, Stecher B, Böhm A, Hardt WD, Ackermann M (2008) A simple screen to identify promoters conferring high levels of phenotypic noise. PLoS Genet 4: e1000307.

Freeman-Gallant CR, Meguerdichian M, Wheelwright NT, Sollecito SV (2003) Social pairing and female mating fidelity predicted by restriction fragment length polymorphism similarity at the major histocompatibility complex in a songbird. Mol Ecol 12: 3077-83.

Freeman-Gallant CR, Wheelwright NT, Meiklejohn KE, Sollecito SV (2006) Genetic similarity, extrapair paternity, and offspring quality in Savannah sparrows (Passerculus sandwichensis). Behav Ecol 17: 952–8.

Freilich S, Kreimer A, Borenstein E, Gophna U, Sharan R, et al. (2010) Decoupling environment-dependent and independent genetic robustness across bacterial species. PLoS Comput Biol 6: e1000690.

French LE, Hahne M, Viard I, Radlgruber G, Zanone R, et al. (1996) Fas and Fas ligand in embryos and adult mice: ligand expression in several immune-privileged tissues and coexpression in adult tissues characterized by apoptotic cell turnover. J Cell Biol 133: 335–43.

French NR, McBride R, Detmer J (1965) Fertility and population density of the black-tailed jackrabbit. J Wildlife Manage 29: 14-26.

French R, Messinger A (1994) Genes, phenes and the Baldwin effect: Learning and evolution in a simulated population. Artif Life 4: 277–82.

Fretz PC, Sandlow JI (2002) Varicocele: current concepts in pathophysiology, diagnosis, and treatment. Urol Clin North Am 29: 921–38.

Fréville H, Silvertown J (2005) Analysis of interspecific competition in perennial plants using life table response experiments. Plant Ecol 176: 69-78.

Fricke C, Arnqvist G (2007) Rapid adaptation to a novel host in a seed beetle (Callosobruchus maculatus): The role of sexual selection. Evolution 61: 440–54.

Fridolfsson AK, Ellegren H (2000) Molecular evolution of the avian CHD1 genes on the Z and W sex chromosomes. Genetics 155: 1903–12.

Friedberg EC (2005) Suffering in silence: The tolerance of DNA damage. Nat Rev Mol Cell Biol 6: 943-53.

Friedberg EC (2006) Mutagenesis and translesion synthesis in prokaryotes. In: Friedberg EC, Walker GC, Siede W, Wood RD, Schultz RA, Ellenberger T, eds. DNA Repair and Mutagenesis. Washington, DC: ASM Press. pp 407–522.

Friedberg EC, Walker GC, Siede W (1995) DNA repair and mutagenesis. New York, NY: Academic Press.

Friedberg EC, Wagner R, Radman M (2002) Specialized DNA polymerases, cellular survival, and the genesis of mutations. Science 296: 1627-30.

Friedberg EC, Walker GC, Siede W, Wood RD, Schultz RA, Ellenberger T (2006) DNA repair and mutagenesis. 2nd edn. Washington, DC: ASM Press.

Friedenberg NA (2003) Experimental evolution of dispersal in spatiotemporally variable microcosms. Ecol Lett 6: 953–9.

Friedman CI, Danforth DR, Herbosa-Encarnacion C, Arbogast L, Alak BM, Seifer DB (1997) Follicular fluid vascular endothelial growth factor concentrations are elevated in women of advanced reproductive age undergoing ovulation induction. Fertil Steril 68: 607-12.

Friedman CI, Seifer DB, Kennard EA, Arbogast L, Alak B, Danforth DR (1998) Elevated level of follicular fluid vascular endothelial growth factor is a marker of diminished pregnancy potential. Fertil Steril 70: 836–9.

Frost LS, Leplae R, Summers AO, Toussaint A (2005) Mobile genetic elements: the agents of open source evolution. Nat Rev Microbiol 3: 72232.

Frost RJ, Hamra FK, Richardson JA, Qi X, Bassel-Duby R, Olson EN (2010) MOV10L1 is necessary forprotection of spermatocytes against retrotransposons by Piwi-interacting RNAs. Proc Natl Acad Sci USA 107: 11847-52.

Fruehauf JP, Meyskens FL Jr (2007) Reactive oxygen species: a breath of life or death? Clin Cancer Res 13: 789-94.

Frungieri MB, Zitta K, Pignataro OP, Gonzalez-Calvar SI, Calandra RS (2002) Interactions between testicular serotonergic, catecholaminergic and corticotropin-releasing factor systems modulating the cAMP and testosterone production in the Golden hamster. Neuroendocrinology 76: 35–46.

Frungieri MB, Mayerhofer A, Zitta K, Pignataro OP, Calandra RS, Gonzalez-Calvar SI (2005) Direct effect of melatonin on Syrian hamster testes: melatonin subtype 1a receptors, inhibition of androgen production, and interaction with the local corticotropin-releasing hormone system. Endocrinology 146: 1541–52.

Fry JD (1990) Tradeoffs in fitness on different hosts: evidence from a selection experiment with a phytophagous mite. Am Nat 136: 569–80.

Fry JD (1996) The evolution of host specialization: are trade-offs overrated? Am Nat 148(suppl.): S84–S107.

Fry JD (2003) Detecting ecological trade-offs using selection experiments. Ecology 84: 1672–8.

Fry JD (2004) On the rate and linearity of viability declines in Drosophila mutation-accumulation experiments: genomic mutation rates and synergistic epistasis revisited. Genetics 166: 797–806.

Fry JD, Keightley PD, Heinsohn SL, Nuzhdin SV (1999) New estimates of rates and effects of mildly deleterious mutation in Drosophila melanogaster. Proc Natl Acad Sci USA 96: 574-9.

Fryer HR, Frater J, Duda A, Roberts MG; SPARTAC Trial Investigators, et al. (2010) Modelling the evolution and spread of HIV immune escape mutants. PLoS Pathog 6: e1001196.

Fryxell KJ, Zuckerkandl E (2000) Cytosine deamination plays a primary role in the evolution of mammalian isochores. Mol Biol Evol 17: 1371–83.

Fu AQ, Genereux DP, Stöger R, Laird CD, Stephens M (2010) Statistical inference of transmission fidelity of DNA methylation patterns over somatic cell divisions in mammals. Ann Appl Stat 4: 871-92.

Fu S (2007) An illustrated introduction to the basic biological principles. arXiv:0712.2108v5 [q-bio.PE].

Fu YH, Kuhl DP, Pizzuti A, Pieretti M, Sutcliffe JS, et al. (1991) Variation of the CGG repeat at the fragile X site results in genetic instability: resolution of the Sherman paradox. Cell 67: 1047-58.

Fu Z, Kato H, Kotera N, Noguchi T, Sugahara K, Kubo T (2001) Regulation of hydroxyindole-O-methyltransferase gene expression in Japanese quail (Coturnix japonica). Biosci Biotechnol Biochem 65: 2504–11.

Fuchs RP, Fujii S, Wagner J (2004) Properties and functions of Escherichia coli: Pol IV and Pol V. Adv Protein Chem 69: 229–64.

Fudyk TC, Maclean IW, Simonsen JN, Njagi EN, Kimani J, et al. (1999) Genetic diversity and mosaicism at the por locus of Neisseria gonorrhoeae. J Bacteriol 181: 5591–9.

Fuentes-Mascorro G, Serrano H, Rosado A (2000) Sperm chromatin. Arch Androl 45: 215–25.

Fujihara N, Howarth B (1978) Lipid peroxidation in fowl spermatozoa. Poult Sci 57: 1766–8.

Fujii J, Iuchi Y, Matsuki S, Ishii T (2003) Cooperative function of antioxidant and redox systems against oxidative stress in male reproductive tissues. Asian J Androl 5: 231-42.

Fujii J, Iuchi Y, Okada F (2005) Fundamental roles of reactive oxygen species and protective mechanisms in the female reproductive system. Reprod Biol Endocrinol 3: 43.

Fujii J, Tsunoda S (2011) Redox regulation of fertilisation and the spermatogenic process. Asian J Androl 13: 420-3.

Fujimoto M, Nakai A (2010) The heat shock factor family and adaptation to proteotoxic stress. FEBS J 277: 4112–25.

Fujimoto Y, Matsui M, Fujita T (1982) The accumulation of ascorbic acid and iron in rat liver mitochondria after lipid peroxidation. Jpn J Pharmacol 32: 397-9.

Fujita M, Fujita Y, Noutoshi Y, Takahashi F, Narusaka Y, et al. (2006) Crosstalk between abiotic and biotic stress responses: a current view from the points of convergence in the stress signaling networks. Curr Opin Plant Biol 9: 436–42.

Fuks F (2005) DNA methylation and histone modifications: teaming up to silence genes. Curr Opin Genet Dev 15: 490-5.

Fukuda T, Hedinger C, Groscurth P (1975) Ultrastructure of developing germ cells in the fetal human testis. Cell Tissue Res 161: 55-70.

Fulda S, Gorman AM, Hori O, Samali A (2010) Cellular stress responses: cell survival and cell death. Int J Cell Biol 2010: 214074.

Funes JM, Quintero M, Henderson S, Martinez D, Qureshi U, et al. (2007) Transformation of human mesenchymal stem cells increases their dependency on oxidative phosphorylation for energy production. Proc Natl Acad Sci USA 104: 6223–8.

Fung H, Bennett RA, Demple B (2001) Key role of a downstream specificity protein 1 site in cell cycle-regulated transcription of the AP endonuclease gene APE1/APEX in NIH3T3 cells. J Biol Chem 276: 42011–7.

Fung H, Liu P, Demple B (2007) ATF4-dependent oxidative induction of the DNA repair enzyme Ape1 counteracts arsenite cytotoxicity and suppresses arsenite-mediated mutagenesis. Mol Cell Biol 27: 8834–47.

Furlong RF (2005) Insights into vertebrate evolution from the chicken genome sequence. Genome Biol 6: 207.

Furuchi T, Masuko K, Nishimune Y, Obinata M, Matsui Y (1996) Inhibition of testicular germ cell apoptosis and differentiation in mice misexpressing Bcl-2 in spermatogonia. Development 122: 1703–9.

Fussmann GF, Ellner SP, Hairston NG Jr (2003) Evolution as a critical component of plankton dynamics. Proc R Soc Lond B Biol Sci 270: 1015–22.

Fussmann GF, Loreau M, Abrams PA (2007) Eco-evolutionary dynamics of communities and ecosystems. Funct Ecol 21: 465–77.

Futuyma D (1998) Evolutionary Biology, 3rd edn. Sunderland, MA: Sinauer Associates.

Futuyma DJ (2010) Evolutionary constraint and ecological consequences. Evolution 64: 1865–84.

Futuyma DJ, Slatkin M, eds. (1983) Coevolution. Sunderland, MA: Sinauer Associates.

Futuyma DJ, Moreno G (1988) The evolution of ecological specialization. Annu Rev Ecol Syst 19: 207–33.

Gabor CR, Ryan MJ (2001) Geographical variation in reproductive character displacement in mate choice by male sailfin mollies. Proc R Soc Lond B 268: 1063-70.

Gabriel W (2005) How stress selects for reversible phenotypic plasticity. J Evol Biol 18: 873-83.

Gabriel W, Lynch M, Bürger R (1993) Muller's ratchet and mutational meltdowns. Evolution 47: 1744-57.

Gackowski D, Kruszewski M, Bartlomiejczyk T, Jawien A, Ciecierski M, Olinski R (2002) The level of 8-oxo-7,8-dihydro-2’-deoxyguanosine is positively correlated with the size of the labile iron pool in human lymphocytes. J Biol Inorg Chem 7: 548–50.

Gade B, Parker ED Jr (1997) The effect of life cycle stage and genotype on desiccation tolerance in the colonizing parthenogenetic cockroach Pycnoscelus surinamensis and its sexual ancestor P. indicus. J Evol Biol 10: 479–93.

Gadgil M, Solbrig OT (1972) The concept of r- and K-selection: Evidence from wild flowers and some theoretical considerations. Am Nat 106: 14–31.

Gaedicke S, Zhang X, Schmelzer C, Lou Y, Doering F, et al. (2008) Vitamin E dependent microRNA regulation in rat liver. FEBS Lett 582: 3542-6.

Gaeta RT, Pires CJ (2010) Homoeologous recombination in allopolyploids: the polyploid ratchet. New Phytol 186: 18-28.

Gaeth AP, Short RV, Renfree MB (1999) The developing renal, reproductive, and respiratory systems of the African elephant suggest an aquatic ancestry. Proc Natl Acad Sci USA 96: 5555-8.

Gaffé J, McKenzie C, Maharjan RP, Coursange E, Ferenci T, Schneider D (2011) Insertion sequence-driven evolution of Escherichia coli in chemostats. J Mol Evol 72: 398–412.

Gaffney P, Bushak D (1996) Genetic aspects of disease resistance in oysters. J Shellfish Res 15: 135–40.

Gafter U, Malachi T, Ori Y, Breitbart H (1997) The role of calcium in human lymphocyte DNA repair ability. J Lab Clin Med 130: 33-41.

Gage MJ (1991) Risk of sperm competition directly affects ejaculate size in the Mediterranean fruit fly. Anim Behav 42: 1036–7.

Gage MJG (1994) Associations between body size, mating pattern, testis size and sperm lengths across butterflies. Proc R Soc Lond B 258: 247-54.

Gage MJ, Parker GA, Nylin S, Wiklund C (2002) Sexual selection and speciation in mammals, butterflies and spiders. Proc R Soc Lond B Biol Sci 269: 2309-16.

Gage MJG, Macfarlane CP, Yeates S, Ward RG, Searle JB, et al. (2004) Spermatozoal traits and sperm competition in Atlantic Salmon: relative sperm velocity is the primary determinant of fertilization success. Curr Biol 14: 44–7.

Gagliano M, McCormick MI (2009) Hormonally mediated effects shape offspring survival potential in stressful environments. Oecologia 160: 657–65.

Gaiddon C, Moorthy NC, Prives C (1999) Ref-1 regulates the transactivation and pro-apoptotic functions of p53 in vivo. EMBO J 18: 5609–21.

Galeotti P, Rubolini D, Fea G, Ghia D, Nardi PA, Gherardi F, Fasola M (2006) Female freshwater crayfish adjust egg and clutch size in relation to multiple male traits. Proc R Soc B 273: 1105–10.

Galewski T, Tilak M, Sanchez S, Chevret P, Paradis E, Douzery E (2006) The evolutionary radiation of arvicolinae rodents (voles and lemmings): relative contribution of nuclear and mitochondrial DNA phylogenies. BMC Evol Biol 6: 80.

Galhardo RS, Hastings PJ, Rosenberg SM (2007) Mutation as a stress response and the regulation of evolvability. Crit Rev Biochem Mol Biol 42: 399-435.

Galhardo RS, Do R, Yamada M, Friedberg EC, Hastings PJ, Nohmi T, Rosenberg SM (2009) DinB upregulation is the sole role of the SOS response in stress-induced mutagenesis in Escherichia coli. Genetics 182: 55-68.

Gallant P (2006) Myc/Max/Mad in invertebrates: the evolution of the Max network. Curr Top Microbiol Immunol 302: 235-53.

Gallo CM, Munro E, Rasoloson D, Merritt C, Seydoux G (2008) Processing bodies and germ granules are distinct RNA granules that interact in C. elegans embryos. Dev Biol 323: 76–87.

Gallou-Kabani C, Junien C (2005) Nutritional epigenomics of metabolic syndrome: new perspective against the epidemic. Diabetes 54: 1899–906.

Galloway LF, Etterson JR (2007) Transgenerational plasticity is adaptive in the wild. Science 318: 1134–6.

Galluzzi L, Morselli E, Kepp O, Vitale I, Pinti M, Kroemer G (2011) Mitochondrial liaisons of p53. Antioxid Redox Signal 15: 1691-714.

Galtier N, Piganeau G, Mouchiroud D, Duret L (2001) GC-content evolution in mammalian genomes: the biased gene conversion hypothesis. Genetics 159: 907–11.

Gamaley IA, Klyubin IV (1999) Roles of reactive oxygen species: signaling and regulation of cellular functions. Int Rev Cytol 188: 203–55.

Gambino V, De Michele G, Venezia O, Migliaccio P, Dall'olio V, et al. (2013) Oxidative stress activates a specific p53 transcriptional response that regulates cellular senescence and aging. Aging Cell 12: 435-45.

Gamfeldt L, Wallen J, Jonsson PR, Berntsson KM, Havenhand JN (2005) Increasing intraspecific diversity enhances settling success in a marine invertebrate. Ecology 86: 3219–24.

Gamfeldt L, Kallstrom B (2007) Increasing intraspecific diversity increases predictability in population survival in the face of perturbations. Oikos 116: 700–5.

Gandini L, Lombardo F, Paoli D, Caponecchia L, Familiari G, et al. (2000) Study of apoptotic DNA fragmentation in human spermatozoa. Hum Reprod 15: 830–9.

Gandolfi A, Bonilauri P, Rossi V, Menozzi P (2001) Intraindividual and intraspecies variability of ITS1 sequences in the ancient asexual Darwinula stevensoni (Crustacea: Ostracoda). Heredity 87: 449-55.

Gandolfi A, Sanders IR, Rossi V, Menozzi P (2003) Evidence of recombination in putative ancient asexuals. Mol Biol Evol 20: 754–61.

Gangaraju VK, Yin H, Weiner MM, Wang J, Huang XA, Lin H (2011) Drosophila Piwi functions in Hsp90-mediated suppression of phenotypic variation. Nat Genet 43: 153–8.

Gantenbein B, Fet V, Gantenbein-Ritter IA, Balloux F (2005) Evidence for recombination in scorpion mitochondrial DNA (Scorpiones: Buthidae). Proc R Soc B: Biol Sci 272: 697–704.

Gao H-B, Ge R-S, Lakshmi V, Marandici A, Hardy MP (1997) Hormonal regulation of oxidative and reductive activities of 11β-hydroxysteroid dehydrogenase in rat Leydig cells. Endocrinology 138: 156–61.

Gao M, Arkov AL (2012) Next generation organelles: Structure and role of germ granules in the germline. Mol Reprod Dev doi: 10.1002/mrd.22115. [Epub ahead of print]

Gao S, Liu Y (2012) Intercepting noncoding messages between germline and soma. Genes Dev 26: 1774-9.

Gao X, Ge L, Shao J, Su C, Zhao H, et al. (2010) Tudor-SN interacts with and co-localizes with G3BP in stress granules under stress conditions. FEBS Lett 584: 3525-32.

Gapper C, Dolan L (2006) Control of plant development by reactive oxygen species. Plant Physiol 141: 341–5.

Garamszegi LZ, Eens M, Hurtrez-Boussès S, Møller AP (2005) Testosterone, testes size, and mating success in birds: a comparative study. Horm Behav 47: 389-409.

García-Díaz M, Domínguez O, López-Fernández LA, de Lera LT, Saníger ML, et al. (2005) Sperm viability matters in insect sperm competition. Curr Biol 15: 271–5.

García-González F, Simmons LW (2011) Good genes and sexual selection in dung beetles (Onthophagus taurus): genetic variance in egg-to-adult and adult viability. PLoS ONE 6: e16233.

García-Martínez J, Castro JA, Ramón M, Latorre A, Moya A (1998) Mitochondrial DNA haplotype frequencies in natural and experimental populations of Drosophila subobscura. Genetics 149: 1377-82.

Garcia-Ramos G, Kirkpatrick M (1997) Genetic models of adaptation and gene flow in peripheral populations. Evolution 51: 21–8.

Gardner A, Grafen A (2009) Capturing the superorganism: a formal theory of group adaptation. J Evol Biol 22: 659–71.

Gardner L, Anderson T, Place AR, Elizur A (2003) Sex change strategy and the aromatase genes. Fish Physiol Biochem 28: 147–8.

Gardner L, Anderson T, Place AR, Dixon B, Elizur A (2005) Sex change strategy and the aromatase genes. J Steroid Biochem Mol Biol 94: 395-404.

Garfinkel MR, Skapaderas S (2007) Economics of conflict: an overview. In: Sandler T, Hartley K, eds. Handbook of defense economics. Amsterdam, The Netherlands: Springer. pp 649–709.

Garten C Jr (1976) Relationships between aggressive behavior and genetic heterozygosity in the oldfield mouse, Peromyscus polionotus. Evolution 41: 80–91.

Gartner A, Milstein S, AhmedS, Hodgkin J, Hengartner MO (2000) A conserved checkpoint pathway mediates DNA damage–induced apoptosis and cell cycle arrest in C. elegans. Mol Cell 5: 435-443.

Gartner A, Boag PR, Blackwell TK (2008) Germline survival and apoptosis. WormBook 4: 1–20.

Garvin JC, Abroe B, Pedersen MC, Dunn PO, Whittingham LA (2006) Immune response of nestling warblers varies with extra-pair paternity and temperature. Mol Ecol 15: 3833–40.

Garzon R, Liu S, Fabbri M, Liu Z, Heaphy CE, et al. (2009) MicroRNA-29b induces global DNA hypomethylation and tumor suppressor gene reexpression in acute myeloid leukemia by targeting directly DNMT3A and 3B and indirectly DNMT1. Blood 113: 6411-8.

Gasior SL, Wakeman TP, Xu B, Deininger PL (2006) The human LINE-1 retrotransposon creates DNA double-strand breaks. J Mol Biol 357: 1383-93.

Gasparini C, Simmons LW, Beveridge M, Evans JP (2010) Sperm swimming velocity predicts competitive fertilization success in the Green Swordtail Xiphophorus helleri. PLoS ONE 5: e12146.

Gaston KJ (1996) Biodiversity — latitudinal gradients. Prog Phys Geogr 20: 466–76.

Gaston KJ (2000) Global patterns in biodiversity. Nature 405: 220–7.

Gaston KJ (2003) The structure and dynamics of geographic ranges. Oxford, UK: Oxford University Press.

Gaston KJ (2009) Geographic range limits: achieving synthesis. Proc R Soc B 276: 1395–406.

Gatewood JM, Cook GR, Balhorn R, Bradbury EM, Schmid CW (1987) Sequence-specific packaging of DNA in human-sperm chromatin. Science 236: 962–4.

Gatti JL, Castella S, Dacheux F, Ecroyd H, Metayer S, et al. (2004) Post-testicular sperm environment and fertility. Anim Reprod Sci 82-83: 321-39.

Gaudet F, Hodgson JG, Eden A, Jackson-Grusby L, Dausman J, et al. (2003) Induction of tumors in mice by genomic hypomethylation. Science 300: 489-92.

Gaudet F, Rideout WM 3rd, Meissner A, Dausman J, Leonhardt H, Jaenisch R (2004) Dnmt1 expression in pre- and postimplantation embryogenesis and the maintenance of IAP silencing. Mol Cell Biol 24: 1640-8.

Gaudin E, Rosado M, Agenes F, McLean A, Freitas AA (2004) B-cell homeostasis, competition, resources, and positive selection by self-antigens. Immunol Rev 197: 102-15.

Gause GF (1934) The Struggle for Existence. Baltimore, MA: Williams and Wilkins Co.

Gauss KA, Bunger PL, Larson TC, Young CJ, Nelson-Overton LK, et al. (2005) Identification of a novel tumor necrosis factor alpha-responsive region in the NCF2 promoter. J Leukoc Biol 77: 267-78.

Gavazza M, Catala A (2003) Melatonin preserves arachidonic and docosapetaenoic acids during ascorbate-Fe++ peroxidation of rat testis microsomes and mitochondria. Int J Biochem Cell Biol 35: 359–66.

Gavella M, Lipovac V (1992) NADH-dependent oxidoreductase (diaphorase) activity and isozyme pattern of sperm in infertile men. Arch Androl 28: 135–41.

Gavrilets S (2000) Rapid evolution of reproductive barriers driven by sexual conflict. Nature 403: 886-9.

Gavrilets S (2010) Rapid transition towards the division of labor via evolution of developmental plasticity. PLoS Comput Biol 6: 621–54.

Gavrilets S, Arnqvist G, Friberg U (2001) The evolution of female mate choice by sexual conflict. Proc R Soc LondSer B Biol Sci 268: 531–9.

Gaymes TJ, Mufti GJ, Rassool FV (2002) Myeloid leukemias have increased activity of the nonhomologous end-joining pathway and concomitant DNA misrepair that is dependent on the Ku70/86 heterodimer. Cancer Res 62: 2791–7.

Geber MA, Dawson TE, Delph LF, eds. (1999) Gender and sexual dimorphism in flowering plants. Berlin, Germany: Springer.

Geber MA, Eckhart VM (2005) Experimental studies of adaptation in Clarkia xantiana. II. Fitness variation across a subspecies border. Evolution 59: 521–31.

Gebhardt F, Zanker KS, Brandt B (1999) Modulation of epidermal growth factor receptor gene transcription by a polymorphic dinucleotide repeat in intron 1. J Biol Chem 274: 13176-80.

Gebhardt F, Burger H, Brandt B (2000) Modulation of EGFR gene transcription by a polymorphic repetitive sequence – a link between genetics and epigenetics. Int J Biol Markers 15: 105-10.

Gechev TS, Van Breusegem F, Stone JM, Denev I, Laloi C (2006) Reactive oxygen species as signals that modulate plant stress responses and programmed cell death. BioEssays 28: 1091–101.

Geedey CK, Tessier AJ, Machledt K (1996) Habitat heterogeneity, environmental change, and the clonal structure of Daphnia populations. Funct Ecol 10: 613–21.

Gehring M, Bubb KL, Henikoff S (2009) Extensive demethylation of repetitive elements during seed development underlies gene imprinting. Science 324: 1447–51.

Geiger W (1998) Population dynamics, life histories and reproductive modes In: Martens K,ed. Sex and Parthenogenesis: Evolutionary Ecology of Reproductive Modes in Non-marine Ostracods, Leiden, The Netherlands: Backhuys Publishers. pp 215–228.

Geiman TM, Muegge K (2010) DNA methylation in early development. Mol Reprod Dev 77: 105-13.

Geissler EN, McFarland EC, Russell ES (1981) Analysis of pleiotropism at the dominant white-spotting (W) locus of the house mouse: a description of ten new W alleles. Genetics 97: 337-61.

Geissler EN, Ryan MA, Housman DE (1988) The dominant-white spotting (W) locus of the mouse encodes the c-kit proto-oncogene. Cell 55: 185–92.

Gemayel R, Vinces MD, Legendre M, Verstrepen KJ (2010) Variable tandem repeats accelerate evolution of coding and regulatory sequences. Annu Rev Genet 44: 445-77.

Gemmell NJ, Western PS, Watson JM, Graves JAM (1996) Evolution of the mammalian mitochondrial control region - comparisons of control region sequences between monotreme and therian mammals. Mol Biol Evol 13: 798–808.

Gemmell NJ, Metcalf VJ, Allendorf FW (2004) Mother’s curse: the effect of mtDNA on individual fitness and population viability. Trends Ecol Evol 19: 238–44.

Gemmill AW, Viney ME, Read AF (1997) Host immune status determines sexuality in a parasitic nematode. Evolution 51: 393-401.

Gena P, Fanelli E, Brenner C, Svelto M, Calamita G (2009) News and views on mitochondrial water transport. Front Biosci 14: 4189–98.

Genereux DP, Miner BE, Bergstrom CT, Laird CD (2005) A population-epigenetic model to infer site-specific methylation rates from double-stranded DNA methylation patterns. Proc Natl Acad Sci USA 102: 5802–7.

Generoso WM, Cain KT, Krishna M, Huff SW (1979) Genetic lesions induced by chemicals in spermatozoa and spermatids of mice are repaired in the egg. Proc Natl Acad Sci USA 76: 435-7.

Geng YP, Pan XY, Xu CY, Zhang WJ, Li B, et al. (2007) Phenotypic plasticity rather than locally adapted ecotypes allows the invasive alligator weed to colonize a wide range of habitats. Biol Invasions 9: 245–56.

Genova ML, Bianchi C, Lenaz G (2003) Structural organization of the mitochondrial respiratory chain. Ital J Biochem 52: 58-61.

Gérard N, Syed V, Bardin W, Genetet N, Jégou B (1991) Sertoli cells are the site of interleukin-1α synthesis in rat testis. Mol Cell Endocrinol 82: R13–R16.

Gerber HP, Seipel K, Georgiev O, Hofferer M, Hug M, et al. (1994) Transcriptional activation modulated by homopolymeric glutamine and proline stretches. Science 263: 808–11.

Gerhart J (2000) Inversion of the chordate body axis : are there alternatives? Proc Natl Acad Sci USA 97: 4445–8.

Gerhart J, Kirschner M (2007) The theory of facilitated variation. Proc Natl Acad Sci USA 104 Suppl 1: 8582-9.

Gerlach VL, Feaver WJ, Fischhaber PL, Richardson JA, Aravind L, et al. (2000) Human DNA polymerase kappa: a novel DNA polymerase of unknown biological function encoded by the DINB1 gene. Cold Spring Harb Symp Quant Biol 65: 41-9.

Gerone PJ, Couch RB, Ketter GV, Douglas RG, Derrenbacher EB, Knight V (1966) Assessment of experimental and natural viral aerosols. Bacteriol Rev 30: 576–88.

Gerrish PJ, Lenski RE (1998) The fate of competing beneficial mutations in an asexual population. Genetica 102-103:127-44.

Gerrish PJ, Colato A, Perelson AS, Sniegowski PD (2007) Complete genetic linkage can subvert natural selection. Proc Natl Acad Sci USA 104: 6266–71.

Gerrish PJ, Colato A, Sniegowski PD (2012) Genomic mutation rates that neutralize adaptive evolution and natural selection. arXiv: 2012.0974. [q-bio.PE].

Gerritsen J (1980) Sex and parthenogenesis in sparse populations. Am Nat 115: 718–42.

Gersani M, Brown JS, O’Brien EE, Maina GM, Abramsky Z (2001) Tragedy of the commons as a result of root competition. J Ecol 89: 660-9.

Gerstein AC (2012) Mutational effects depend on ploidy level: all else is not equal. Biol Lett 9: 20120614.

Gerstein AC, Otto SP (2009) Ploidy and the causes of genomic evolution. J Hered 100: 571-81.

Gerstein AC, Otto SP (2011) Cryptic fitness advantage: diploids invade haploid populations despite lacking any apparent advantage as measured by standard fitness assays. PLoS ONE 6: e26599.

Gessler DDG, Xu SZ (2000) Meiosis and the evolution of recombination at low mutation rates. Genetics 156: 449–56.

Gethmann RC (1988) Crossing over in males of higher Diptera (Brachycera). J Hered 79: 344-50.

Geva E, Jaffe RB (2000a) Role of vascular endothelial growth factor in ovarian physiology and pathology. Fertil Steril 74: 429–38.

Geva E, Jaffe RB (2000b) Role of angiopoietins in reproductive tract angiogenesis. Obstet Gynecol Surv 55: 511–9.

Ghabrial AS, Krasnow MA (2006) Social interactions among epithelial cells during tracheal branching morphogenesis. Nature 441: 746–9.

Ghafari F, Pelengaris S, Walters E, Hartshorne GM (2009) Influence of p53 and genetic background on prenatal oogenesis and oocyte attrition in mice. Hum Reprod 24: 1460–72.

Ghalambor CK, McKay JK, Carroll SP, Reznick DN (2007) Adaptive versus non-adaptive phenotypic plasticity and the potential for contemporary adaptation in new environments. Funct Ecol 21: 394–407.

Giaccia AJ, Simon MC, Johnson R (2004) The biology of hypoxia: the role of oxygen sensing in development, normal function, and disease. Genes Dev 18: 2183–94.

Giannakakis A, Sandaltzopoulos R, Greshock J, Liang S, Huang J, et al. (2008) miR-210 links hypoxia with cell cycle regulation and is deleted in human epithelial ovarian cancer. Cancer Biol Ther 7: 255–64.

Gibbons JRH (1979) A model for sympatric speciation in Megarhyssa (Hymenoptera: Ichneumonidae): competitive speciation. Am Nat 114: 719-41.

Gibbs AJ, Fargette D, García-Arenal F, Gibbs MJ (2009) Time – the emerging dimension of plant virus studies. J Gen Virol 91: 13–22.

Gibbs HL, Grant PR (1987) Oscillating selection on Darwin’s finches. Nature 327: 511–3.

Gibbs RB (1997) Effects of estrogen on basal forebrain cholinergic neurons vary as a function of dose and duration of treatment. Brain Res 757: 10-6.

Gibson G, Wagner G (2000) Canalization in evolutionary genetics: a stabilizing theory? BioEssays 22: 372–80.

Gibson G, Dworkin I (2004) Uncovering cryptic genetic variation. Nat Rev Genet 5: 681–90.

Gibson JL, Lombardo MJ, Thornton PC, Hu KH, Galhardo RS, et al. (2010) The sigma(E) stress response is required for stress-induced mutation and amplification in Escherichia coli. Mol Microbiol 77: 415-30.

Gibson TC, Scheppe ML, Cox EC (1970) Fitness of an Escherichia coli mutator gene. Science 169: 686-8.

Giehl KM (2001) Trophic dependencies of rodent corticospinal neurons. Rev Neurosci 12: 79–94.

Giesel JT (1976) Reproductive strategies as adaptations to life in temporally heterogeneous environments. Annu Rev Ecol Syst 7: 57-79.

Gil D, Graves J, Hazon N, Wells A (1999) Male attractiveness and differential testosterone investment in zebra finch eggs. Science 286: 126–8.

Gil D, Heim C, Bulmer E, Rocha M, Puerta M, Naguib M (2004) Negative effects of early developmental stress on yolk testosterone levels in a passerine bird. J Exp Biol 207: 2215–20.

Gil-Guzman E, Ollero M, Lopez MC, Sharma RK, Alvarez JG, et al. (2001) Differential production of reactive oxygen species by subsets of human spermatozoa at different stages of maturation. Hum Reprod 16: 1922-30.

Gilbert C, Schaack S, Pace JK, Brindley PJ, Feschotte C (2010) A role for host-parasite interactions in the horizontal transfer of transposons across phyla. Nature 464: 1347–50.

Gilbert JJ (1998) Asexual diapause in the rotifer Synchaeta: diversified bet-hedging, energetic cost and age effects. Ergeb Limnol 52: 97–107.

Gilbert JJ (2003) Environmental and endogenous control of sexuality in a rotifer life cycle: Development and population biology. Evol Dev 5: 19–24.

Gilbert JJ (2004) Population density, sexual reproduction and diapause in monogonont rotifers: new data for Brachionus and a review. J Limnol 63(Suppl. 1): 32-6.

Gilbert JJ, Schreiber DK (1995) Induction of diapausing amictic eggs in Synchaeta pectinata. Hydrobiologia 313/314: 345–50.

Gilbert JJ, Schreiber DK (1998) Asexual diapause induced by food limitation in the rotifer Synchaeta pectinata. Ecology 79: 1371–81.

Gilbert N, Lutz-Prigge S, Moran JV (2002) Genomic deletions created upon LINE-1 retrotransposition. Cell 110: 315-25.

Gilbert SF, Bolker JA (2001) Homologies of process and modular elements of embryonic construction. J Exp Zool 291: 1–12.

Gilchrist RB, Lane M, Thompson JG (2008) Oocyte-secreted factors: regulators of cumulus cell function and oocyte quality. Hum Reprod Update 14: 159–77.

Gildenhuys P (2004) Darwin, Herschel, and the role of analogy in Darwin’s origin. Stud Hist Phil Biol & Biomed Sci 35: 593–611.

Giles NM, Giles GI, Jacob C (2003) Multiple roles of cysteine in biocatalysis. Biochem Biophys Res Commun 300: 1–4.

Giles RE, Blanc H, Cann HM, Wallace DC (1980) Maternal inheritance of human mitochondrial DNA. Proc Natl Acad Sci USA 77: 6715–9.

Gilks N, Kedersha N, Ayodele M, Shen L, Stoecklin G, et al. (2004) Stress granule assembly is mediated by prion-like aggregation of TIA-1. Mol Biol Cell 15: 5383-98.

Gill DE, Chao L, PerkinsSL, Wolf JB (1995) Genetic mosaicism in plants and clonal animals. Annu Rev Ecol Syst26: 423–44.

Gillespie DOS, Russell AF, Lummaa V (2008) When fecundity does not equal fitness: evidence of an offspring quantity vs. quality trade-off in pre-industrial humans. Proc R Soc B 275: 713–22.

Gillespie JH (1973) Natural selection with varying selection coefficients - a haploid model. Genet Res 21: 115-20.

Gillespie JH (1974a) Natural selection for within-generation variance in offspring numbers. Genetics 76: 601–6.

Gillespie JH (1974b) Polymorphism in patchy environments. Am Nat 108: 145-51.

Gillespie JH (1977) Natural selection for variances in offspring numbers: a new evolutionary principle. Am Nat 111: 1010-4.

Gillespie JH (1981) Evolution of the mutation rate at a heterotic locus. Proc Natl Acad Sci USA 78: 2452–4.

Gillespie JH (1991a) Mutation modification in a random environment. Evolution 35: 468–76.

Gillespie JH (1991b) The causes of molecular evolution. New York, NY: Oxford University Press.

Gillespie JH (1999) The role of population size in molecular evolution. Theor Popul Biol. 55: 145–56.

Gillespie JH (2000) Genetic drift in an infinite population. The pseudohitchhiking model. Genetics 155: 909–19.

Gillespie JH (2001) Is the population size of a species relevant to its evolution? Evolution 55: 2161–9.

Gilligan DM, Woodworth LM, Montgomery ME, Briscoe DA, Frankham R (1997) Is mutation accumulation a threat to the survival of endangered populations? Conserv Biol 11: 1235-41.

Gillman LN, Keeling DJ, Ross HA, Wright SD (2009) Latitude, elevation and the tempo of molecular evolution in mammals. Proc Biol Sci 276: 3353-9.

Gillman LN, Keeling DJ, Gardner RC, Wright SD (2010) Faster evolution of highly conserved DNA in tropical plants. J Evol Biol 23: 1327-30.

Gillooly JF, Brown JH, West GB, Savage VM, Charnov EL (2001) Effects of size and temperature on metabolic rate. Science 293: 2248–51.

Gillooly JF, Allen AP, West GB, Brown JH (2005) The rate of DNA evolution: effects of body size and temperature on the molecular clock. Proc Natl Acad Sci USA 102: 140–5.

Gillooly JF, McCoy MW, Allen AP (2007) Effects of metabolic rate on protein evolution. Biol Lett 3: 655–9.

Gilpin M (1975) Group Selection in Predator-Prey Communities. Princeton, NJ: Princeton University Press.

Gingerich PD (1983) Rates of evolution: effects of time and temporal scaling. Science 222: 159–61.

Gingerich PD (2001) Rates of evolution on the time scale of the evolutionary process. Genetica 112–113: 127–44.

Gingerich PD (2009) Rates of evolution. Annu Rev Ecol Evol Syst 40: 657–75.

Girard A, Sachidanandam R, Hannon GJ, Carmell MA (2006) A germline-specific class of small RNAs binds mammalian Piwi proteins. Nature 442: 199–202.

Girard M, Couvert P, Carrié A, Tardieu M, Chelly J, Beldjord C, Bienvenu T (2001) Parental origin of de novo MECP2 mutations in Rett syndrome. Eur J Hum Genet 9: 231–6.

Giraud A, Matic I, Tenaillon O, Clara A, Radman M, et al. (2001a) Costs and benefits of high mutation rates: adaptive evolution of bacteria in the mouse gut. Science 291: 2606–8.

Giraud A, Radman M, Matic I, Taddei F (2001b) The rise and fall of mutator bacteria. Curr Opin Microbiol 4: 582-5.

Giraud A, Matic I, Radman M, Fons M, Taddei F (2002) Mutator bacteria as a risk factor in treatment of infectious diseases. Antimicrob Agents Chemother 46: 863–5.

Girod JP, Brotman DJ (2004) Does altered glucocorticoid homeostasis increase cardiovascular risk? Cardiovasc Res 64: 217–26.

Gissis SB, Jablonka E, eds. (2011) Transformations of Lamarckism: from subtle fluids to molecular biology. Cambridge, MA: MIT Press.

Gittleman JL, Thompson SD (1988) Energy allocation in mammalian reproduction. Am Zool 28: 863–75.

Gjerstorff MF, Johansen LE, Nielsen O, Kock K, Ditzel HJ (2006) Restriction of GAGE protein expression to subpopulations of cancer cells is independent of genotype and may limit the use of GAGE proteins as targets for cancer immunotherapy. Br J Cancer 94: 1864-73.

Glaab WE, Risinger JI, Umar A, Barrett JC, Kunkel TA, et al. (1998) Characterization of distinct human endometrial carcinoma cell lines deficient in mismatch repair that originated from a single tumor. J Biol Chem 41: 26662-9.

Gladden LB (2004) Lactate metabolism: a new paradigm for the third millennium. J Physiol 558: 5-30.

Gladyshev EA, Meselson M, Arkhipova IR (2008) Massive horizontal gene transfer in bdelloid rotifers. Science 320: 1210–3.

Gladyshev EA, Meselson M (2008) Extreme resistance of bdelloid rotifers to ionizing radiation. Proc Natl Acad Sci USA 105: 5139–44.

Gladyshev EA, Arkhipova IR (2010) Genome structure of bdelloid rotifers: shaped by asexuality or desiccation? J Hered 101 Suppl 1: S85–93.

Glaser RL, Lis JT (1990) Multiple, compensatory regulatory elements specify spermatocyte-specific expression of the Drosophila melanogaster hsp26 gene. Mol Cell Biol 10: 131-7.

Glaser RL, Jiang W, Boyadjiev SA, Tran AK, Zachary AA, et al. (2000) Paternal origin of FGFR2 mutations in sporadic cases of Crouzon syndrome and Pfeiffer syndrome. Am J Hum Genet 66: 768-77.

Glaser RL, Jabs EW (2004) Dear old dad. Sci Aging Knowledge Environ 2004: re1.

Glassner BJ, Rasmussen LJ, Najarian MT, Posnick LM, Samson LD (1998) Generation of a strong mutator phenotype in yeast by imbalanced base excision repair. Proc Natl Acad Sci USA 95: 9997–10002.

Glesener RR (1979) Recombination in a simulated predator-prey interaction. Am Zool 19: 763–71.

Glesener RR, Tilman D (1978) Sexuality and the components of environmental uncertainty: clues from geographic parthenogenesis in terrestrial animals. Am Nat 112: 659–73.

Glickman SE, Short RV, Renfree MB (2005) Sexual differentiation in three unconventional mammals: spotted hyenas, elephants and tammar wallabies. Horm Behav 48: 403-17.

Gloire G, Legrand-Poels S, Piette J (2006) NF-κB activation by reactive oxygen species: fifteen years later. Biochem Pharmacol 72: 1493–505.

Gluckman P, Hanson M (2006) Mismatch: why our world no longer fits our bodies. Oxford, UK: Oxford University Press.

Gluckman PD, Hanson MA, Beedle AS (2007) Early life events and their consequences for later disease: a life history and evolutionary perspective. Am J Hum Biol 19: 1–19.

Go YM, Jones DP (2010) Redox control systems in the nucleus: mechanisms and functions. Antioxid Redox Signal 13: 489-509.

Goday C, Esteban MR (2001) Chromosome elimination in sciarid flies. Bioessays 23: 242–50.

Goddard KA, Dawley RM (1990) Clonal inheritance of a diploid nuclear genome by a hybrid freshwater minnow (Phoxinus eos-neogaeus, Pisces: Cyprinidae). Evolution 44: 1052–65.

Goddard KA, Megwinoff O, Wessner LL, Giaimo F (1998) Confirmation of gynogenesis in Phoxinus eos-neogaeus (Pisces:Cyprinidae). J Hered 89: 151–7.

Goddard MR, Godfray HC, Burt A (2005) Sex increases the efficacy of natural selection in experimental yeast populations. Nature 434: 636–40.

Godeau F, Persson H, Gray HE, Pardee AB (1986) C-myc expression is dissociated from DNA synthesis and cell division in Xenopus oocyte and early embryonic development. EMBO J 5: 3571-7.

Godfray HCJ (1995) Evolutionary theory of parent-offspring conflict. Nature 376: 133–8.

Godfrey SS, Sussman M (1982) The genetics of development in Dictyostelium discoideum. Annu Rev Genet 16: 385-404.

Godfrey-Smith P (2001) Three kinds of adaptationism. In: Orzack S, Sober E (eds) Adaptationism and optimality. Cambridge, UK: Cambridge University Press. pp 335–357.

Godfrey-Smith P (2009) Darwinian populations and natural selection. Oxford, UK: Oxford University Press.

Godoy VG, Gizatullin FS, Fox MS (2000) Some features of the mutability of bacteria during nonlethal selection. Genetics 154: 49–59.

Goedecke W, Eijpe M, Offenberg HH, van Aalderen M, Heyting C (1999) Mre11 and Ku70 interact in somatic cells, but are differentially expressed in early meiosis. Nat Genet 23: 194-8.

Goerlich O, Quillardet P, Hofnung M (1989) Induction of the SOS response by hydrogen peroxide in various Escherichia coli mutants with altered protection against oxidative DNA damage. J Bacteriol 171: 6141-7.

Goetz R, Fuglsang A (2005) Correlation of codon bias measures with mRNA levels: analysis of transcriptome data from Escherichia coli. Biochem Biophys Res Commun 327: 4–7.

Gogolevsky KP, Vassetzky NS, Kramerov DA (2008) Bov-B-mobilised SINEs in vertebrate genomes. Gene 407:75–85.

Gogvadze E, Buzdin A (2009) Retroelements and their impact on genome evolution and functioning. Cellular and Molecular Life Sci 66: 3727–3742.

Gohli J, Anmarkrud JA, Johnsen A, Kleven O, Borge T, Lifjeld JT (2013) Female promiscuity is positively associated with neutral and selected genetic diversity in passerine birds. Evolution 67: 1406-19.

Goho S, Bell G (2000) Mild environmental stress elicits mutations affecting fitness in Chlamydomonas. Proc R Soc Lond Ser B Biol Sci 267: 123–9.

Gojobori T, Li WH, Graur D (1982) Patterns of nucleotide substitutions in pseudogenes and functional genes. J Mol Evol 18: 360-9.

Gokhale CS, Iwasa Y, Nowak MA, Traulsen A (2009) The pace of evolution across fitness valleys. J Theor Biol 259: 613–20.

Gold B, Fujimoto H, Kramer JM, Erickson RP, Hecht NB (1983) Haploid accumulation and translational control of phosphoglycerate kinase-2 messenger RNA during mouse spermatogenesis. Dev Biol 98: 392–9.

Goldbach R (1987) Genome similarities between plant and animal RNA viruses. Microbiol Sci 4: 197–202.

Goldberg AD, Allis CD, Bernstein E (2007) Epigenetics: a landscape takes shape. Cell 128: 635–8.

Goldstein M, Eid JF (1989) Elevation of intratesticular and scrotal skin surface-temperature in man with varicocele. J Urol 142: 173–5.

Golstein P, Kroemer G (2007) Cell death by necrosis: towards a molecular definition. Trends Biochem Sci 32: 37-43.

Goldstein S, Czapski G (1986) The role and mechanism of metal ions and their complexes in enhancing damage in biological systems or in protecting these systems from the toxicity of O2¯. J Free Radic Biol Med 2: 3-11.

Goll MG, Bestor TH (2005) Eukaryotic cytosine methyltransferases. Annu Rev Biochem 74: 481–514.

Golovatch SI, Kime RD (2009) Millipede (Diplopoda) distributions: A review. Soil Organisms 81: 565–97.

Gomendio M, Roldan ERS (1993) Coevolution between male ejaculates and female reproductive biology in eutherian mammals. Proc R Soc Lond B 252: 7–12.

Gomer RH (1994) Intercellular signalling. Knowing that you’re among friends. Curr Biol 4: 734-5.

Gomes WR (1970) Chemical agents affecting testicular function and male fertility. In: Johnson AD, Gomes WR, VanDemark NL, eds. The Testis, Vol. III. New York, NY: Academic Press. pp 483-554.

Gómez A, Carvalho GR (2000) Sex, parthenogenesis and genetic structure of rotifers. Microsatellite analysis of contemporary and resting egg bank populations. Mol Ecol 9: 203–14.

Gómez P, Buckling A (2011) Bacteria-phage antagonistic coevolution in soil. Science 332: 106-9.

Gomez-Cabrera MC, Viña J, Ji LL (2009) Interplay of oxidants and antioxidants during exercise: implications for muscle health. Phys Sportsmed 37: 116–23.

Gomulkiewicz R, Kirkpatrick M (1992) Quantitative genetics and the evolution of reaction norms. Evolution 46: 390–411.

Gonfloni S, Di Tella L, Caldarola S, Cannata SM, Klinger FG, et al. (2009) Inhibition of the c-Abl–TAp63 pathway protects mouse oocytes from chemotherapy-induced death. Nat Med 15: 1179–85.

Gong WJ, Golic KG (2004) Genomic deletions of the Drosophila melanogaster Hsp70 genes. Genetics 168: 1467–76.

Gong Z, Morales-Ruiz T, Ariza RR, Roldán-Arjona T, David L, Zhu JK (2002) ROS1, a repressor of transcriptional gene silencing in Arabidopsis, encodes a DNA glycosylase/lyase. Cell 111: 803–14.

Gonzales GF, Gasco M, Tapia V, Gonzales-Castañeda C (2009) High serum testosterone levels are associated with excessive erythrocytosis of chronic mountain sickness in men. Am J Physiol Endocrinol Metab 296: E1319-25.

Gonzales GF, Tapia V, Gasco M, Gonzales-Castañeda C (2011) Serum testosterone levels and score of chronic mountain sickness in Peruvian men natives at 4340 m. Andrologia 43: 189-95.

Gonzalez C, Hadany L, Ponder RG, Price M, Hastings PJ, Rosenberg SM (2008) Mutability and importance of a hypermutable cell subpopulation that produces stress-induced mutants in Escherichia coli. PLoS Genet 4: e1000208.

González F, Georgieva D, Vanoli F, Shi ZD, Stadtfeld M, et al. (2013) Homologous recombination DNA repair genes play a critical role in reprogramming to a pluripotent state. Cell Rep 3: 651-60.

González G, Alemán S, Infante D (2003) Asexual genetic variability in Agave fourcroydes. II: selection among individuals in a clonally propagated population. Plant Sci 165: 595–601.

Gonzalez G, Behringer RR (2009) Dicer is required for female reproductive tract development and fertility in the mouse. Mol Reprod Dev 76: 678-88.

González J, Lenkov K, Lipatov M, Macpherson JM, Petrov DA (2008) High rate of recent transposable element-induced adaptation in Drosophila melanogaster. PLoS Biol 6: e251.

González J, Petrov DA (2009) The adaptive role of transposable elements in the Drosophila genome. Gene 448: 124–33.

González J, Macpherson JM, Petrov DA (2009) A recent adaptive transposable element insertion near highly conserved developmental loci in Drosophila melanogaster. Mol Biol Evol 26: 1949–61.

González J, Karasov TL, Messer PW, Petrov DA (2010) Genome-wide patterns of adaptation to temperate environments associated with transposable elements in Drosophila. PLoS Genet 6: e1000905.

Gonzalez-Garcia I, Solé RV, Costa J (2002) Metapopulation dynamics and spatial heterogeneity in cancer. Proc Natl Acad Sci USA 99: 13085–9.

Gonzalo S, Jaco I, Fraga MF, Chen T, Li E, et al. (2006) DNA methyltransferases control telomere length and telomere recombination in mammalian cells. Nat Cell Biol 8: 416-24.

Good BH, Rouzine IM, Balick DJ, Hallatschek O, Desai MM (2012) Distribution of fixed beneficial mutations and the rate of adaptation in asexual populations. Proc Natl Acad Sci USA 109: 4950-5.

Good DA, Wright JW (1984) Allozymes and hybrid origin of the parthenogenetic lizard Cnemidophorus exanguis. Experientia 40: 1012–4.

Good JM,Nachman MW (2005) Rates of protein evolution are positively correlated with developmental timing of expression during mouse spermatogenesis. Mol Biol Evol 22: 1044–52.

Good RDO (1931) A theory of plant geography. New Phytol 30: 149–71.

Good-Avila SV, Souza V, Gaut BS, Eguiarte LE (2006) Timing and rate of speciation in Agave (Agavaceae). Proc Natl Acad Sci USA 103: 9124-9.

Goodenough U, Lin H, Lee JH (2007) Sex determination in Chlamydomonas. Semin Cell Dev Biol 18: 350–61.

Goodier JL, Kazazian HH (2008) Retrotransposons revisited: the restraint and rehabilitaion of parasites. Cell 135: 23–35.

Goodman AE, Hild E, Marshall KC, Hermansson M (1993) Conjugative plasmid transfer between bacteria under simulated marine oligotrophic conditions. Appl Environ Microbiol 59: 1035-40.

Goodman FR, Mundlos S, Muragaki Y, Donnai D, Giovannucci-Uzielli ML, et al. (1997) Synpolydactyly phenotypes correlate with size of expansions in HOXD13 polyalanine tract. Proc Natl Acad Sci USA 94: 7458–63.

Goodman MF (2002) Error-prone repair DNA polymerases in prokaryotes and eukaryotes. Annu Rev Biochem 71: 17-50.

Goodman SD, Scocca JJ (1988) Identification and arrangement of the DNA sequence recognized in specific transformation of Neisseria gonorrhoeae. Proc Natl Acad Sci USA 85: 6982-6.

Goodnight C, Rauch E, Sayama H, de Aguiar MAM, Baranger M, Bar-Yam Y (2008) Evolution in spatial predator-prey models and the “prudent predator”: the inadequacy of steady-state organism fitness and the concept of individual and group selection. Complexity 13: 23–44.

Goodson ML, Park-Sarge OK, Sarge KD (1995) Tissue-dependent expression of heat shock factor 2 isoforms with distinct transcriptional activities. Mol Cell Biol 15: 5288-93.

Goodwillie C, Kalisz S, Eckert CG (2005) The evolutionary enigma of mixed mating systems in plants: occurrence, theoretical explanations, and empirical evidence. Annu Rev Ecol Evol Syst 36: 47–79.

Goodwin TDJ, Poulter RTM (2000) Multiple LTR retrotransposon families in the asexual yeast Candida albicans. Genome Res 10: 174–91.

Goos HJ, Consten D (2002) Stress adaptation, cortisol and pubertal development in the male common carp, Cyprinus carpio. Mol Cell Endocrinol 197: 105-16.

Gopalkrishnan K, Kholkute S, Anand Kumar TC (1987) Estimation of daily sperm production in rat and monkeys. J Biosci 12: 93–7.

Gordo I, Charlesworth B (2000) The degeneration of asexual haploid populations and the speed of Muller’s ratchet. Genetics 154: 1379–87.

Gordo I, Charlesworth B (2001) Genetic linkage and molecular evolution. Curr Biol 11: R684-6.

Gordon DM (1992) Phenotypic plasticity. In: Lloyd EA, Kell EF, eds. Keywords in evolutionary biology. Cambridge, MA: Harvard University Press. pp 255–262.

Gore A, Li Z, Fung HL, Young JE, Agarwal S, et al. (2011) Somatic coding mutations in human induced pluripotent stem cells. Nature 471: 63–7.

Gorelick R, Heng HHQ (2011) Sex reduces genetic variation: A multidisciplinary review. Evolution 65: 1088–98.

Gorelick R, Laubichler M, Massicotte R (2011) Asexuality and epigenetic variation. In: Hallgrimsson B, Hall BK, eds. Epigenetics: Linking Genotype and Phenotype in Development and Evolution. Berkeley, CA: University of California Press. pp 87-102.

Goriely A, McVean GA, Rojmyr M, Ingemarsson B, Wilkie AO (2003) Evidence for selective advantage of pathogenic FGFR2 mutations in the male germ line. Science 301: 643–6.

Goriely A, McVean GA, van Pelt AM, O’Rourke AW, Wall SA, et al. (2005) Gain-of-function amino acid substitutions drive positive selection of FGFR2 mutations in human spermatogonia. Proc Natl Acad Sci USA 102: 6051–6.

Goriely A, Wilkie AOM (2012) Paternal age effect mutations and selfish spermatogonial selection: causes and consequences for human disease. Am J Hum Genet 90: 175–200.

Görlach A (2009) Regulation of HIF-1alpha at the transcriptional level. Curr Pharm Des 15: 3844-52.

Görlach A, Bonello S (2008) The cross-talk between NF-kappaB and HIF-1: further evidence for a significant liaison. Biochem J 412: e17-9.

Gorman OT, Bean WJ, Webster RG (1992) Evolutionary processes in influenza viruses: divergence, rapid evolution, and stasis. Curr Top Microbiol Immunol 176: 75-97.

Gorodetsky P, Tannenbaum E (2008) A dual role for sex? arXiv:0809.0029v1 [q-bio.PE].

Gorokhova E, Dowling TE, Weider LJ, Crease TJ, Elser JJ (2002) Functional and ecological significance of rDNA intergenic spacer variation in a clonal organism under divergent selection for production rate. Proc R Soc B 269: 2373–9.

Gorospe M, Tominaga K, Wu X, Fahling M, Ivan M (2011) Post-transcriptional control of the hypoxic response by RNA-binding proteins and MicroRNAs. Front Mol Neurosci 4: 7.

Gosden RG, Byatt-Smith JG (1986) Oxygen concentration gradient across the ovarian follicular epithelium: model, predictions and implications. Hum Reprod 1:65-8.

Gosling LM (1986) Selective abortion of entire litters in the coypu: adaptive control of offspring production in relation to quality and sex. Am Nat 127: 772–95.

Gossmann TI, Keightley PD, Eyre-Walker A (2012) The effect of variation in the effective population size on the rate of adaptive molecular evolution in eukaryotes. Genome Biol Evol 4: 658-67.

Gostincar C, Grube M, de Hoog S, Zalar P, Gunde-Cimerman N (2010) Extremotolerance in fungi: evolution on the edge. FEMS Microbiol Ecol 71:2-11.

Goto Y, Kubota S, Kohno S (1998) Highly repetitive DNA sequences that are restricted to the germline in the hagfish Eptatretus cirrhatus: A mosaic of eliminated elements. Chromosoma 107: 17–32.

Gotthard K, Nylin S (1995) Adaptive plasticity and plasticity as adaptation: a selective review of plasticity in animal morphology and life history. Oikos 74: 3–17.

Gottipati P, Cassel TN, Savolainen L, Helleday T (2008) Transcription-associated recombination is dependent on replication in mammalian cells. Mol Cell Biol 28: 154-64.

Gottlieb Y, Zchori-Fein E, Werren JH, Karr TL (2002) Diploidy restoration in Wolbachia-infected Muscidifurax uniraptor (Hymenoptera: Pteromalidae). J Invertebr Pathol 81: 166–74.

Gottschald OR, Malec V, Krasteva G, Hasan D, Kamlah F, et al. (2010) TIAR and TIA-1 mRNA-binding proteins co-aggregate under conditions of rapid oxygen decline and extreme hypoxia and suppress the HIF-1α pathway. J Mol Cell Biol 2: 345–56.

Götz C, Montenarh M (1996) p53: DNA damage, DNA repair, and apoptosis. Rev Physiol Biochem Pharmacol 127: 65-95.

Goud P, Goud A, Van Oostveldt P, Van der Elst J, Dhont M (1999) Fertilization abnormalities and pronucleus size asynchrony after intracytoplasmic sperm injection are related to oocyte postmaturity. Fertil Steril 72: 245–52.

Gougeon A (2004) Dynamics of human follicular growth: morphologic, dynamic and functional aspects. In: Leung PKC, Addashi EY, eds. The Ovary, 2nd edn. San Diego, CA: Elsevier Academic Press. pp 25-43.

Gould E, Woolley CS, McEwen BS (1990) Short-term glucocorticoid manipulations affect neuronal morphology and survival in the adult dentate gyrus. Neuroscience 37: 367–75.

Gould SJ (2002) The structure of evolutionary theory. Cambridge, MA: Belknap Press of Harvard University Press.

Gould SJ, Eldredge N (1977) Punctuated equilibrium: the tempo and mode of evolution reconsidered. Paleobiology 3: 115–51.

Gould S, Lewontin R (1979) The spandrels of San Marco and the Panglossian paradigm: a critique of the adaptationist programme. Proc R Soc Lond B Biol Sci 205: 581–98.

Gould SJ, Eldredge N (1993) Punctuated equilibrium comes of age. Nature 366: 223–7.

Goulet I, Boisvenue S, Mokas S, Mazroui R, Côté J (2008) TDRD3, a novel Tudor domain-containing protein, localizes to cytoplasmic stress granules. Hum Mol Genet 17: 3055-74.

Gourbière S, Menu F (2009) Adaptive dynamics of dormancy duration variability: evolutionary trade-off and priority effect lead to suboptimal adaptation. Evolution 63: 1879-92.

Gouy M, Gautier C (1982) Codon usage in bacteria: correlation with gene expressivity. Nucleic Acids Res 10: 7055–74.

Govin J, Lestrat C, Caron C, Pivot-Pajot C, Rousseaux S, Khochbin S (2006) Histone acetylation-mediated chromatin compaction during mouse spermatogenesis. Ernst Schering Res Found Workshop 57: 155–72.

Govin J, Escoffier E, Rousseaux S, Kuhn L, Ferro M, et al. (2007) Pericentric heterochromatin reprogramming by new histone variants during mouse spermiogenesis. J Cell Biol 176: 283–94.

Goyal S, Balick DJ, Jerison ER, Neher RA, Shraiman BI, Desai MM (2012) Dynamic mutation-selection balance as an evolutionary attractor. Genetics 191: 1309-19.

Grad I, Cederroth CR, Walicki J, Grey C, Barluenga S, et al. (2010) The molecular chaperone Hsp90α is required for meiotic progression of spermatocytes beyond pachytene in the mouse. PLoS ONE 5: e15770.

Grafen A (1990) Sexual selection unhandicapped by the Fisher process. J Theor Biol 144: 473–516.

Grafen A (1999) Formal Darwinism, the individual-as-maximising-agent analogy, and bet-hedging. Proc R Soc Ser B 266: 799–803.

Grafen A (2006) Optimization of inclusive fitness. J Theor Biol 238: 541–63.

Graham LA, Lougheed SC, Ewart KV, Davies PL (2008) Lateral transfer of a lectin-like antifreeze protein gene in fishes. PLoS ONE 3: e2616.

Grandbastien MA (1998) Activation of plant retrotransposons under stress conditions. Trends Plant Sci 3: 181–7.

Grande-Pérez A, Sierra S, Castro MG, Domingo E, Lowenstein PR (2002) Molecular indetermination in the transition to error catastrophe: systematic elimination of lymphocytic choriomeningitis virus through mutagenesis does not correlate linearly with large increases in mutant spectrum complexity. Proc Natl Acad Sci USA 99: 12938–43.

Grandori C, Cowley SM, James LP, Eisenman RN (2000) The Myc/Max/Mad network and the transcriptional control of cell behavior. Annu Rev Cell Dev Biol 16: 653–99.

Grant PR, Grant BR (2002) Unpredictable evolution in a 30-year study of Darwin’s finches. Science 296: 707–11.

Grant PR, Grant BR (2006) Evolution of character displacement in Darwin’s finches. Science 313: 224–6.

Grant P, Grant R (2008) How and Why Species Multiply: The Radiation of Darwin's Finches. Princeton, NJ: Princeton University Press.

Grant V (1977) Organismic Evolution. San Francisco, CA: Freeman.

Grant V (1991) The evolutionary process. 2nd edn. New York, NY: Columbia University Press.

Grant-Downton RT, Dickinson HG (2006) Epigenetics and its implications for plant biology 2. The ‘epigenetic epiphany’: epigenetics, evolution and beyond. Ann Bot 97: 11–27.

Grativol C, Hemerly AS, Ferreira PC (2012) Genetic and epigenetic regulation of stress responses in natural plant populations. Biochim Biophys Acta 1819: 176-85.

Graziewicz M, Wink DA, Laval F (1996) Nitric oxide inhibits DNA ligase activity: potential mechanisms for NO-mediated DNA damage. Carcinogenesis 17: 2501-5.

Green DR (2010) Cell competition: pirates on the tangled bank. Cell Stem Cell 6: 287-8.

Green DR, Chipuk JE (2006) p53 and metabolism: Inside the TIGAR. Cell 126: 30-2.

Green P, Lipman D, Hillier L, Waterston R, States D, Claverie JM (1993) Ancient conserved regions in new gene sequences and the protein databases. Science 259: 1711–6.

Green P, Ewing B, Miller W, Thomas PJ, Program NCS, et al. (2003) Transcription-associated mutational asymmetry in mammalian evolution. Nat Genet 33: 514–7.

Green PM, Saad S, Lewis CM, Giannelli F (1999) Mutation rates in humans. I. Overall and sex-specific rates obtained from a population study of hemophilia B. Am J Hum Genet 65: 1572-9.

Green RF, Noakes DLG (1995) Is a little bit of sex as good as a lot? J Theor Biol 174: 87–96.

Greenaway JB, Connor K, Pedersen HG, Coomer BL, LaMarre J, Petrik J (2004) Vascular endothelial growth factor and its receptor, Flk-1/KDR are cytoprotective in the extravascular compartment of the ovarian follicle. Endocrinology 145: 2896–905.

Greenberg BD, Newbold JE, Sugino A (1983) Intraspecific nucleotide sequence variability surrounding the origin of replication in human mitochondrial DNA. Gene 21: 33–49.

Greenberg JT, Demple B (1988) Overproduction of peroxide-scavenging enzymes in Escherichia coli suppresses spontaneous mutagenesis and sensitivity to redox-cycling agents in oxyR-mutants. EMBO J 7: 2611-7.

Greenberg N, Wingfield JC (1987) Stress and reproduction: reciprocal relationships. In: Norris DO, Jones RE, eds. Reproductive endocrinology of fishes, amphibians and reptiles. New York, NY: Plenum Press. pp 389–426.

Greenblatt MS, Bennett WP, Hollstein M, Harris CC (1994) Mutations in the p53 tumor suppressor gene: clues to cancer etiology and molecular pathogenesis. Cancer Res 54: 4855–78.

Greene CM, Hall JE, Guilbault KR,Quinn TP (2010) Improved viability of populations with diverse life-history portfolios. Biol Lett 6: 382–86.

Greene JC, Whitworth AJ, Kuo I, Andrews LA, Feany MB,Pallanck LJ (2003) Mitochondrial pathology and apoptotic muscle degeneration in Drosophila parkin mutants. Proc Natl Acad Sci USA 100: 4078-83.

Greenslade PJM (1983) Adversity selection and the habitat template. Am Nat 122: 352–65.

Greenwald GS (1989) Temporal and topographic changes in DNA synthesis after induced follicular atresia. Biol Reprod 40: 175-81.

Greenwald GS, Terranova PF (1988) Follicular selection and its control. In: Knobil E, Neill JD, eds. The Physiology of Reproduction. New York, NY: Raven Press. pp 387–445.

Greenwood JA, Murphy-Ullrich JE (1998) Signaling of de-adhesion in cellular regulation and motility. Microsc Res Tech 43: 420-32.

Greenwood PJ, Wheeler P (1985) The evolution of sexual size dimorphism in birds and mammals: a “hot blooded hypothesis”. In: Greenwood PJ, Harvey PH, Slatkin M, eds. Evolution: Essays in Honour of John Maynard Smith. Cambridge, UK: Cambridge University Press. pp 287-299.

Greer S, Zamenhof SJ (1962) Studies on depurination of DNA by heat. J Mol Biol 4: 123–41.

Gregersen N, Hansen J, Palmfeldt J (2012) Mitochondrial proteomics – a tool for the study of metabolic disorders. J Inherit Metab Dis 35: 715–26.

Grégoire MC, Leduc F, Boissonneault G (2011) Spermiogenesis in sperm genetic integrity. In: Zini A, Agarwal A, eds. Sperm chromatin: biological and clinical applications in male infertility and assisted reproduction. New York, NY: Springer. pp 307–320.

Grégoire MC, Massonneau J, Simard O, Gouraud A, Brazeau MA, et al. (2013) Male-driven de novo mutations in haploid germ cells. Mol Hum Reprod. 2013 Apr 7. [Epub ahead of print]

Gregory WC (1955) X-ray breeding of peanuts (Arachis hypogea L.). Agron J 47: 396-9.

Greig D, Borts RH, Louis EJ (1998) The effect of sex on adaptation to high temperature in heterozygous and homozygous yeast. Proc R Soc Lond B Biol Sci 265: 1017-23.

Greives TJ, Kingma SA, Beltrami G, Hau M (2012) Melatonin delays clutch initiation in a wild songbird. Biol Lett 8: 330–2.

Grell RF (1971) Heat-induced exchange in the fourth chromosome of diploid females of Drosophila melanogaster. Genetics 69: 523–7.

Grell RF (1978) A comparison of heat and interchromosomal effects on recombination and interference in Drosophila melanogaster. Genetics 89: 65–77.

Gremer JR, Crone EE, Lesica P (2012) Are dormant plants hedging their bets? Demographic consequences of prolonged dormancy in variable environments. Am Nat 179: 315-27.

Grenfell BT, Pybus OG, Gog JR, Wood JL, Daly JM, et al. (2004) Unifying the epidemiological and evolutionary dynamics of pathogens. Science 303: 327–32.

Grewal SI, Klar AJ (1996) Chromosomal inheritance of epigenetic states in fission yeast during mitosis and meiosis. Cell 86: 95–101.

Grey C, Barthès P, Chauveau-Le Friec G, Langa F, Baudat F, et al. (2011) Mouse PRDM9 DNA-binding specificity determines sites of histone H3 lysine 4 trimethylation for initiation of meiotic recombination. PLoS Biol 9: e1001176.

Griendling KK (2004) Novel NAD(P)H oxidases in the cardiovascular system. Heart 90: 491–3.

Grier HJ, Taylor RG (1998) Testicular maturation and regression in common snook. J Fish Biol 53: 521–42.

Griffith SC, Owens IPF, Thuman KA (2002) Extra-pair paternity in birds: A review of interspecific variation and adaptive function. Mol Ecol 11: 2195–212.

Griffith TM,Watson MA (2006) Is evolution necessary for range expansion? Manipulating reproductive timing of a weedy annual transplanted beyond its range. Am Nat 167: 153–64.

Griffiths D (1998) Sampling effort, regression method, and the shape and slope of size-abundance relations. J Anim Ecol 67: 795–804.

Griffiths HI, Butlin RK (1995) A timescale for sex versus parthenogenesis: Evidence from subfossil ostracods. Proc R Soc Lond Ser B Biol Sci 260: 65–71.

Griffiths L, Dachs GU, Bicknell R, Harris AL, Stratford IJ (1997) The influence of oxygen tension and pH on the expression of platelet derived endothelial cell growth factor/thymidine phosphorylase in human breast tumor cells grown in vitro and in vivo. Cancer Res 57: 570–2.

Griggio M, Biard C, Penn DJ, Hoi H (2011) Female house sparrows "count on" male genes: experimental evidence for MHC-dependent mate preference in birds. BMC Evol Biol 11: 44.

Griggs RF (1914) Observations on the behavior of some species at the edges of their ranges. Bull Torrey Bot Club 41: 25–49.

Grimberg B, Zeyl C (2005) The effects of sex and mutation rate on adaptation in test tubes and to mouse hosts by Saccharomyces cerevisiae. Evolution 59: 431–8.

Grime JP (1977) Evidence for the existence of three primary strategies in plants and its relevance to ecological and evolutionary theory. Am Nat 111: 1169–94.

Grime JP, Thompson K, Hunt R, Hodgson JG, Cornelissen JHC, et al. (1997) Integrated screening validates primary axes of specialisation in plants. Oikos 79: 259–81.

Grisham MB, Jourd'heuil D, Wink DA (2000) Review article: chronic inflammation and reactive oxygen and nitrogen metabolism--implications in DNA damage and mutagenesis. Aliment Pharmacol Ther 14 Suppl 1: 3-9.

Grishkan I, Korol AB, Nevo E, Wasser SP (2003) Ecological stress and sex evolution in soil microfungi. Proc R Soc B Lond Ser B 270: 13–8.

Grishok A, Tabara H, Mello CC (2000) Genetic requirements for inheritance of RNAi in C. elegans. Science 287: 2494–7.

Griswold CK (2006) Pleiotropic mutation, modularity and evolvability. Evol Dev 8: 81–93.

Griswold MD (1995) Interactions between germ cells and Sertoli cells in the testis. Biol Reprod 52: 211–6.

Griswold MD (1998) The central role of Sertoli cells in spermatogenesis. Semin Cell Dev Biol 9: 411–6.

Griveau JF, Le Lannou D (1997) Reactive oxygen species and human spermatozoa: physiology and pathology. Int J Androl 20: 61–9.

Grivna ST, Pyhtila B, Lin H (2006a) MIWI associates with translational machinery and PIWI-interacting RNAs (piRNAs) in regulating spermatogenesis. Proc Natl Acad Sci USA 103: 13415-20.

Grivna ST, Beyret E, Wang Z, Lin H (2006b) A novel class of small RNAs in mouse spermatogenic cells. Genes Dev 20: 1709–14.

Groeger G, Quiney C, Cotter TG (2009) Hydrogen peroxide as a cell-survival signaling molecule. Antioxid Redox Signal 11: 2655-71.

Groenendyk J, Michalak M (2005) Endoplasmic reticulum quality control and apoptosis. Acta Biochim Pol 52: 381-95.

Grogan DW (2004) Stability and repair of DNA in hyperthermophilic Archaea. Curr Issues Mol Biol 6: 137-44.

Grogan DW, Carver GT, Drake JW (2001) Genetic fidelity under harsh conditions: Analysis of spontaneous mutation in the thermoacidophilic archaeon Sulfolobus acidocaldarius. Proc Natl Acad Sci USA 98: 7928-33.

Groot TV, Janssen A, Pallini A, Breeuwer JA (2005) Adaptation in the asexual false spider mite Brevipalpus phoenicis: evidence for frozen niche variation. Exp Appl Acarol 36: 165-76.

Grootegoed JA, Jansen R, Van der Molen HJ (1984) The role of glucose, pyruvate and lactate in ATP production by rat spermatocytes and spermatids. Biochim Biophys Acta 767: 248–56.

Grosberg RK, Strathmann RR (1998) One cell, two cell, red cell, blue cell, the persistence of a unicellular stage in multicellular life histories. Trends Ecol Evol 13: 112-6.

Grösch S, Fritz G, Kaina B (1998) Apurinic endonuclease (Ref-1) is induced in mammalian cells by oxidative stress and involved in clastogenic adaptation. Cancer Res 58: 4410–6.

Grösch S, Kaina B (1999) Transcriptional activation of apurinic/apyrimidinic endonuclease (Ape, Ref-1) by oxidative stress requires CREB. Biochem Biophys Res Commun 261: 859–63.

Gross MD, Siegel EC (1981) Incidence of mutator strains in Escherichia coli and coliforms in nature. Mutat Res 91: 107–10.

Grossman AD (1995) Genetic networks controlling the initiation of sporulation and the development of genetic competence in Bacillus subtilis. Annu Rev Genet 29: 477-508.

Großhans H, Filipowicz W (2008) Molecular biology: the expanding world of small RNAs. Nature 451: 414–6.

Grover JP (1997) Resource competition. London, UK: Chapman & Hall.

Gruber M, Mathew LK, Runge AC, Garcia JA, Simon MC (2010) EPAS1 is required for spermatogenesis in the postnatal mouse testis. Biol Reprod 82: 1227-36.

Grün D, Wang YL, Langenberger D, Gunsalus KC, Rajewsky N (2005) microRNA target predictions across seven Drosophila species and comparison to mammalian targets. PLoS Comp Biol 1: e13.

Grunewald M, Johnson S, Lu D, Wang Z, Lomberk G, et al. (2012) Mechanistic role for a novel glucocorticoid-KLF11 (TIEG2) protein pathway in stress-induced monoamine oxidase A expression. J Biol Chem 287: 24195-206.

Gruppi CM, Zakeri ZF, Wolgemuth DJ (1991) Stage and lineage-regulated expression of two hsp90 transcripts during mouse germ cell differentiation and embryogenesis. Mol Reprod Dev 28: 209-17.

Gu SG, Pak J, Guang S, Maniar JM, Kennedy S, Fire A (2012) Amplification of siRNA in Caenorhabditis elegans generates a transgenerational sequence-targeted histone H3 lysine 9 methylation footprint. Nat Genet 44: 157-64.

Gu X (1997) The age of the common ancestor of eukaryotes and prokaryotes: Statistical inferences. Mol Biol Evol 14: 861–6.

Gu ZL, Steinmetz LM, Gu X, Scharfe C, Davis RW, et al. (2003) Role of duplicate genes in genetic robustness against null mutations. Nature 421: 63–6.

Gudmand-Høyer E (1994) The clinical significance of disaccharide maldigestion. Am J Clin Nutr 59: 735–41S.

Guerrero-Bosagna C, Skinner MK (2012) Environmentally induced epigenetic transgenerational inheritance of phenotype and disease. Mol Cell Endocrinol 354: 3–8.

Guibert S, Forné T, Weber M (2012) Global profiling of DNA methylation erasure in mouse primordial germ cells. Genome Res 22: 633-41.

Guillemaud T, Mieuzet L, Simon JC (2003) Spatial and temporal genetic variability in French populations of the peach-potato aphid, Myzus persicae. Heredity 91: 143–52.

Guillette LJ Jr, Cree A, Rooney AA (1995) Biology of stress: interactions  with reproduction, immunology and intermediary metabolism. In: Warwick C, Frye FL, Murphy JB, eds. Health and Welfare of Captive Reptiles. London, UK: Chapman and Hall. pp 32–81.

Guilmatre A, Sharp AJ (2012) Parent of origin effects. Clin Genet 81: 201-9.

Gull I, Geva E, Lerner-Geva L, Lessing J, Wolman I, Amit A (1999) Anaerobic glycolysis. The metabolism of the preovulatory human oocyte. Eur J Obstet Gynecol Reprod Biol 85: 225-8.

Gumienny TL, Lambie E, Hartwieg E, Horvitz HR, Hengartner MO (1999) Genetic control of programmed cell death in the Caenorhabditis elegans hermaphrodite germline. Development 126: 1011–22.

Gunawardane LS, Saito K, Nishida KM, Miyoshi K, Kawamura Y, et al. (2007) A slicer-mediated mechanism for repeat-associated siRNA 5′ end formation in Drosophila. Science 315: 1587–90.

Guneli E, Tugyan K, Ozturk H, Gumustekin M, Cilaker S, Uysal N (2008) Effect of melatonin on testicular damage in streptozotocin-induced diabetes rats. Eur Surg Res 40: 354-60.

Guo G, Wang W, Bradley A (2004) Mismatch repair genes identified using genetic screens in Blm-deficient embryonic stem cells. Nature 429: 891-5.

Guo HH, Loeb LA (2003) Tumbling down a different pathway to genetic instability. J Clin Invest 112: 1793-5.

Guo J, Jia Y, Tao SX, Li YC, Zhang XS, et al. (2009) Expression of nitric oxide synthase during germ cell apoptosis in testis of cynomolgus monkey after testosterone and heat treatment. J Androl 30: 190-9.

Guo JU, Su Y, Zhong C, Ming GL, Song H (2011) Hydroxylation of 5-methylcytosine by TET1 promotes active DNA demethylation in the adult brain. Cell 145: 423–34.

Guo R, Yu Z, Guan J, Ge Y, Ma J, Li S, et al. (2004) Stage-specific and tissue-specific expression characteristics of differentially expressed genes during mouse spermatogenesis. Mol Reprod Dev 67: 264–72.

Guo Y, Levin HL (2010) High-throughput sequencing of retrotransposon integration provides a saturated profile of target activity in Schizosaccharomyces pombe. Genome Res 20: 239–48.

Guppy M, Withers P (1999) Metabolic depression in animals: physiological perspectives and biochemical generalizations. Biol Rev Camb Philos Soc 74: 1-40.

Gupta S, Sekhon L, Aziz N, Agarwal A (2008) The impact of oxidative stress on female reproduction and ART: an evidence-based review. In: Makrigiannakis A, Rizk B, Garcia-Velasco JA; Sallam HN, eds. New York, NY: Cambridge University Press. pp 629-642.

Gupta GS (2005) Proteomics of spermatogenesis. Berlin, Germany: Springer.

Gupta S, Haldar C (2013) Physiological crosstalk between melatonin and glucocorticoid receptor modulates T-cell mediated immune responses in a wild tropical rodent, Funambulus pennanti. J Steroid Biochem Mol Biol 134: 23-36.

Gupta SC, Siddique HR, Mathur N, Mishra RK, Mitra K, et al. (2007) Adverse effect of organophosphate compounds, dichlorvos and chlorpyrifos in the reproductive tissues of transgenic Drosophila melanogaster: 70kDa heat shock protein as a marker of cellular damage. Toxicology 238:1-14.

Guraya SS (1979) Recent advances in the morphology, cytochemistry, and function of Balbiani's vitelline body in animal oocytes. Int Rev Cytol 59: 249–321.

Gusev O, Nakahara Y, Vanyagina V, Malutina L, Cornette R, et al. (2010) Anhydrobiosis-associated nuclear DNA damage and repair in the sleeping chironomid: linkage with radioresistance. PLoS ONE 5: e14008.

Gustafson EA, Wessel GM (2010) Vasa genes: Emerging roles in the germ line and in multipotent cells. Bioessays 32: 626–37.

Gustafsson H, Söderdahl T, Jönsson G, Bratteng JO, Forsby A (2004) Insulin-like growth factor type 1 prevents hyperglycemia-induced uncoupling protein 3 down-regulation and oxidative stress. J Neurosci Res 77: 285–91.

Gustafsson L, Sutherland WJ (1988) The costs of reproduction in the collared flycatcher Ficedula albicollis. Nature 335: 813–5.

Gutierrez-Marcos JF, Dickinson HG (2012) Epigenetic reprogramming in plant reproductive lineages. Plant Cell Physiol 53: 817–23.

Gutsaeva DR, Carraway MS, Suliman HB, Demchenko IT, Shitara H, et al. (2008) Transient hypoxia stimulates mitochondrial biogenesis in brain subcortex by a neuronal nitric oxide synthase-dependent mechanism. J Neurosci 28: 2015-24.

Gutteridge JM (1994) Biological origin of free radicals and mechanisms of antioxidant protection. Chem Biol Interact 91:133-40.

Guyomarc'h C, Lumineau S, Vivien-Roels B, Richard J, Deregnaucourt S (2001) Effect of melatonin supplementation on the sexual development in European quail (Coturnix coturnix). Behav Processes 53: 121–30.

Guzmán NV, Lanteri AA, Confalonieri VA (2012) Colonization ability of two invasive weevils with different reproductive modes. Evol Ecol 26: 1371-90.

Guzy RD, Hoyos B, Robin E, Chen H, Liu L, et al. (2005) Mitochondrial complex III is required for hypoxia-induced ROS production and cellular oxygen sensing. Cell Metab 1: 401-8.

Guzy RD, Schumacker PT (2006) Oxygen sensing by mitochondria at complex III: the paradox of increased reactive oxygen species during hypoxia. Exp Physiol 91: 807-19.

Gwinner E, Hau M, Heigl S (1997) Melatonin: generation and modulation of avian circadian rhythms. Brain Res Bull 44: 439-44.

Gyllensten U, Wharton D, Josefsson A, Wilson AC (1991) Paternal inheritance of mitochondrial-DNA in mice. Nature 352: 255–7.

Gyllström M, Hansson LA (2004) Dormancy in freshwater zooplankton: Induction, termination and the importance of benthic-pelagic coupling. Aquat Sci 66: 274–95.

Ha TY, Kim HS, Shin T (2004) Expression of constitutive endothelial, neuronal and inducible nitric oxide synthase in the testis and epididymis of horse.   J Vet Med Sci 66: 351-6.

Haag CR, Ebert D (2004) A new hypothesis to explain geographical parthenogenesis. Ann Zool Fenn 41: 539–44.

Haag WR, Garton DW (1995) Variation in genotype frequencies during the life history of the bivalve, Dreissena polymorpha. Evolution 49: 1284–8.

Haag-Liautard C, Dorris M, Maside X, Macaskill S, Halligan DL, et al. (2007) Direct estimation of per nucleotide and genomic deleterious mutation rates in Drosophila. Nature 445: 82–5.

Haag-Liautard C, Coffey N, Houle D, Lynch M, Charlesworth B, Keightley PD (2008) Direct estimation of the mitochondrial DNA mutation rate in Drosophila melanogaster. PLoS Biol 6: e204.

Haaland RE, Hawkins PA, Salazar-Gonzalez J, Johnson A, Tichacek A, et al. (2009) Inflammatory genital infections mitigate a severe genetic bottleneck in heterosexual transmission of subtype A and C HIV-1. PLoS Pathog 5: e1000274.

Haber JE (1999) DNA recombination: the replication connection. Trends Biochem Sci 24: 271–5.

Habu T, Wakabayashi N, Yoshida K, Yomogida K, Nishimune Y, Morita T (2004) P53 protein interacts specifically with the meiosis-specific mammalian RecA-like protein DMC1 in meiosis. Carcinogenesis 25: 889–93.

Haccou P, Iwasa Y (1995) Optimal mixed strategies in stochastic environments. Theor Popul Biol 47: 212–43.

Haccou P, Schneider MV (2004) Modes of reproduction and the accumulation of deleterious mutations with multiplicative fitness effects. Genetics 166: 1093-104.

H-Acevedo D, Currie DJ (2003) Does climate determine broad-scale patterns of species richness? A test of the causal link by natural experiment. Global Ecol Biogeogr 12: 461–73.

Hackett RW, Lis JT (1983) Localization of the hsp83 transcript within a 3292 nucleotide sequence from the 63B heat shock locus of D. melanogaster. Nucleic Acids Res 11: 7011-30.

Hadany L, Beker T (2003a) Fitness-associated recombination on rugged adaptive landscapes. J Evol Biol 16: 862–70.

Hadany L, Beker T (2003b) On the evolutionary advantage of fitness-associated recombinaton. Genetics 165: 2167–79.

Hadany L, Otto SP (2007) The evolution of condition-dependent sex in the face of high costs. Genetics 176: 1713–27.

Hadany L, Otto SP (2009) Condition-dependent sex and the rate of adaptation. Am Nat 174(suppl.): S71–S78.

Hadchouel M, Farza H, Simon D, Tiollais P, Pourcel C (1987) Maternal inhibition of hepatitis B surface antigen gene expression in transgenic mice correlates with de novo methylation. Nature 329: 454–6.

Haddal CFB, Pombal JP, Batistic RF (1994) Natural hybridization between diploid and tetraploid species of leaf frog, genus Phyllomedusa (Amphibia). J Herpetol 28: 425–30.

Haddrill PR, Halligan DL, Tomaras D, Charlesworth B (2007) Reduced efficiency of selection in regions of the Drosophila genome that lack crossing over. Genome Biol 8: R18.

Haddrill PR, Charlesworth B (2008) Non-neutral processes drive the nucleotide composition of non-coding sequences in Drosophila. Biol Lett 4: 438–41.

Haddrill PR, Bachtrog D, Andolfatto P (2008) Positive and negative selection on noncoding DNA in Drosophila simulans. Mol Biol Evol 25: 1825–34.

Haddy JA, Pankhurst NW (1999) Stress-induced changes in concentrations of plasma sex steroids in black bream. J Fish Biol 55: 1304–16.

Haegeman A, Jones JT, Danchin EG (2011) Horizontal gene transfer in nematodes: a catalyst for plant parasitism? Mol Plant Microbe Interact 24: 879–87.

Haerty W, Jagadeeshan S, Kulathinal RJ, Wong A, Ravi Ram K, et al. (2007) Evolution in the fast lane: rapidly evolving sex-related genes in Drosophila. Genetics 177: 1321–35.

Hafez ESE (1964) Effects of high temperature on reproduction. Int J Biometeor 7: 223-30.

Haga S, Terui K, Zhang HQ, Enosawa S, Ogawa W, et al. (2003) Stat3 protects against Fas-induced liver injury by redox-dependent and -independent mechanisms. J Clin Invest 112: 989–98.

Hagen DW (1964) Evidence of adaptation to environmental temperatures in three species of Gambusia (Poeciliidae). Southwest Nat 9: 6-19.

Hagen EH, Barrett HC, Price ME (2006) Do human parents face a quantity-quality tradeoff? Evidence from a Shuar Community. Am J Phys Anthropol 130: 405–18.

Hagen TM, Ingersoll RT, Wehr CM, Lykkesfeldt J, Vinarsky V, et al. (1998) Acetyl-L-carnitine fed to old rats partially restores mitochondrial function and ambulatory activity. Proc Natl Acad Sci USA 95: 9562-6.

Hagimori T, Abe Y, Date S, Miura K (2006) The first finding of a Rickettsia bacterium associated with parthenogenesis induction among insects. Curr Microbiol 52: 97–101.

Hahn JS, Hu Z, Thiele DJ, Iyer VR (2004) Genome-wide analysis of the biology of stress responses through heat shock transcription factor. Mol Cell Biol 24: 5249–56.

Hahn MW (2008) Toward a selection theory of molecular evolution. Evolution 62: 255-65.

Haig D (1990) Brood reduction and optimal parental investment when offspring differ in quality. Am Nat 136: 550-6.

Haig D (2007) Weismann Rules! OK? Epigenetics and the Lamarckian temptation. Biol Philos 22: 415–28.

Haig D, Grafen A (1991) Genetic scrambling as a defense against meiotic drive. J Theor Biol 153: 531-58.

Hairston NG (1996) Zooplankton egg banks as biotic reservoirs in changing environments. Limnol Oceanogr 41: 1087–92.

Hairston NG, Munns WR (1984) The timing of copepod diapause as an evolutionary stable strategy. Am Nat 123: 733–51.

Hairston NG Jr, Dillon TA (1990) Fluctuating selection and response in a population of freshwater copepods. Evolution 44: 1796–805.

Hairston NGJr, Ellner S, Kearns CM (1996) Overlapping generations: the storage effect and the maintenance of biotic diversity. In: Rhodes Jr OE, Chesser RK, Smith MH, eds. Population Dynamics in Ecological Space and Time. Chicago, IL: University of Chicago Press. pp 109–145.

Hairston NG Jr, Ellner SP, Geber MA, Yoshida T, FoxJA (2005) Rapid evolution and the convergence of ecological and evolutionary time. Ecol Lett 8: 1114–27.

Hajkova P, Erhardt S, Lane N, Haaf T, El-Maarri O, et al. (2002) Epigenetic reprogramming in mouse primordial germ cells. Mech Dev 117: 15–23.

Hajkova P, Ancelin K, Waldmann T, Lacoste N, Lange UC, et al. (2008) Chromatin dynamics during epigenetic reprogramming in the mouse germ line. Nature 452: 877–881.

Hajkova P, Jeffries SJ, Lee C, Miller N, Jackson SP, Surani MA (2010) Genome-wide reprogramming in the mouse germ line entails the base excision repair pathway. Science 329: 78–82.

Hakoyama H, Nishimura T, Matsubara N, Iguchi KI (2001) Difference in parasite load and nonspecific immune reaction between sexual and gynogenetic forms of Carassius auratus. Biol J Linn Soc 72: 401–7.

Hakoyama H, Iwasa Y (2004) Coexistence of a sexual and an unisexual form stabilized by parasites. J Theor Biol 226: 185–94.

Halazonetis TD, Gorgoulis VG, Bartek J (2008) An oncogene-induced DNA damage model for cancer development. Science 319: 1352–5.

Haldane JBS (1927) A mathematical theory of natural and artificial selection. Part V: selection and mutation. Proc Camb Philos Soc 23: 838–44.

Haldane JBS (1932) The Causes of Evolution. London, UK: Longmans, Green & Co.

Haldane JBS (1935) The rate of spontaneous mutation of a human gene. J Genet 31: 317–326.

Haldane JBS (1937) The effect of variation on fitness. Am Nat 71: 337-49.

Haldane JBS (1947) The mutation rate of the gene for hemophilia and its segregation ratios in males and females. Ann Eugen 13: 262–271.

Haldane JBS (1949a) Suggestions as to quantitative measurement of rates of evolution. Evolution 3: 51-6.

Haldane JBS (1949b) Disease and evolution. Ricerca Sci 19 (Suppl. 1): 68–76.

Haldane JBS (1956) The relation between density regulation and natural selection. Proc R Soc Lond B 145: 306–8.

Hales BF, Robaire B (2010) The male germ cell as a target for toxicants. In: Richburg J, Hoyer P, eds. Reproductive and Endocrine Toxicology. Oxford, UK: Elsevier Ltd. pp 115-129.

Hales CN, Barker DJ (1992) Type 2 (non-insulin-dependent) diabetes mellitus: the thrifty phenotype hypothesis. Diabetologia 35: 595-601.

Hales DB (1992) Interleukin-1 inhibits Leydig cell steroidogenesis primarily by decreasing 17α-hydroxylase/C17-20 lyase cytochrome P450 expression. Endocrinology 131: 2165-72.

Hales DB (2002) Testicular macrophage modulation of Leydig cell steroidogenesis. J Reprod Immmunol 57: 3-18.

Hales DB, Xiong Y, Tur-Kaspa I (1992) The role of cytokines in the regulation of Leydig cell P450c17 gene expression. J Steroid Biochem Mol Biol 43: 907–17.

Hales DB, Diemer T, Hales KH (1999) Role of cytokines in testicular function. Endocrine 10: 201–17.

Hales KG (2010) Iron testes: sperm mitochondria as a context for dissecting iron metabolism. BMC Biol 8: 79.

Halfmann R, Lindquist S (2010) Epigenetics in the extreme: prions and the inheritance of environmentally acquired traits. Science 330: 629–32.

Halicka DH, Ardelt B, Li X, Melamed MR, Darzynkiewicz Z (1995) 2-deoxyglucose enhances sensitivity of human histiocytic U 935 cells to apoptosis induced by Tumor Necrosis Factor. Cancer Res 55: 444-9.

Halkett F, Plantegenest M, Prunier-Leterme N, Mieuzet L, Delmotte F, Simon JC (2005) Admixed sexual and facultatively asexual aphid lineages at mating sites. Mol Ecol 14: 325–36.

Hall AR, Colegrave N (2008) Decay of unused characters by selection and drift. J Evol Biol 21: 610–7.

Hall BG (1992) Selection-induced mutations occur in yeast. Proc Natl Acad Sci USA 89: 4300–3.

Hall BG (1998) Adaptive mutagenesis: a process that generates almost exclusively beneficial mutations. Genetica 102-103: 109-25.

Hall BG (1999) Transposable elements as activators of cryptic genes in E. coli. Genetica 107: 181-7.

Hall BG (2002) Predicting evolution by in vitro evolution requires determining evolutionary pathways. Antimicrob Agents Chemother 46: 3035–8.

Hall BK (2003) Descent with modification: the unity underlying homology and homoplasy as seen through an analysis of development and evolution. Biol Rev 78: 409-33.

Hall C, Brachat S, Dietrich FS (2005) Contribution of horizontal gene transfer to the evolution of Saccharomyces cerevisiae. Eukaryot Cell 4: 1102–15.

Hall JL, Wang X, Adamson V, Zhao Y, Gibbons GH (2001) Overexpression of Ref-1 inhibits hypoxia and tumor necrosis factor-induced endothelial cell apoptosis through nuclear factor-kappaB-independent and –dependent pathways. Circ Res 88: 1247–53.

Hall MD, Bussière LF, Hunt J, Brooks R (2008) Experimental evidence that sexual conflict influences the opportunity, form and intensity of sexual selection. Evolution 62: 2305-15.

Hallatschek O (2011) The noisy edge of traveling waves. Proc Natl Acad Sci USA 108: 1783-7.

Hallgrimsson B, Hall BK, eds. (2011) Epigenetics: Linking Genotype and Phenotype in Development and Evolution. Berkeley, CA: University of California Press.

Halligan DL, Oliver F, Eyre-Walker A, Harr B, Keightley PD (2010) Evidence for pervasive adaptive protein evolution in wild mice. PLoS Genet 6: e1000825.

Halliwell B (2007) Biochemistry of oxidative stress. Biochem Soc Trans 35: 1147-50.

Halliwell B, Gutteridge JMC (1984) Oxygen toxicity, oxygen radicals, transition metals and disease. J Biochem 219:1-14.

Halliwell B, Aruoma OI (1991) DNA damage by oxygen-derived species. Its mechanism and measurement in mammalian systems. Fed Eur Biochem Soc Lett 281: 9–19.

Halliwell B, Gutteridge JMC (1999) Free Radicals in Biology and Medicine. New York, NY: Oxford University Press.

Halliwell B, Gutteridge JMC (2007) Free radicals in biology and medicine. 4th edn. Oxford, UK: Oxford University Press.

Hallmann A (2011) Evolution of reproductive development in the volvocine algae. Sex Plant Reprod 24: 97-112.

Hallmann A, Milczarek R, Lipinski M, Kossowska E, Spodnik JH, et al. (2004) Fast perinuclear clustering of mitochondria in oxidatively stressed human choriocarcinoma cells. Folia Morphol (Warsz) 63: 407-12.

Hamada H, Seidman M, Howard BH, Gorman CM (1984) Enhanced gene expression by the poly (dT-dG) poly (dC-dA) sequence. Mol Cell Biol 4: 2622-30.

Hamblin MT, Aquadro CF (1999) DNA sequence variation and the recombinational landscape in Drosophila pseudoobscura: a study of the second chromosome. Genetics 153: 859–69.

Hambuch T, Parsch J (2005) Patterns of synonymous codon usage in Drosophila melanogaster genes with sex-biased expression. Genetics 170: 1691-700.

Hamburger V, Oppenheim RW (1982) Naturally occurring neuronal death in vertebrates. Neurosci Comment 1: 39–55.

Hamer G, Roepers-Gajadien HL, van Duyn-Goedhart A, Gademan IS, Kal HB, et al. (2003) DNA double-strand breaks and gamma-H2AX signaling in the testis. Biol Reprod 68: 628-34.

Hamilton WD (1964) The genetical evolution of social behaviour. J Theor Biol 7: 1–52.

Hamilton WD (1967) Extraordinary sex ratios. Science 156: 477–88.

Hamilton WD (1980) Sex versus non-sex versus parasite. Oikos 35: 282–90.

Hamilton WD (1993) Inbreeding in Egypt and this book: a childish perspective. In: Thornhill NW, ed. The natural history of inbreeding and outbreeding—theoretical and empirical perspectives. Chicago, IL: University of Chicago Press. pp 429–450.

Hamilton WD, Zuk M (1982) Heritable true fitness and bright birds: a role for parasites? Science 218: 384-7.

Hamilton WD, Zuk M (1989) Parasites and sexual selection. Nature 341: 289-90.

Hamilton WD, Axelrod R, Tanese R (1990) Sexual selection as an adaptation to resist parasites (a review). Proc Natl Acad Sci USA 87: 3566-73.

Hammer M, Wallwork JA (1979) A review of the world distribution of oribatid mites (Acari: Cryptostigmata) in relation to continental drift. Biol Skr Dan Vid Selsk 22: 1–31.

Hammerstein P, ed. (2003) Genetic and cultural evolution of cooperation. Cambridge, MA: MIT Press.

Hammoud SS, Nix DA, Zhang H, Purwar J, Carrell DT, Cairns BR (2009) Distinctive chromatin in human sperm packages genes for embryo development. Nature 460: 473–8.

Han JS, Boeke JD (2005) LINE-1 retrotransposons: modulators of quantity and quality of mammalian gene expression? Bioessays 27: 775-84.

Hanawalt PC, Cooper PK, Ganesan AK, Smith CA (1979) DNA repair in bacteria and mammalian cells. Annu Rev Biochem 48: 783–836.

Hancock AM, Witonsky DB, Ehler E, Alkorta-Aranburu G, Beall C, et al. (2010) Human adaptations to diet, subsistence, and ecoregion are due to subtle shifts in allele frequency. Proc Natl Acad Sci USA 107 Suppl 2: 8924–30.

Hancock JM (1999) Microsatellites and other simple sequences: genomic context and mutational mechanisms. In: Goldstein DB, Schlötterer C, eds. Microsatellites: Evolution and Applications. Oxford, UK: Oxford University Press. pp 1-9.

Hancock JT (1997) Superoxide, hydrogen peroxide and nitric oxide as signalling molecules: their production and role in disease. Br J Biomed Sci 54: 38-46.

Handa RJ, Nunley KM, Lorens SA, Louie JP, McGivern RF, Bollnow MR (1994) Androgen regulation of adrenocorticotropin and corticosterone secretion in the male rat following novelty and foot shock stressors. Physiol Behav 55: 117-24.

Hanley KA, Fisher RN, Case TJ (1995) Lower mite infestations in an asexual gecko compared with its sexual ancestors. Evolution 49: 418–26.

Hansen JM, Go YM, Jones DP (2006) Nuclear and mitochondrial compartmentation of oxidative stress and redox signaling. Annu Rev Pharmacol Toxicol 46: 215–34.

Hansen PJ (2009) Effects of heat stress on mammalian reproduction. Phil Trans R Soc B Biol Sci 364: 3341–50.

Hansen RS, Wijmenga C, Luo P, Stanek AM, Canfield TK, et al. (1999) The DNMT3B DNA methyltransferase gene is mutated in the ICF immunodeficiency syndrome. Proc Natl Acad Sci USA 96: 14412-7.

Hansen TF, Houle D (2004) Evolvability, stabilizing selection, and the problem of stasis. In: Pigliucci M, Preston K, eds. Phenotypic integration: studying the ecology and evolution of complex phenotypes. Oxford, UK: Oxford University Press. pp 130-150.

Hanson ES, Leibold EA (1998) Regulation of iron regulatory protein 1 during hypoxia and hypoxia/reoxygenation. J Biol Chem 273: 7588-93.

Hanson S, Kim E, Deppert W (2005) Redox factor 1 (Ref-1) enhances specific DNA binding of p53 by promoting p53 tetramerization. Oncogene 24: 1641–7.

Hanukoglu I (2006) Antioxidant protective mechanisms against reactive oxygen species (ROS) generated by mitochondrial P450 systems in steroidogenic cells. Drug Metab Rev 38: 171–96.

Hanyu-Nakamura K, Kobayashi S, Nakamura A (2004) Germ cell-autonomous Wunen2 is required for germline development in Drosophila embryos. Development 131: 4545–53.

Hara T, Kouno J, Nakamura K, Kusaka M, Yamaoka M (2005) Possible role of adaptive mutation in resistance to antiandrogen in prostate cancer cells. Prostate 65: 268–75.

Harada T, Koi H, Kubota T, Aso T (2004a) Haem oxygenase augments porcine granulosa cell apoptosis in vitro. J Endocrinol 181: 191-205.

Harada T, Harada C, Wang YL, Osaka H, Amanai K, et al. (2004b) Role of ubiquitin carboxy terminal hydrolase-L1 in neural cell apoptosis induced by ischemic retinal injury in vivo. Am J Pathol 164: 59–64.

Haraguchi Y, Sasaki A (1996) Host-parasite arms race in mutation modifications: Indefinite escalation despite a heavy load? J Theor Biol 183: 121-37.

Haraguchi Y, Sasaki A (2000) The evolution of parasite virulence and transmission rate in a spatially structured population. J Theor Biol 203: 85–96.

Harcourt AH, Harvey PH, Larson SG, Short RV (1981) Testis weight, body weight and breeding system in primates. Nature 293: 55-7.

Harcourt AH, Purvis A, Liles L (1995) Sperm competition: mating system, not breeding season, affects testes size of primates. Funct Ecol 9: 468–76.

Hardeland R (2005) Antioxidative protection by melatonin: multiplicity of mechanisms from radical detoxification to radical avoidance. Endocrine 27: 119–30.

Hardeland R, Fuhrberg B (1996) Ubiquitous melatonin – Presence and effects in unicells, plants and animals. Trends Comp Biochem Physiol 2: 25-45.

Hardeland R, Poeggeler B (2007) Actions of melatonin, its structural and functional analogs in the central nervous system and the significance of metabolism. Cent Nerv Syst Agents Med Chem 7: 289-303.

Hardeland R, Poeggeler B (2008) Melatonin beyond its classical functions. Open Physiol J 1: 1-22.

Hardin G (1968) The tragedy of the commons. Science 162: 1243-8.

Hardin JA, Wallace LE, Wong JF, O'Loughlin EV, Urbanski SJ, et al. (2004) Aquaporin expression is downregulated in a murine model of colitis and in patients with ulcerative colitis, Crohn’s disease and infectious colitis. Cell Tissue Res 318: 313–23.

Harding RM, Boyce AJ, Clegg JB (1992) The evolution of tandemly repetitive DNA: recombination rules. Genetics 132: 847-59.

Hardingham GE, Bading H (1998) Nuclear calcium: a key regulator of gene expression. Biometals 11: 345-58.

Hardison RC, Roskin KM, Yang S, Diekhans M, Kent WJ, et al. (2003) Covariation in frequencies of substitution, deletion, transposition, and recombination during eutherian evolution. Genome Res 13: 13–26.

Hardling R, Bergsten J (2006) Nonrandom mating preserves intrasexual polymorphism and stops population differentiation in sexual conflict. Am Nat 167: 401-9.

Hardy AC, Cheng L (1986) Studies in the distribution of insects by aerial currents. III. Insect drift over the sea. Ecol Entomol 11: 283–90.

Hardy ICW, Griffiths NT, Godfray HCJ(1992) Clutch size in a parasitoid wasp: a manipulation experiment. J Anim Ecol 61: 121–9.

Hardy K, Handyside AH, Winston RML (1989) The human blastocyst: cell number, death and allocation during late preimplantation development in vitro. Development 107: 597-604.

Hardy MP, Gao HB, Dong Q, Ge R, Wang Q, et al. (2005) Stress hormone and male reproductive function. Cell Tissue Res 322: 147-53.

Haridas V, Ni J, Meager A, Su J, Yu GL, et al. (1998) TRANK, a novel cytokine that activates NF-κB and c-Jun N-terminal kinase. J Immunol 161: 1–6.

Harley CD (2003) Abiotic stress and herbivory interact to set range limits across a two-dimensional stress gradient. Ecology 84: 1477-88.

Harper JL (1977) Population Biology of Plants. London, UK: Academic Press.

Harper JL, Lovell PH, Moore KG(1970) The shapes and sizes of seeds. Annu Rev Ecol Syst 1: 256–327.

Harper ME, Bevilacqua L, Hagopian K, Weindruch R, Ramsey JJ (2004) Aging, oxidative stress, and mitochondrial uncoupling. Acta Physiol Scand 182: 321–31.

Harradine AR, Whalley RDB (1980) Reproductive development and seedling establishment of Aristida ramosa R.Br. in northern New South Wales. Rangel J 2: 124–35.

Harrington R (1994) Aphid layer (letter). Antenna 18: 50-1.

Harris AN, Macdonald PM (2001) Aubergine encodes a Drosophila polar granule component required for pole cell formation and related to eIF2C. Development 128: 2823–32.

Harris EE (2010) Nonadaptive processes in primate and human evolution. Am J Phys Anthropol 143 Suppl 51: 13-45.

Harris EE, Meyer D (2006) The molecular signature of selection underlying human adaptations. Am J Phys Anthropol Suppl 43: 89-130.

Harris EH (1989) The Chlamydomonas Sourcebook. San Diego, CA: Academic Press.

Harris EH, Stern DB, Witman GB (2009) The Chlamydomonas sourcebook, 2nd edn. San Diego, CA: Academic Press.

Harris JR (1998) Placental endogenous retrovirus (ERV): structural, functional, and evolutionary significance. BioEssays 20: 307–16.

Harris KD, Bartlett NJ, Lloyd VK (2012) Daphnia as an emerging epigenetic model organism. Genet Res Int 2012: 147892.

Harris PD (1989) Interactions between population growth and sexual reproduction in the viviparous monogenean Gyrodactylus turnbulli Harris, 1986 from the guppy Poecilia reticulata Peters. Parasitology 98: 245-51.

Harris RA (2002) Carbohydrate metabolism. I. Major metabolic pathways and their control. In: Devlin TM, ed. Textbook of Biochemistry with Clinical Correlations. New York, NY: Wiley. pp 597–664.

Harris WE, Uller T (2009) Reproductive investment when mate quality varies: differential allocation versus reproductive compensation. Phil Trans R Soc B 364: 1039–48.

Harrison F, Buckling A (2005) Hypermutability impedes cooperation in pathogenic bacteria. Curr Biol 15: 1968-71.

Harrison F, Buckling A (2011) Wider access to genotypic space facilitates loss of cooperation in a bacterial mutator. PLoS ONE 6: e17254.

Harrison GW (1995) Comparing predator–prey models to Luckinbill’s experiment with Didinium and Paramecium. Ecology 76: 357–74.

Harrison RG, Rand DM, Wheeler WC (1985) Mitochondrial-DNA size variation within individual crickets. Science 228: 1446–8.

Harrouk W, Codrington A, Vinson R, Robaire B, Hales BF (2000) Paternal exposure to cyclophosphamide induces DNA damage and alters the expression of DNA repair genes in the rat preimplantation embryo. Mutat Res 461: 229–41.

Harshman LG, Futuyma DJ (1985) The origin and distribution of clonal diversity in Alsophila pometaria (Lepidoptera: Geometridae). Evolution 39: 315–24.

Harshman LG, Zera AJ (2007) The cost of reproduction: the devil in the details. Trends Ecol Evol 22: 80-6.

Hartfield M, Keightley PD (2012) Current hypotheses for the evolution of sex and recombination. Integr Zool 7: 192-209.

Hartl DL, Clark AG (1997) Principles of Population Genetics, 3rd edn. Sunderland, MA: Sinauer Associates.

Hartl FU (1996) Molecular chaperones in cellular protein folding. Nature 381: 571-80.

Hartl GB, Hell P (1994) Maintenance of high levels of allelic variation in spite of a severe bottleneck in population size: the brown bear (Ursus arctos) in the Western Carpathians. Biodivers Conserv 3: 546-54.

Hartl M, Mitterstiller AM, Valovka T, Breuker K, Hobmayer B, Bister K (2010) Stem cell-specific activation of an ancestral myc protooncogene with conserved basic functions in the early metazoan Hydra. Proc Natl Acad Sci USA 107: 4051-6.

Hartman JL IV, Garvik B, Hartwell L (2001) Principles for the buffering of genetic variation. Science 291: 1001–4.

Hartmann A, Wantia J, Torres JA, Heinze J (2003) Worker policing without genetic conflicts in a clonal ant. Proc Natl Acad Sci USA 100: 12836-40.

Hartshorne GM, Lyrakou S, Hamoda H, Oloto E, Ghafari F (2009) Oogenesis and cell death in human prenatal ovaries: what are the criteria for oocyte selection? Mol Hum Reprod 15: 805-19.

Harvell CD, Grosberg RK (1988) The timing of sexual maturity in clonal animals: An empirical study. Ecology 69: 1855-64.

Harvell D (2004) Ecology and evolution of host–pathogen interactions in nature. Am Nat 164: S1–S5.

Häsler J, Strub K (2006) Alu RNP and Alu RNA regulate translation initiation in vitro. Nucleic Acids Res 34: 2374–85.

Häsler J, Samuelsson T, Strub K (2007) Useful ‘junk’: Alu RNAs in the human transcriptome. Cell Mol Life Sci 64: 1793–800.

Hashimoto H, Ishikawa T, Yamaguchi K, Shiotani M, Fujisawa M (2009) Experimental ischaemia-reperfusion injury induces vascular endothelial growth factor expression in the rat testis. Andrologia 41: 216-21.

Hashimoto S, Minami N, Takakura R, Yamada M, Imai H, Kashima N (2000) Low oxygen tension during in vitro maturation is beneficial for supporting the subsequent development of bovine cumulus-oocyte complexes. Mol Reprod Dev 57: 353-60.

Hashimoto S, Minami N, Takakura R, Imai H (2002) Bovine immature oocytes acquire developmental competence during meiotic arrest in vitro. Biol Reprod 66: 1696-701.

Hashmi G, Hashmi S, Selvan S, Grewal P, Gaugler R (1997) Polymorphism in heat shock protein gene (Hsp70) in entomopathogenic nematodes (rhabditida). J Thermal Biol 22: 143–9.

Hasselquist D, Bensch S, von Schantz T (1996) Correlation between male song repertoire, extra-pair paternity and offspring survival in the great reed warbler. Nature 381: 229–32.

Hastings IM (1989) Potential germline competition in animals and its evolutionary implications. Genetics 123: 191–7.

Hastings IM (1991) Germline selection: population genetics of the sexual/asexual lifecycle. Genetics 129: 1167–76.

Hastings KEM (1996) Strong evolutionary conservation of broadly expressed protein isoforms in the troponin I gene family and other vertebrate gene families. J Mol Evol 42: 631–40.

Hastings PJ, Rosenberg SM (2002) In pursuit of a molecular mechanism for adaptive gene amplification. DNA Repair (Amst) 1: 111-23.

Hastings PJ, Slack A, Petrosino JF, Rosenberg SM (2004) Adaptive amplification and point mutation are independent mechanisms: evidence for various stress-inducible mutation mechanisms. PLoS Biol 2: e399.

Hauert C, Holmes M, Doebeli M (2006) Evolutionary games and population dynamics: maintenance of cooperation in public goods games. Proc Biol Sci 273: 3131-2.

Haugen TB, Landmark BF, Josefen GM, Hansson V, Högset A (1994) The mature form of interleukin-1α is constitutively expressed in immature male germ cells from rats. Mol Cell Endocrinol 105: R19–R23.

Hausfater G, Hrdy SB (2008) Infanticide: Comparative and evolutionary Perspectives. New Brunswick, NJ: Aldine Transaction.

Hauswirth WW, Laipis PJ (1982) Mitochondrial DNA polymorphism in a maternal lineage of Holstein cows. Proc Natl Acad Sci USA 79: 4686–90.

Havecker ER, Wallbridge LM, Fedito P, Hardcastle TJ, Baulcombe DC (2012) Metastable differentially methylated regions within Arabidopsis inbred populations are associated with modified expression of non-coding transcripts. PLoS ONE 7: e45242.

Hawkins BA, Porter EE (2001) Area and the latitudinal diversity gradient for terrestrial birds. Ecol Lett 4: 595–601.

Hawkins BA, Porter EE (2003) Relative influences of current and historical factors on mammal and bird diversity patterns in deglaciated North America. Global Ecol Biogeogr 12: 475–81.

Hawkins BA, Field R, Cornell HV, Currie DJ, Guegan J-F, et al. (2003a) Energy, water, and broad-scale geographic patterns of species richness. Ecology 84: 3105–17.

Hawkins BA, Porter EE, Diniz-Filho JAF (2003b) Productivity and history as predictors of the latitudinal diversity gradient of terrestrial birds. Ecology 84: 1608–23.

Hawkins BA, Diniz-Filho JAF (2004) ‘Latitude’ and geographic patterns. Ecography 27: 268–72.

Hawkins BA, Diniz-Filho JAF, Jaramillo CA, Soeller SA (2007) Climate, niche conservatism, and the global bird diversity gradient. Am Nat 170(Suppl.): S16–S27.

Hawkins BJ, Madesh M, Kirkpatrick CJ, Fisher AB (2007) Superoxide flux in endothelial cells via the chloride channel-3 mediates intracellular signaling. Mol Biol Cell 18: 2002–12.

Hawks J, Wang ET, Cochran GM, Harpending HC, Moyzis RK (2007) Recent acceleration of human adaptive evolution. Proc Natl Acad Sci USA 104: 20753-8.

Hayashi K, Yoshida K, Matsui Y (2005) A histone H3 methyltransferase controls epigenetic events required for meiotic prophase. Nature 438: 374–8.

Hayashi K, Chuva de Sousa Lopes SM, Kaneda M, Tang F, Hajkova P, et al. (2008) MicroRNA biogenesis is required for mouse primordial germ cell development and spermatogenesis. PLoS ONE 3: e1738.

Hayashita Y, Osada H, Tatematsu Y, Yamada H, Yanagisawa K, et al. (2005) A polycistronic microRNA cluster, miR-17-92, is overexpressed in human lung cancers and enhances cell proliferation. Cancer Res 65: 9628-32.

Hayden EC (2008) Scandal! Sex-starved and still surviving. Nature 452: 678–80.

Hayden EJ, Weikert C, Wagner A (2012) Directional selection causes decanalization in a group I ribozyme. PLoS ONE 7: e45351.

Hayward LS, Wingfield JC (2004) Maternal corticosterone is transferred to avian yolk and may alter offspring growth and adult phenotype. Gen Comp Endocrinol 135: 365–71.

Hazlerigg D, Loudon A (2008) New insights into ancient seasonal life timers. Curr Biol 18: R795-R804.

Hazzard TM, Stouffer RL (2000) Angiogenesis in ovarian follicular and luteal development. Baillieres Best Pract Res Clin Obstet Gynaecol 14: 883-900.

He H, Soncin F, Grammatikakis N, Li Y, Siganou A, et al. (2003) Elevated expression of heat shock factor (HSF) 2A stimulates HSF1-induced transcription during stress. J Biol Chem 278: 35465-75.

He XJ, Chen T, Zhu JK (2011) Regulation and function of DNA methylation in plants and animals. Cell Res 21: 442-65.

He YF, Li BZ, Li Z, Liu P, Wang Y, et al. (2011) Tet-mediated formation of 5-carboxylcytosine and its excision by TDG in mammalian DNA. Science 333: 1303–7.

Head ML, Hunt J, Brooks R (2006) Genetic association between male attractiveness and female differential allocation. Biol Lett 2: 341–4.

Heams T (2012) Selection within organisms in the nineteenth century: Wilhelm Roux's complex legacy. Prog Biophys Mol Biol 110: 24-33.

Heasman J, Quarmby J, Wylie CC (1984) The mitochondrial cloud of Xenopus oocytes: the source of germinal granule material. Dev Biol 105: 458-69.

Hebert C, Norris K, Scheper MA, Nikitakis N, Sauk JJ (2007) High mobility group A2 is a target for miRNA-98 in head and neck squamous cell carcinoma. Mol Cancer 6: 5.

Hebert PDN (1987) Genotypic characteristics of cyclic parthenogens and their obligately asexual derivatives. In: Stearns, SJ, ed. The Evolution of Sex and its Consequences. Basel, Switzerland: Birkhäuser Verlag. pp 175–195.

Hecht AD, Agan B (1972) Diversity and age relationships in Recent and Miocene bivalves. Syst Zool 21: 308–12.

Hedgecock D (1994) Does variance in reproductive success limit effective population size of marine organisms? In: Beaumont A, ed. Genetics and evolution of aquatic organisms. London, UK: Chapman and Hall. pp 122–134.

Hedges DJ, Deininger PL (2007) Inviting instability: Transposable elements, double-strand breaks, and the maintenance of genome integrity. Mutat Res 616: 46-59.

Hedges SB, Bogart JP, Maxson LR (1992) Ancestry of unisexual salamanders. Nature 356: 708–10.

Hedrick PW (1986) Genetic polymorphism in heterogeneous environments: a decade later. Annu Rev Ecol Syst 17: 535–66.

Hedrick PW (1994) Evolutionary genetics of the major histocompatibility complex. Am Nat 143: 945–64.

Hedrick PW, Kim TJ, Parker KM (2001) Parasite resistance and genetic variation in the endangered Gila topminnow. Anim Conserv 4: 103–9.

Hedrick P (2005) Large variance in reproductive success and the Ne/N ratio. Evolution 59: 1596-9.

Hedrick PW (2006) Genetic polymorphism in heterogeneous environments: The age of genomics. Annu Rev Ecol Evol Syst 37: 67–93.

Heethoff M, Domes K, Laumann M, Maraun M, Norton RA, Scheu S (2007) High genetic divergences indicate ancient separation of parthenogenetic lineages of the oribatid mite Platynothrus peltifer (Acari, Oribatida). J Evol Biol 20: 392-402.

Heethoff M, Norton RA, Scheu S, Maraun M (2009) Parthenogenesis in oribatid mites (Acari, Oribatida): evolution without sex. In: Schön I, Martens K, van Dijk P, eds. Lost sex. New York, NY: Springer; pp 241–257.

Heethoff M, Koerner L, Norton RA, Raspotnig G (2011) Tasty but protected—first evidence of chemical defense in oribatid mites. J Chem Ecol 37: 1037-43.

Hegarty MJ, Hiscock SJ (2008) Genomic clues to the evolutionary success of polyploid plants. Curr Biol 18: R435–R444.

Hegner RW (1914) The germ-cell cycle in animals. New York, NY: Macmillan.

Hegreness M, Shoresh N, Hartl D, Kishony R (2006) An equivalence principle for the incorporation of favorable mutations in asexual populations. Science 311: 1615–7.

Heidenreich E, Novotny R, Kneidinger B, Holzmann V, Wintersberger U (2003) Non-homologous end joining as an important mutagenic process in cell cycle-arrested cells. EMBO J 22: 2274–83.

Heilbronn LK, Ravussin E (2003)Calorie restriction and aging: review of the literature and implications for studies in humans.Am J Clin Nutr 78: 361-9.

Heimberg AM, Sempere LF, Moy VN, Donoghue PCJ, Peterson KJ (2008) MicroRNAs and the advent of vertebrate morphological complexity. Proc Natl Acad Sci USA 105: 2946–50.

Heininger K (1999) A unifying hypothesis of Alzheimer’s disease. I. Ageing sets the stage. Hum Psychopharmacol Clin Exp 14: 363-414.

Heininger K (2001) The deprivation syndrome is the driving force of phylogeny, ontogeny and oncogeny. Rev Neurosci 12: 217-87.

Heininger K (2002) Aging is a deprivation syndrome driven by a germ-soma conflict. Ageing Res Rev 1: 481-536.

Heininger K (2012) The germ-soma conflict theory of aging and death: Obituary to the “evolutionary theories of aging”. WebmedCentral AGING 3: WMC003275.

Heino M, Metz JAJ, Kaitala V (1998) The enigma of frequency-dependent selection. Trends Ecol Evol 13: 367–70.

Heip C (1977) On the evolution of reproductive potentials in a brackish water meiobenthic community. Mikrofauna Meeresboden 61: 105-12.

Heise K, Puntarulo S, Pörtner HO, Abele D (2003) Production of reactive oxygen species by isolated mitochondria of the Antarctic bivalve Laternula elliptica (King and Broderip) under heat stress. Comp Biochem Phys 134C: 79-90.

Heise K, Puntarulo S, Nikinmaa M, Abele D, Pörtner HO (2006) Oxidative stress during stressful heat exposure and recovery in the North Sea eelpout (Zoarces viviparus). J Exp Biol 209: 353-63.

Heisler L, Andersson M, Arnold SJ, Boake CRB, Borgia G, et al. (1987) The evolution of mating preferences and sexually selected traits: group report. In: Bradbury J, Andersson M, eds. Sexual selection: testing the alternatives. New York, NY: Wiley. pp 96–118.

Heiss RS, Schoech SJ (2012) Oxidative cost of reproduction is sex specific and correlated with reproductive effort in a cooperatively breeding bird, the Florida scrub jay. Physiol Biochem Zool 85: 499-503.

Heitman J (2006) Sexual reproduction and the evolution of microbial pathogens. Curr Biol 16: R711–R725.

Helanterä H, Uller T (2010) The Price equation and extended inheritance. Philos Theory Biol 2: 1–17.

Held T, Paprotta I, Khulan J, Hemmerlein B, Binder L, et al. (2006) Hspa4l-deficient mice display increased incidence of male infertility and hydronephrosis development. Mol Cell Biol 26: 8099-108.

Held T, Barakat AZ, Mohamed BA, Paprotta I, Meinhardt A, et al. (2011) Heat-shock protein HSPA4 is required for progression of spermatogenesis. Reproduction 142: 133-44.

Helfenstein F, Losdat S, Saladin V, Richner H (2008) Females of carotenoid-supplemented males are more faithful and produce higher quality offspring. Behav Ecol 19: 1165–72.

Helfenstein F, Losdat S, Møller AP, Blount JD, Richner H (2010) Sperm of colourful males are better protected against oxidative stress. Ecol Lett 13: 213-22.

Hellberg ME, Vacquier VD (1999) Rapid evolution of fertilization selectivity and lysin cDNA sequences in teguline gastropods. Mol Biol Evol 16: 839–48.

Hellberg ME, Moy GW, Vacquier VD (2000) Positive selection and propeptide repeats promote rapid interspecific divergence of a gastropod sperm protein. Mol Biol Evol 17: 458–66.

Helleday T (2003) Pathways for mitotic homologous recombination in mammalian cells. Mutat Res 532: 103-15.

Hellmann I, Ebersberger I, Ptak SE, Pääbo S, Przeworski M (2003) A neutral explanation for the correlation of diversity with recombination rates in humans. Am J Hum Genet 72: 1527–35.

Hellriegel B, Ward PI (1998) Complex female reproductive tract morphology: its possible use in postcopulatory female choice. J Theor Biol 190: 179–86.

Helmrich A, Ballarino M, Nudler E, Tora L (2013) Transcription-replication encounters, consequences and genomic instability. Nat Struct Mol Biol 20: 412-8.

Hempel CG (1965) Aspects of Scientific Explanation. New York, NY: The Free Press.

Hempel CG (1966) Philosophy of Natural Science. Englewood Cliffs: Prentice-Hall.

Hendrickson H, Slechta ES, Bergthorsson U, Andersson DI, Roth JR (2002) Amplification-mutagenesis: evidence that "directed" adaptive mutation and general hypermutability result from growth with a selected gene amplification. Proc Natl Acad Sci USA 99: 2164–9.

Hendriks AJ, Mulder C (2008) Scaling of offspring number and mass to plant and animal size: model and meta-analysis. Oecologia 155: 705–16.

Hendriks G, Jansen JG, Mullenders LH, de Wind N (2010) Transcription and replication: far relatives make uneasy bedfellows. Cell Cycle 9: 2300-4.

Hendry AP, Kinnison MT (1999) The pace of modern life: measuring rates of contemporary microevolution. Evolution 53: 1637–53.

Heng HH (2009) The genome-centric concept: resynthesis of evolutionary theory. BioEssays 31: 512–25.

Hengge-Aronis R (1993) Survival of hunger and stress: the role of rpoS in early stationary phase gene regulation in E. coli. Cell 72: 165-8.

Hengherr S, Worland MR, Reuner A, Brummer F, Schill RO (2009) High-temperature tolerance in anhydrobiotic tardigrades is limited by glass transition. Physiol Biochem Zool 82: 749–55.

Henke JI, Goergen D, Zheng J, Song Y, Schüttler CG, et al. (2008) microRNA-122 stimulates translation of hepatitis C virus RNA. EMBO J 27: 3300-10.

Henle ES, Linn S (1997) Formation, prevention, and repair of DNA damage by iron/hydrogen peroxide. J Biol Chem 272: 19095-8.

Henriksen K, Hakovirta H, Parvinen M (1995) Testosterone inhibits and induces apoptosis in rat seminiferous tubules in a stage-specific manner: in situ quantification in squash preparations after administration of ethane dimethane sulfonate. Endocrinology 136: 3285–91.

Henter HJ, Via S (1995) The potential for coevolution in a host-parasitoid system. I. Genetic variation within an aphid population in susceptibility to a parasitic wasp. Evolution 49: 427–38.

Henzler T, Steudle E (2000) Transport and metabolic degradation of hydrogen peroxide: model calculations and measurements with the pressure probe suggest transport of H2O2 across water channels. J Exp Bot 51: 2053–66.

Heo M, Kang L, Shakhnovich EI (2009) Emergence of species in evolutionary ‘‘simulated annealing’’. Proc Natl Acad Sci USA 106: 1869–74.

Heo M, Shakhnovich EI (2010) Interplay between pleiotropy and secondary selection determines rise and fall of mutators in stress response. PLoS Comput Biol 6: e1000710.

Hepburn PA, Margison GP, Tisdale MJ (1991) Enzymatic methylation of cytosine in DNA is prevented by adjacent O6-methylguanine residues. J Biol Chem 266: 7985–7.

Herbeck JT, Nickle DC, Learn GH, Gottlieb GS, Curlin ME, et al. (2006) Human immunodeficiency virus type 1 env evolves toward ancestral states upon transmission to a new host. J Virol 80: 1637–44.

Herberstein ME, Schneider JM, Uhl G, Michalik P (2011) Sperm dynamics in spiders. Behav Ecol 22: 692-5.

Herbert PDN, Crease TJ (1980) Clonal coexistence in Daphnia pulex (Leydig): Another planktonic paradox. Science 207: 1363–5.

Hercus MJ, Hoffmann AA (2000) Maternal and grandmaternal age influence offspring fitness in Drosophila. Proc R Soc Lond B 267: 2105–10.

Herédia F, Loreto EL, Valente VL (2004) Complex evolution of gypsy in Drosophilid species. Mol Biol Evol 21: 1831–42.

Hereford J (2009) A quantitative survey of local adaptation and fitness trade-offs. Am Nat 173: 579-88.

Hereford J, Hansen TF, Houle D (2004) Comparing strengths of directional selection: How strong is strong? Evolution 58: 2133–43.

Herin FA, Booth NH, Johnson RM (1960) Thermoregulatory effects of abdominal air sacs on spermatogenesis in domestic fowl. Am J Physiol 198: 1343-5.

Herman RK, Dworkin NB (1971) Effect of gene induction on the rate of mutagenesis by ICR-191 in Escherichia coli. J Bacteriol 106: 543–50.

Hermann A, Goyal R, Jeltsch A (2004) The Dnmt1 DNA-(cytosine-C5)-methyltransferase methylates DNA processively with high preference for hemimethylated target sites. J Biol Chem 279: 48350–9.

Hermann F (1889) Beiträge zur Histologie des Hodens. Arch Mikrosk Anat 34: 58–105.

Hermisson J, Wagner GP (2004) The population genetic theory of hidden variation and genetic robustness. Genetics 168: 2271–84.

Hermisson J, Pennings PS (2005) Soft sweeps: molecular population genetics of adaptation from standing genetic variation. Genetics 169: 2335–52.

Hernández-García D, Wood CD, Castro-Obregón S, Covarrubias L (2010) Reactive oxygen species: A radical role in development? Free Radic Biol Med 49: 130-43.

Hernández-Rodríguez M, Bückle-Ramirez LF, Espina S (2002) Temperature preference in Poecilia sphenops (Pisces, Poeciliidae). Aquac Res 33: 933-40.

Herold S, Herkert B, Eilers M (2009) Facilitating replication under stress: an oncogenic function of MYC? Nat Rev Cancer 9: 441-4.

Heron SE, Scheffer IE, Iona X, Zuberi SM, Birch R, et al. (2010) De novo SCN1A mutations in Dravet syndrome and related epileptic encephalopathies are largely of paternal origin. J Med Genet 47: 137–41.

Herre EA, West SA (1997) Conflict of interest in a mutualism: documenting the elusive fig–wasp–seed tradeoff. Proc R Soc Lond Ser B 264: 1501–7.

Herre EA, Knowlton N, Mueller UG, Rehner SA (1999) The evolution of mutualisms: exploring the paths between conflict and cooperation. Trends Ecol Evol 14: 49-53.

Herring A, Donath A, Yarmolenko M, Uslar E, Conzen C, et al. (2012) Exercise during pregnancy mitigates Alzheimer-like pathology in mouse offspring. FASEB J 26: 117-28.

Herring CD, Blattner FR (2004) Conditional lethal amber mutations in essential Escherichia coli genes. J Bacteriol 186: 2673–81.

Herron MD, Doebeli M (2013) Parallel evolutionary dynamics of adaptive diversification in Escherichia coli. PLoS Biol 11: e1001490.

Hersch-Green EI, Myburg H, Johnson MT (2012) Adaptive molecular evolution of a defence gene in sexual but not functionally asexual evening primroses. J Evol Biol 25: 1576-86.

Herschel JFW (1987) Apreliminary discourse on the study of natural philosophy. Chicago, IL: University of Chicago Press. (First published 1830)

Hersh MN, Ponder RG, Hastings PJ, Rosenberg SM (2004) Adaptive mutation and amplification in Escherichia coli: two pathways of genome adaptation under stress. Res Microbiol 155: 352-9.

Hershberg R, Petrov DA (2008) Selection on codon bias. Annu Rev Genet 42: 287–99.

Hershko A, Ciechanover A (1998) The ubiquitin system. Annu Rev Biochem 67: 425–79.

Hertel J, Lindemeyer M, Missal K, Fried C, Tanzer A, et al. (2006) The expansion of the metazoan microRNA repertoire. BMC Genomics 7: 25.

Herzig A (1987) The analysis of planktonic rotifer populations: A plea for long–term investigations. Hydrobiologia 147: 163–80.

Hess RA, Cooke PS, Bunick D, Kirby JD (1993) Adult testicular enlargement induced by neonatal hypothyroidism is accompanied by increased Sertoli and germ cell numbers. Endocrinology 132: 2607-13.

Heubel KU, Rankin DJ, Kokko H (2009) How to go extinct by mating too much: population consequences of male mate choice and efficiency in a sexual-asexual species complex. Oikos 118: 513–20.

Hey J (1999) The neutralist, the fly and the selectionist. Trends Ecol Evol 14: 35–8.

Heyland A, Hodin J, Reitzel AM (2005) Hormone signaling in evolution and development: a non-model system approach. Bioessays 27: 64–75.

Hickey DA (1982) Selfish DNA: a sexually-transmitted nuclear parasite. Genetics 101: 519–31.

Hicks WM, Kim M, Haber JE (2010) Increased mutagenesis and unique mutation signature associated with mitotic gene conversion. Science 329: 82–5.

Hidaka Y, Hasegawa M, Nakahara T, Hoshino T (1994) The entire population of Thermus thermophilus cells is always competent at any growth phase. Biosci Biotechnol Biochem 58: 1338-9.

Hiddink JG, ter Hofstede R (2008) Climate induced increases in species richness of marine fishes. Glob Change Biol 14: 453-60.

Higashi M, Takimoto G, Yamamura N (1999) Sympatric speciation by sexual selection. Nature 402: 523-6.

Higginson DM, Pitnick S (2011) Evolution of intra-ejaculate sperm interactions: do sperm cooperate? Biological Reviews 86: 249–70.

Hijri M, Sanders IR (2005) Low gene copy number shows that arbuscular mycorrhizal fungi inherit genetically different nuclei. Nature 433: 160–3.

Hilbert L (2011) Shifting gears: Thermodynamics of genetic informationstorage suggest stress-dependence of mutation rate, whichcan accelerate adaptation. arXiv:1104.1930v1 [physics.bio-ph].

Hilbricht T, Varotto S, Sgaramella V, Bartels D, Salamini F, Furini A (2008) Retrotransposons and siRNA have a role in the evolution of desiccation tolerance leading to resurrection of the plant Craterostigma plantagineum. New Phytol 179: 877–87.

Hildenbrand C, Stock T, Lange C, Rother M, Soppa J (2011) Genome copy numbers and gene conversion in methanogenic archaea. J Bacteriol 193: 734–43.

Hill AS, Foot NJ, Chaplin TL, Young BD (2000) The most frequent constitutional translocation in humans, the t(11;22) (q23;q11) is due to a highly specific alu-mediated recombination. Hum Mol Genet 9: 1525–32.

Hill GE (1991) Plumage coloration is a sexually selected indicator of male quality. Nature 350: 337-9.

Hill R, Madureira PA, Waisman DM, Lee PW (2011) DNAPKCS binding to p53 on the p21WAF1/CIP1 promoter blocks transcription resulting in cell death. Oncotarget 2: 1094-108.

Hill WG (1982a) Predictions of response to artificial selection from new mutations. Genet Res 40: 255-78.

Hill WG (1982b) Rates of change in quantitative traits from fixation of new mutations. Proc Natl Acad Sci USA 79: 142–5.

Hill WG, Robertson A (1966) The effect of linkage on limits to artificial selection. Genet Res 8: 269-94.

Hill WG, Caballero A (1992) Artificial selection experiments. Annu Rev Ecol Syst 23:287–310.

Hillebrand H (2004a) On the generality of the latitudinal diversity gradient. Am Nat 163: 192–211.

Hillebrand H (2004b) Strength, slope and variability of marine latitudinal gradients. Mar Ecol Prog Ser 273: 251–67.

Hillis DM (2007) Asexual evolution: can species exist without sex? Curr Biol 17: R543-R544.

Hine E, McGuigan K, Blows MW (2011) Natural selection stops the evolution of male attractiveness. Proc Natl Acad Sci USA 108: 3659-64.

Hines WGS, Moore WS (1981) An analysis of sex in random environments. Adv Appl Prob 13: 342-52.

Hinton GE, Nowlan SJ (1987) How learning can guide evolution. Complex Syst 1: 495–502.

Hintze A, Adami C (2008) Evolution of complex modular biological networks. PLoS Comput Biol 4: e23.

Hintze A, Adami C (2010) Modularity and anti-modularity in networks with arbitrary degree distribution. Biol Direct 5: 32.

Hirochika H, Sugimoto K, Otsuki Y, Tsugawa H, Kanda M (1996) Retrotransposons of rice involved in mutations induced by tissue culture. Proc Natl Acad Sci USA 93: 7783-8.

Hirose F, Hotta Y, Yamaguchi M, Matsukage A (1989) Difference in the expression level of DNA polymerase β among mouse tissues: high expression in the pachytene spermatocytes. Exp Cell Res 181: 169–180.

Hiroshige T, Abe K, Wada S, Kaneko M (1973) Sex difference in circadian periodicity of CRF activity in the rat hypothalamus. Neuroendocrinology 11: 306-20.

Hirota K, Murata M, Sachi Y, Nakamura H, Takeuchi J, et al. (1999) Distinct roles of thioredoxin in the cytoplasm and in the nucleus: a two-step mechanism of redox regulation of transcription factor NF-kappaB. J Biol Chem 274: 27891–7.

Hirschhorn JN (2009) Genomewide association studies—Illuminating biologic pathways. N Engl J Med 360: 1699–1701.

Hirsh AE, Fraser HB (2001) Protein dispensability and rate of evolution. Nature 411: 1046-9.

Hirshleifer J (1988) The analytics of continuing conflict. Synthese 76: 201-33.

Hitchler MJ, Domann FE (2007) An epigenetic perspective on the free radical theory of development. Free Radic Biol Med 43: 1023-36.

Hitchmough JD, Curtain H, Hammersley I, Kellow J (1996) Effect of gap width and turf type on the establishment of the Australian forb Bulbine bulbosa. Restoration Ecol 4: 25–32.

Ho CC, Hau PM, Marxer M, Poon RY (2010) The requirement of p53 for maintaining chromosomal stability during tetraploidization. Oncotarget 1: 583-95.

Ho SYW, Phillips MJ, Cooper A, Drummond AJ (2005) Time dependency of molecular rate estimates and systematic overestimation of recent divergence times. Mol Biol Evol 22: 1561–8.

Ho SYW, Saarma U, Barnett R, Haile J, Shapiro B (2008) The effect of inappropriate calibration: three case studies in molecular ecology. PLoS ONE 3: e1615.

Ho SYW, Lanfear R, Bromham L, Phillips MJ, Soubrier J, et al. (2011) Time-dependent rates of molecular evolution. Mol Ecol 20: 3087–101.

Ho YS, Gargano M, Cao J, Bronson RT, Heimler I, Hutz RJ (1998) Reduced fertility in female mice lacking copper–zinc superoxide dismutase. J Biol Chem 273: 7765–9.

Hobbs MM, Seiler A, Achtman M, Cannon JG (1994) Microevolution within a clonal population of pathogenic bacteria: recombination, gene duplication and horizontal genetic exchange in the opa gene family of Neisseria meningitidis. Mol Microbiol 12: 171–80.

Hobza P, Zahradník R, Müller-Dethlefs K (2006) The world of non-covalent interactions: 2006. Collect Czech Chem Commun 71: 443-531.

Hochedlinger K, Plath K (2009) Epigenetic reprogramming and induced pluripotency. Development 136: 509–23.

Hochereau-de Reviers MT, Lincoln GA (1978) Seasonal variation in the histology of the testis of the red deer, Cervus elaphus. J Reprod Fertil 54: 209–13.

Hoede C, Denamur E, Tenaillon O (2006) Selection acts on DNA secondary structures to decrease transcriptional mutagenesis. PLoS Genet 2: e176.

Hoeijmakers JHJ (2001) Genome maintenance mechanisms for preventing cancer. Nature 411: 366–74.

Hoekstra RF (2005) Evolutionary biology: why sex is good. Nature 434: 571-3.

Hoelzer MA, Michod RE (1991) DNA repair and the evolution of transformation in Bacillus subtilis. III. Sex with damaged DNA. Genetics 128: 215-23.

Hoenigsberg H (2002) The future of selection: individuality, the twin legacies of Lamarck & Darwin. Genet Mol Res 1: 39-50.

Hoey AS, McCormick MI (2004) Selective predation for low body condition at the larval-juvenile transition of a coral reef fish. Oecologia 139: 23-9.

Hoffman EK, Trusko SP, Murphy M, George DL (1990) An S1 nuclease-sensitive homopurine/homopyrimidine domain in the c-Ki-ras promoter interacts with a nuclear factor. Proc Natl Acad Sci USA 87: 2705-9.

Hoffman JI, Forcada J, Trathan PN, Amos W (2007) Females fur seals show active choice for males that are heterozygous and unrelated. Nature 445: 912–4.

Hoffman PF, Kaufman AJ, Halverson GP, Schrag DP (1998) A Neoproterozoic snowball Earth. Science 281: 1342–6.

Hoffmann AA, Parsons PA (1991) Evolutionary geneticsand environmental stress. Oxford, UK: Oxford University Press.

Hoffmann AA, Blows MW (1994) Species borders: ecological and evolutionary perspectives. Trends Ecol Evol 9: 223-27.

Hoffmann AA, Parsons PA (1997) Extreme environmental change and evolution. Cambridge, UK: Cambridge University Press.

Hoffmann AA, Hercus MJ (2000) Environmental stress as an evolutionary force. Bioscience 50: 217–26.

Hoffmann AA, Willi Y (2008) Detecting genetic responses to environmental change. Nat Rev Genet 9: 421-32.

Hoffmann AA, Tracy Reynolds K, Nash MA, Weeks AR (2008) A high incidence of parthenogenesis in agricultural pests. Proc R Soc B 275: 2473-81.

Hoffmann MJ, Schulz WA (2005) Causes and consequences of DNA hypomethylation in human cancer. Biochem Cell Biol 83: 296–321.

Hofmann GE, Somero GN (1995) Evidence for protein damage at environmental temperatures: seasonal changes in levels of ubiquitin conjugates and hsp70 in the intertidal mussel Mytilus trossulus. J Exp Biol 198: 1509–18.

Hofseth LJ, Khan MA, Ambrose M, Nikolayeva O, Xu-Welliver M, et al. (2003) The adaptive imbalance in base excision–repair enzymes generates microsatellite instability in chronic inflammation. J Clin Invest 112: 1887–94.

Hogan B, Constantini F, Lacey E (1986) Manipulating the Mouse Embryo. A Laboratory Manual. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press.

Höglund J, Alatalo RV (1995) Leks. Princeton, NJ: Princeton University Press.

Högstrand K, Böhme J (1997) Gene conversion of major histocompatibility complex genes in the mouse spermatogenesis is a premeiotic event. Mol Biol Cell 8: 2511–7.

Holcik M, Sonenberg N (2005) Translational control in stress and apoptosis. Nat Rev Mol Cell Biol 6: 318-27.

Holland B (2002) Sexual selection fails to promote adaptation to a new environment. Evolution 56: 721–30.

Holland B, Rice WR (1999) Experimental removal of sexual selection reverses intersexual antagonistic coevolution and removes a reproductive load. Proc Natl Acad SciUSA 96: 5083–8.

Holland JH (1995) Hidden order. How adaptation builds complexity. New York, NY: Perseus Books.

Holland JJ, Spindler K, Horodyski F, Grabau E, Nichol S, Vande Pol S (1982) Rapid evolution of RNA genomes. Science 215: 1577–85.

Holland JJ, de la Torre JC, Clarke DK, Duarte E (1991) Quantitation of relative fitness and great adaptability of clonal populations of RNA viruses. J Virol 65: 2960-7.

Holland JN, Ness JH, Boyle A, Bronstein JL (2005) Mutualisms as consumer-resource interactions. In: Barbosa P, Castellanos I, eds. Ecology of Predator-Prey Interactions. New York, NY: Oxford University Press. pp 17–33.

Holland JN, DeAngelis DL (2009) Consumer-resource theory predicts dynamic transitions between outcomes of interspecific interactions. Ecol Lett 12: 1357–66.

Holland JN, DeAngelis DL (2010) A consumer-resource approach to the density-dependent population dynamics of mutualism. Ecology 91: 1286–95.

Holland LZ, Holland ND (1992) Early development in the lancelet (=Amphioxus) Branchiostoma floridae from sperm entry through pronuclear fusion: presence of vegetal pole plasm and lack of conspicuous ooplasmic segregation. Biol Bull 182: 77-96.

Holley AK, Dhar SK, St Clair DK (2010)Manganese superoxide dismutase versus p53: the mitochondrial center. Ann NY Acad Sci 1201: 72-8.

Holliday R (1984) The biological significance of meiosis. Symp Soc Exp Biol 38: 381-94.

Holliday R, Pugh JE (1975) DNA modification mechanisms and gene activity during development. Science 187: 226–32.

Holliday R, Grigg GW (1993) DNA methylation and mutation. Mutat Res 285: 61-7.

Hollingsworth ML (2000) Evidence for massive clonal growth in the invasive weed Fallopia japonica (Japanese Knotweed). Bot J Linn Soc 133: 463–72.

Hollis B, Fierst JL, Houle D (2009) Sexual selection accelerates the elimination of a deleterious mutant in Drosophila melanogaster. Evolution 63: 324–33.

Hollis B, Houle D (2011) Populations with elevated mutation load do not benefit from the operation of sexual selection. J Evol Biol 24: 1918-26.

Hollstein M, Rice K, Greenblatt MS, Soussi T, Fuchs R, et al. (1994) Database of p53 gene somatic mutations in human tumors and cell lines. Nucleic Acids Res 22: 3551–5.

Holman L, Snook RR (2006) Spermicide, cryptic female choice and the evolution of sperm form and function. J Evol Biol 19: 1660–70.

Holman L, Snook RR (2008) A sterile sperm caste protects brother fertile sperm from female-mediated death in Drosophila pseudoobscura. Curr Biol 18: 292–6.

Holmes EC (2003a) Patterns of intra- and interhost nonsynonymous variation reveal strong purifying selection in dengue virus. J Virol 77: 11296-8.

Holmes EC (2003b) Error thresholds and the constraints to RNA virus evolution. Trends Microbiol 11: 543–6.

Holmes TH, McCormick MI (2006) Location influences size-selective predation on newly settled reef fish. Mar Ecol Prog Ser 317: 203–9.

Holmgren A (1995) Thioredoxin structure and mechanism: conformational changes on oxidation of the active-site sulfhydryls to a disulfide. Structure 3: 239–43.

Holsberger DR, Cooke PS (2005) Understanding the role of thyroid hormone in Sertoli cell development: a mechanistic hypothesis. Cell Tissue Res 322: 133-40.

Holsinger KE, Feldman MW (1983) Modifiers of mutation rate: evolutionary optimum with complete selfing. Proc Natl Acad Sci USA 80: 6732–4.

Holstein AF, Schulze W, Davidoff M (2003) Understanding spermatogenesis is a prerequisite for treatment. Reprod Biol Endocrinol 1: 107.

Holt RD (1977) Predation, apparent competition, and the structure of prey communities. Theor Pop Biol 12: 197–229.

Holt RD (2003) On the evolutionary ecology of species’ ranges. Evol Ecol Res 5: 159–78.

Holt RD, Keitt TH (2005) Species’ borders: a unifying theme in ecology. Oikos 108: 3-6.

Holt WV, Shenfield F, Leonard T, Hartman TD, North RD, Moore HD (1989) The value of sperm swimming speed measurements in assessing the fertility of human frozen semen. Hum Reprod 4: 292–7.

Hölzel M, Burger K, Mühl B, Orban M, Kellner M, Eick D (2010) The tumor suppressor p53 connects ribosome biogenesis to cell cycle control: a double-edged sword. Oncotarget 1: 43-7.

Hong X, Luense LJ, McGinnis LK, Nothnick WB, Christenson LK (2008) Dicer1 is essential for female fertility and normal development of the female reproductive system. Endocrinology 149: 6207-12.

Hongoh Y, Deevong P, Inoue T, Moriya S, Trakulnaleamsai S, et al. (2005) Intra- and interspecific comparisons of bacterial diversity and community structure support coevolution of gut microbiota and termite host. Appl Environ Microbiol 71: 6590-9.

Honkakoski P, Negishi M (2000) Regulation of cytochrome P450 (CYP) genes by nuclear receptors. Biochem J 347: 321-37.

Honnay O, Bossuyt B (2005) Prolonged clonal growth: escape route or route to extinction? Oikos 108: 427–32.

Honnay O, Bossuyt B, Jacquemyn H, Shimono A, Uchiyama K (2008) Can a seed bank maintain the genetic variation in the above ground plant population? Oikos 117: 1–5.

Hood DW, Deadman ME, Jennings MP, Bisercic M, Fleischmann RD, et al. (1996) DNA repeats identify novel virulence genes in Haemophilus influenzae. Proc Natl Acad Sci USA 93: 11121-5.

Hoogeboom D, Burgering BM (2009) Should I stay or should I go: beta-catenin decides under stress. Biochim Biophys Acta 1796: 63-74.

Hoogenboom MO, Connolly SR, Anthony KR (2008) Interactions between morphological and physiological plasticity optimize energy acquisition in corals. Ecology 89: 1144-54.

Hooijmaijers C, Rhee JY, Kwak KJ, Chung GC, Horie T, et al. (2012) Hydrogen peroxide permeability of plasma membrane aquaporins of Arabidopsis thaliana. J Plant Res 125: 147-53.

Hook EB (1983) Down syndrome rates and relaxed selection at older maternal ages. Am J Hum Genet 35: 1307–13.

Hopper KR (1999) Risk-spreading and bet-hedging in insect population biology. Annu Rev Entomol 44: 535–60.

Hõrak P, Saks L, Zilmer M, Karu U, Zilmer K (2007) Do dietary antioxidants alleviate the cost of immune activation? An experiment with greenfinches. Am Nat 170: 625–35.

Hörandl E (2006) The complex causality of geographical parthenogenesis. New Phytol 171: 525-38.

Hörandl E (2008) Evolutionary implications of self-compatibility and reproductive fitness in the apomictic Ranunculus auricomus polyploid complex (Ranunculaceae). Int J Plant Sci 169: 1219–28.

Hörandl E (2009) A combinational theory for maintenance of sex. Heredity 103: 445-57.

Hörandl E, Paun O(2007) Patterns and sources of genetic diversity in apomictic plants: implications for evolutionary potentials and ecology. In: Hörandl E, Grossniklaus U, Van Dijk P, Sharbel T, eds. Apomixis: evolution, mechanisms and perspectives. Ruggell, Liechtenstein: Gantner. pp 169-194.

Hörandl E, Cosendai AC, Temsch EM (2008) Understanding the geographic distributions of apomictic plants: a case for a pluralistic approach. Plant Ecol Divers 1:309–20.

Hori A, Yoshida M, Shibata T, Ling F (2009) Reactive oxygen species regulate DNA copy number in isolated yeast mitochondria by triggering recombination-mediated replication. Nucleic Acids Res 37: 749–61.

Horikawa I, Fujita K, Harris CC (2011) p53 governs telomere regulation feedback too, via TRF2. Aging (Albany NY) 3: 26-32.

Hormaza JI, Herrero M (1994) Gametophytic competition and selection. In: Williams EG, Clarke AE, Knox RB, eds. Genetic Control of Self-Incompatibility and Reproductive Development in Flowering Plants. Dordrecht, The Netherlands: Kluiwer Academic Publishers. pp 372-400.

Horne DJ, Danielopol DL, Martens K (1998a) Reproductive behaviour. In: Martens K, ed. Sex and parthenogenesis: evolutionary ecology of reproductive modes in non-marine ostracods. Leiden, The Netherlands: Backhuys Publishers. pp 157–96.

Horne DJ, Martens K, Mösslacher F (1998b) A short note: is there brood selection in Darwinula stevensoni? In: Crasquin-Soleau S, Braccini E, Lethiers F, eds. What about Ostracoda! Actes du 3e Congrés Européen des Ostracodologistes, Paris-Bierville, France, 8–12 Juillet 1996. Bull Centre Rech Elf Explor Prod Mem 20: 33–5.

Horne DJ, Martens K (1999) Geographical parthenogenesis in European non-marine ostracods: post-glacial invasion or Holocene stability? Hydrobiologia 391: 1–7.

Hornstein E, Shomron N (2006) Canalization of development by microRNAs. Nat Genet 38 Suppl: S20-4.

Horvath JE (2003) On gamma-ray bursts and their biological effects: a case for an extrinsic trigger of the Cambrian explosion? arXiv:astro-ph/0310034.

Horwich MD, Li C, Matranga C, Vagin V, Farley G, et al. (2007) The Drosophila RNA methyltransferase, DmHen1, modifies germline piRNAs and single-stranded siRNAs in RISC. Curr Biol 17: 1265-72.

Hosken DJ (1997) Sperm competition in bats. Proc Biol Sci 264: 385–92.

Hosken DJ (2001) Sex and death: microevolutionary trade-offs between reproductive and immune investment in dung flies. Curr Biol 11: R379-80.

Hosken DJ, Garner TWJ, Ward PI (2001) Sexual conflict selects for male and female characters. Curr Biol 11: 489-93.

Hosken DJ, Ward PI (2001) Experimental evidence for testis size evolution via sperm competition. Ecol Lett 4: 10–3.

Hosken DJ, Garner TWJ, Tregenza T, Wedell N, Ward PI (2003) Superior sperm competitors sire higher-quality young. Proc R Soc Lond B 270: 1933–8.

Hötzel MJ, Markey CM, Walkden-Brown SW, Blackberry MA, Martin GB (1998) Morphometric and endocrine analyses of the effects of nutrition on the testis of mature Merino rams. J Reprod Fertil 113: 217–30.

Houde AE, Torio AJ (1992) Effect of parasitic infection on male color pattern and female choice in guppies. Behav Ecol 3: 346-51.

Houde ED (1987) Fish early life dynamics and recruitment variability. Am Fish Soc Symp 2: 17-29.

Houde ED (1997) Patterns and trends in larval-stage growth and mortality of teleost fish. J Fish Biol 51A: 52–83.

Houghton FD, Hawkhead JA, Humpherson PG, Hogg JE, Balen AH, et al. (2002) Non-invasive amino acid turnover predicts human embryo developmental capacity. Hum Reprod 17: 999–1005.

Houle D (1989) Allozyme associated heterosis in Drosophila melanogaster. Genetics 123: 789–801.

Houle D (1991) Genetic covariances of fitness correlates: what genetic correlations are made of and why it matters. Evolution 45: 630–48.

Houle D (1992) Comparing evolvability and variability of quantitative traits. Genetics 130: 195–204.

Houle D (1998) How should we explain variation in the genetic variance of traits? Genetica 102/103: 241–53.

Houle D, Kondrashov AS (2001) Coevolution of costly mate choice and condition-dependent display of good genes. Proc R Soc Lond B Biol Sci 269: 97–104.

Houston DW, King ML (2000) Germ plasm and molecular determinants of germ cell fate. Curr Top Dev Biol 50: 155-81.

Houwing S, Kamminga LM, Berezikov E, Cronembold D, Girard A, et al. (2007) A role for Piwi and piRNAs in germ cell maintenance and transposon silencing in zebrafish. Cell 129: 69–82.

Howard G, Eiges R, Gaudet F, Jaenisch R, Eden A (2008) Activation and transposition of endogenous retroviral elements in hypomethylation induced tumors in mice. Oncogene 27: 404-8.

Howard RS, Lively CM (1994) Parasitism, mutation accumulation and the maintenance of sex. Nature 367: 554-7. Reprinted figures in Nature 368:358.

Howard RS, Lively CM (1998) The maintenance of sex by parasitism and mutation accumulation under epistatic fitness functions. Evolution 52: 604–10.

Howard RS, Lively CM (2002) The ratchet and the Red Queen: the maintenance of sex in parasites. J Evol Biol 15: 648–56.

Howard RS, Lively CM (2004) Good vs complementary genes for parasite resistance and the evolution of mate choice. BMC Evol Biol 4: 48.

Howe DK, Denver DR (2008) Muller's Ratchet and compensatory mutation in Caenorhabditis briggsae mitochondrial genome evolution. BMC Evol Biol 8: 62.

Howell N, Halvorson S, Kubacka I, McCullough DA, Bindoff LA, Turnbull DM (1992) Mitochondrial gene segregation in mammals: is the bottleneck always narrow? Hum Genet 90: 117–20.

Howell N, Kubacka I, Mackey DA (1996) How rapidly does the human mitochondrial genome evolve? Am J Hum Genet 59: 501–9.

Howell N, Smejkal CB (2000) Persistent heteroplasmy of a mutation in the human mtDNA control region: Hypermutation as an apparent consequence of simple-repeat expansion/contraction. Am J Hum Genet 66: 1589–98.

Howell N, Smejkal CB, Mackey DA, Chinnery PF, Turnbull DM, Herrnstadt C (2003) The pedigree rate of sequence divergence in the human mitochondrial genome: there is a difference between phylogenetic and pedigree rates. Am J Hum Genet 72: 659–70.

Howlett SK, Reik W (1991) Methylation levels of maternal and paternal genomes during preimplantation development. Development 113: 119–27.

Hoy RR, Hoikkala A, Kaneshiro K (1988) Hawaiian courtship songs: evolutionary innovation in communication signals of Drosophila. Science 240: 217–9.

Hoyes KP, Johnson C, Johnston RE, Lendon RG, Hendry JH, et al. (1995) Testicular toxicity of the transferrin binding radionuclide 114mIn in adult and neonatal rats. Reprod Toxicol 9: 297-305.

Hoy Jensen L, Enghoff H, Frydenberg J, Parker ED Jr (2002) Genetic diversity and the phylogeography of parthenogenesis: comparing bisexual and thelytokous populations of Nemasoma varicorne (Diplopoda: Nemasomatidae) in Denmark. Hereditas 136: 184–94.

Hsia KT, Millar MR, King S, Selfridge J, Redhead NJ, et al. (2003) DNA repair gene Ercc1 is essential for normal spermatogenesis and oogenesis and for functional integrity of germ cell DNA in the mouse. Development 130: 369–78.

Hsieh TF, Ibarra CA, Silva P, Zemach A, Eshed-Williams L, et al. (2009) Genome-wide demethylation of Arabidopsis endosperm. Science 324: 1451–4.

Hsu GW, Ober M, Carell T, Beese LS (2004) Error-prone replication of oxidatively damaged DNA by a high-fidelity DNA polymerase. Nature 431: 217-21.

Hsu K, Passey RJ, Endoh Y, Rahimi F, Youssef P, et al. (2005) Regulation of S100A8 by glucocorticoids. J Immunol 174: 2318-26.

Hu B, Gharaee-Kermani M, Wu Z, Phan SH (2010) Epigenetic regulation of myofibroblast differentiation by DNA methylation. Am J Pathol 177: 21-8.

Hu JJ, Dubin N, Kurland D, Ma BL, Roush GC (1995) The effects of hydrogen peroxide on DNA repair activities. Mutat Res 336: 193–201.

Hu W (2009) The role of p53 gene family in reproduction. Cold Spring Harb Perspect Biol 1: a001073.

Hu W, Feng Z, Teresky AK, Levine AJ (2007) p53 regulates maternal reproduction through LIF. Nature 450: 721–4.

Hu W, Zhang C, Wu R, Sun Y, Levine A, Feng Z (2010) Glutaminase 2, a novel p53 target gene regulating energy metabolism and antioxidant function. Proc Natl Acad Sci USA 107: 7455-60.

Hu W, Alvarez-Dominguez JR, Lodish HF (2012) Regulation of mammalian cell differentiation by long non-coding RNAs. EMBO Rep 13: 971–83.

Hua Z, Lv Q, Ye W, Wong CK, Cai G, et al. (2006) MiRNA-directed regulation of VEGF and other angiogenic factors under hypoxia. PLoS ONE 1: e116.

Huamani J, McMahan CA, Herbert DC, Reddick R, McCarrey JR, et al. (2004) Spontaneous mutagenesis is enhanced in Apex heterozygous mice. Mol Cell Biol 24: 8145–53.

Huang BM, Stocco DM, Hutson JC, Norman RL (1995) Corticotropin-releasing hormone stimulates steroidogenesis in mouse Leydig cells. Biol Reprod 53: 620-6.

Huang BM, Stocco DM, Li PH, Yang HY, Wu CM, Norman RL (1997) Corticotropin-releasing hormone stimulates the expression of the steroidogenic acute regulatory protein in MA-10 mouse cells. Biol Reprod 57: 547-51.

Huang F, Ning H, Xin QQ, Huang Y, Wang H, et al. (2009) Melatonin pretreatment attenuates 2-bromopropane-induced testicular toxicity in rats. Toxicology 256: 75-82.

Huang LE (2008) Carrot and stick: HIF-alpha engages c-Myc in hypoxic adaptation. Cell Death Differ 15: 672-7.

Huang LE, Arany Z, Livingston DM, Bunn HF (1996) Activation of hypoxia-inducible transcription factor depends primarily upon redox-sensitive stabilization of its alpha subunit. J Biol Chem 271: 32253–9.

Huang LE, Gu J, Schau M, Bunn HF (1998) Regulation of hypoxia-inducible factor 1α is mediated by an O2-dependent degradation domain via the ubiquitin-proteasome pathway. Proc Natl Acad Sci USA 95: 7987–92.

Huang LE, Bunn HF (2003) Hypoxia-inducible factor and its biomedical relevance. J Biol Chem 278: 19575–8.

Huang LE, Bindra RS, Glazer PM, Harris AL (2007) Hypoxia-induced genetic instability– a calculated mechanism underlying tumor progression. J Mol Med (Berl) 85: 139-48.

Huang LH, Chen B, Kang L (2007) Impact of mild temperature hardening on thermotolerance, fecundity, and Hsp gene expression in Liriomyza huidobrensis. J Insect Physiol 53: 1199-205.

Huang RP, Adamson ED (1993) Characterization of the DNA-binding properties of the early growth response-1 (Egr-1) transcription factor: evidence for modulation by a redox mechanism. DNA Cell Biol 12: 265–73.

Huang S (2008) The genetic equidistance result of molecular evolution is independent of mutation rates. J Comput Sci Syst Biol 1: 92-102.

Huang S (2009) Reprogramming cell fates: reconciling rarity with robustness. Bioessays 31: 546-60.

Huang S (2012) The molecular and mathematical basis of Waddington's epigenetic landscape: a framework for post-Darwinian biology? Bioessays 34: 149-57.

Huang W, Chang BHJ, Gu X, Hewett-Emmett D, Li W (1997) Sex differences in mutation rate in higher primates estimated from AMG intron sequences. J Mol Evol 44: 463-5.

Huang X (2003) Iron overload and its association with cancer risk in humans: evidence for iron as a carcinogenic metal. Mutat Res 533: 153–71.

Huang X, Le QT, Giaccia AJ (2010) MiR-210-micromanager of the hypoxia pathway. Trends Mol Med 16: 230–7.

Huang X, Darzynkiewicz Z (2006) Cytometric assessment of histone H2AX phosphorylation: a reporter of DNA damage. Methods Mol Biol 314: 73–80.

Huang XA, Yin H, Sweeney S, Raha D, Snyder M, Lin H (2013) A major epigenetic programming mechanism guided by piRNAs. Dev Cell 24: 502-16.

Hua-Van A, Le Rouzic A, Boutin TS, Filée J, Capy P (2011) The struggle for life of the genome’s selfish architects. Biol Direct 6: 19.

Hubbs C (1964) Interactions between bisexual fish species and its gynogenetic sexual parasite. Bull Tex Mem Mus 8: 1–72.

Hubbs C, Schlupp I (2008) Juvenile survival in a unisexual/sexual complex of mollies. Environ Biol Fishes 83: 327–30.

Huckins C (1978) The morphology and kinetics of spermatogonial degeneration in normal adult rats: an analysis using a simplified classification of the germinal epithelium. Anat Rec 190: 905–26.

Huckins C, Oakberg EF (1978) Morphological and quantitative analysis of spermatogonia in mouse testes using whole mounted seminiferous tubules. I. The normal testes. Anat Rec 192: 519–28.

Huda A, Mariño-Ramírez L, Jordan IK (2010) Epigenetic histone modifications of human transposable elements: genome defense versus exaptation. Mob DNA 1: 2.

Hudson RE, Bergthorsson U, Ochman H (2003) Transcription increases multiple spontaneous point mutations in Salmonella enterica. Nucl Acids Res 31: 4517–22.

Huettel B, Kanno T, Daxinger L, Bucher E, van der Winden J, et al. (2007) RNA-directed DNA methylation mediated by DRD1 and Pol IVb: a versatile pathway for transcriptional gene silencing in plants. Biochim Biophys Acta 1769: 358-74.

Huey RB, Hertz PE (1984) Is a jack-of-all-temperatures a master of none? Evolution 38: 441–4.

Huey RB, Gilchrist GW, Carlson ML, Berrigan D, Serra L (2000) Rapid evolution of a geographic cline in size in an introduced fly. Science 287: 308–9.

Huey RB, Carlson M, Crozier L, Frazier M, Hamilton H, et al. (2002) Plants versus animals: do they deal with stress in different ways? Integr Comp Biol 42: 415-23.

Huey RB, Hertz P, Sinervo B (2003) Behavioral drive versus behavioral inertia in evolution: a null model approach. Am Nat 161: 357–66.

Huey RB, Gilchrist GW, Hendry AP (2005) Using invasive species to study evolution: case studies with Drosophila and Salmon. In: Sax DF, Stachowicz JJ, Gaines SD, eds. Species Invasions: Insights into Ecology, Evolution and Biogeography. Sunderland, MA: Sinauer Associates. pp 139–164.

Hufbauer RA, Via S (1999) Evolution of an aphid-parasitoid interaction: variation in resistance to parasitism among aphid populations specialized on different host plants. Evolution 53: 1435–45.

Hughes AL, Hughes MK, Howell CY, Nei M (1994) Natural-selection at the class-II major histocompatibility complex loci of mammals. Philos Trans R Soc Lond B Biol Sci 346: 359-66.

Hughes AL, Hughes MK (1995) Self peptides bound by HLA class I molecules are derived from highly conserved regions of a set of evolutionarily conserved proteins. Immunogenetics 41: 257–62.

Hughes AR, Inouye BD, Johnson MTJ, Underwood N, Vellend M (2008) Ecological consequences of genetic diversity. Ecol Lett 11: 609–23.

Hughes AR, Stachowicz JJ (2004) Genetic diversity enhances the resistance of a seagrass ecosystem to disturbance. Proc Natl Acad Sci USA 101: 8998-9002.

Hughes FM Jr, Gorospe WC (1991) Biochemical identification of apoptosis (programmed cell death) in granulosa cells: evidence for a potential mechanism underlying follicular atresia. Endocrinology 129: 2415–22.

Hughes RN (1989) A functional biology of clonal animals. London, UK: Chapman & Hall.

Huh K, Kwon TH, Kim JS, Park JM (1998) Role of the hepatic xanthine oxidase in thyroid dysfunction: effect of thyroid hormones in oxidative stress in rat liver. Arch Pharmacol Res 21: 236–40.

Hühne J, Pfeiffer H, Brinkmann B (1998) Heteroplasmic substitutions in the mitochondrial DNA control region in mother and child samples. Int J Legal Med 112: 27–30.

Huie RE, Padmaja S (1993) The reaction of NO with superoxide. Free Rad Res 18: 195-9.

Huigens ME, Luck RF, Klaassen RH, Maas MF, Timmermans MJ, Stouthamer R (2000) Infectious parthenogenesis. Nature 405: 178-9.

Hulbert AJ (2003) Life, death and membrane bilayers. J Exp Biol 206: 2303–11.

Hulbert AJ (2007) Membrane fatty acids as pacemakers of animal metabolism. Lipids 42: 811-9.

Hulbert AJ (2008) The links between membrane composition, metabolic rate and lifespan. Comp Biochem Physiol A Mol Integr Physiol 150: 196-203.

Hulbert AJ (2010) Metabolism and longevity: is there a role for membrane fatty acids? Integr Comp Biol 50:808-17.

Hulbert AJ, Else PL (1999) Membranes as possible pacemakers of metabolism. J Theoret Biol 199: 257–74.

Hulbert AJ, Else PL (2000) Mechanisms underlying the cost of living in animals. Annu Rev Physiol 62: 207–35.

Hulbert AJ, Else PL (2004) Basal metabolic rate: history, regulation, and usefulness. Physiol Biochem Zool 77: 869–76.

Hulbert AJ, Else PL (2005) Membranes and the setting of energy demand. J Exp Biol 208: 1593-9.

Huleihel M, Lunenfeld E (2004) Regulation of spermatogenesis by paracrine/autocrine testicular factors. Asian J Androl 6: 259-68.

Hull MGR, Glazener CMA, Kelly NJ, Conway DI, Foster PA, et al. (1985) Population study of causes, treatment and outcome of infertility. Br Med J 291: 1693-7.

Hullé M, Pannetier D, Simon J-C, Vernon P, Frenot Y (2003) Aphids of sub-Antarctic Iles Crozet and Kerguelen: species diversity, host range and spatial distribution. Antarct Sci 15: 203–9.

Humayun MZ (1998) SOS and Mayday: multiple inducible mutagenic pathways in Escherichia coli. Mol Microbiol 30: 905-10.

Huminiecki L, Chan HY, Lui S, Poulsom R, Stamp G, Harris AL, et al. (2001) Vascular endothelial growth factor transgenic mice exhibit reduced male fertility and placental rejection. Mol Hum Reprod 7: 255–64.

Hunt C, Morimoto RI (1985) Conserved features of eukaryotic hsp70 genes revealed by comparison with the nucleotide sequence of human hsp70. Proc Natl Acad Sci USA 82: 6455-9.

Hunt G (2007) The relative importance of directional change, random walks, and stasis in the evolution of fossil lineages. Proc Natl Acad Sci USA 104: 18404–8.

Hunt J, Bussière LF, Jennions MD, Brooks R (2004) What is genetic quality? Trends Ecol Evol 19: 329–33.

Hunt J, Brooks R, Jennions MD (2005) Female mate choice as a condition-dependent life-history trait. Am Nat 166:79–92.

Hunt PA, Hassold TJ (2002) Sex matters in meiosis. Science 296: 2181-3.

Hunter FM, Birkhead TR (2002) Sperm viability and sperm competition in insects. Curr Biol 12: 121–3.

Hunter N (1999) Prion diseases and the central dogma of molecular biology. Trends Microbiol 7: 265-6.

Hunter N (2006) Meiotic recombination. In: Aguilera A, Rothstein R, eds. Topics in Current Genetics, Molecular Genetics of Recombination. Heidelberg, Germany: Springer-Verlag. pp 381-442.

Hunter RHF, Einer-Jenson N, Greve T (2006) Presence and significance of temperature gradients among different ovarian tissues. Microsc Res Tech 69: 501–7.

Huo R, He Y, Zhao C, Guo XJ, Lin M, Sha JH (2008) Identification of human spermatogenesis-related proteins by comparative proteomic analysis: a preliminary study. Fertil Steril 90: 1109–18.

Hur JH, Van Doninck K, Mandigo ML, Meselson M (2009) Degenerate tetraploidy was established before bdelloid rotifer families diverged. Mol Biol Evol 26: 375–83.

Hurst D (2009) Fundamental concepts in genetics: genetics and the understanding of selection. Nat Rev Genet 10: 83–93.

Hurst LD, Hamilton WD (1992) Cytoplasmic fusion and the nature of sexes. Proc R Soc Lond Ser B 247: 189-94.

Hurst LD, McVean G, Moore T (1996) Imprinted genes have few and small introns. Nat Genet 12: 234-7.

Hurst LD, Peck JR (1996) Recent advances in understanding of the evolution and maintenance of sex. Trends Ecol Evol 11: 46–52.

Hurst LD, Ellegren H (1998) Sex biases in the mutation rate. Trends Genet 14: 446-52.

Hurst LD, Smith NG (1999) Do essential genes evolve slowly? Curr Biol 9: 747–50.

Hurst LD, Ellegren H (2002) Human genetics: mystery of the mutagenic male. Nature 420: 365–6.

Husband BC, Schemske DW (1996) Evolution of the magnitude and timing of inbreeding depression in plants. Evolution 50: 54–70.

Hüser J, Rechenmacher CE, Blatter LA (1998) Imaging the permeability pore transition in single mitochondria. Biophys J 74: 2129-37.

Hussain SP, Harris CC (1999) p53 mutation spectrum and load: the generation of hypotheses linking the exposure of endogenous or exogenous carcinogens to human cancer. Mutat Res 428: 23–32.

Hussain SP, Amstad P, He P, Robles A, Lupold S, et al. (2004) p53-induced up-regulation of MnSOD and GPx but not catalase increases oxidative stress and apoptosis. Cancer Res 64: 2350–6.

Hussein SM, Batada NN, Vuoristo S, Ching RW, Autio R, et al. (2011) Copy number variation and selection during reprogramming to pluripotency. Nature 471: 58–62.

Hutchison CA 3rd, Newbold JE, Potter SS, Edgell MH (1974) Maternal inheritance of mammalian mitochondrial DNA. Nature 251: 536–8.

Hutson JC (1992) Secretion of tumor necrosis factor alpha by testicular macrophages. J Reprod Immunol 23: 63-72.

Hutter CM, Rand DM (1995) Competition between mitochondrial haplotypes in distinct nuclear genetic environments: Diosophila pseudoobscura vs. D. persimilis. Genetics 140: 537-48.

Huttley GA, Jakobsen IB, Wilson SR, Easteal S (2000) How important is DNA replication for mutagenesis? Mol Biol Evol 17: 929-37.

Huttunen S (2003) Reproduction of the mosses Pleurozium schreberi and Pohlia nutans in the surroundings of copper smelters at Harjavalta, SW Finland. J Bryol 25: 41–7.

Huvenne H, Smagghe G (2010) Mechanisms of dsRNA uptake in insects and potential of RNAi for pest control: a review. J Insect Physiol 56: 227-35.

Huxley J (1942) Evolution. London, UK: Allen and Unwin.

Huxley J (1947) Evolution and ethics. London, UK: Pilot Press.

Huzurbazar S, Kolesov G, Massey SE, Harris KC, Churbanov A, Liberles DA (2010) Lineage-specific differences in the amino acid substitution process. J Mol Biol 396: 1410-21.

Hwang DG, Green P (2004) Bayesian Markov chain Monte Carlo sequence analysis reveals varying neutral substitution patterns in mammalian evolution. Proc Natl Acad Sci USA 101: 13994-4001.

Hwang GS, Wang SW, Tseng WM, Yu CH, Wang PS (2007) Effect of hypoxia on the release of vascular endothelial growth factor and testosterone in mouse TM3 Leydig cells. Am J Physiol Endocrinol Metab 292: E1763-9.

Hwang GS, Chen ST, Chen TJ, Wang SW (2009) Effects of hypoxia on testosterone release in rat Leydig cells. Am J Physiol Endocrinol Metab 297: E1039–45.

Hwang IY, Li PL, Zhang LH, Piper KR, Cook DM, et al. (1994) TraI, a LuxI homolog, is responsible for production of conjugation factor, the Ti plasmid N-acylhomoserine lactone autoinducer. Proc Natl Acad Sci USA 91: 4639-43.

Hybertson BM, Gao B, Bose SK, McCord JM (2011) Oxidative stress in health and disease: the therapeutic potential of Nrf2 activation. Mol Aspects Med 32: 234–46.

Hyde WT, Crowley TJ, Baum SK, Peltier WR (2000) Neoproterozoic ‘snowball earth’ simulations with a coupled climate/icesheet model. Nature 405: 425-9.

Hyman LH (1939) North American triclad Turbellaria. IX. The priority of Dugesia Girard 1850 over Euplanaria Hesse 1897 with notes on American species of Dugesia. Trans Amer Micros Soc 58: 264–75.

Iannello RC, Kola I (2001) Oxidative stress in Down syndrome: A paradigm for the pathogenesis of neurodegenerative disorders. In: Mattson MP, ed. Pathogenesis of Neurodegenerative Disorders. Totowa, NJ: Humana Press Inc. pp 139-148.

Ibanez-Ventoso C, Vora M, Driscoll M (2008) Sequence relationships among C. elegans, D. melanogaster and human microRNAs highlight the extensive conservation of microRNAs in biology. PLoS One 3: e2818.

Igaki T (2009) Correcting developmental errors by apoptosis: lessons from Drosophila JNK signaling. Apoptosis 14: 1021-8.

Iglesias MC, G Bell (1989) The small-scale spatial distribution of male and female plants. Oecologia 80: 229–35.

Ikeda K, Nakayashiki H, Takagi M, Tosa Y, Mayama S (2001) Heat shock, copper sulfate and oxidative stress activate the retrotransposon MAGGY resident in the plant pathogenic fungus Magnaporthe grisea. Mol Genet Genomics 266: 318-25.

Ikehata H, Takatsu M, Saito Y, Ono T (2000) Distribution of spontaneous CpG-associated G:C → A:T mutations in the lacZ gene of Muta mice: effects of CpG methylation, the sequence context of CpG sites, and severity of mutations on the activity of the lacZ gene product. Environ Mol Mutagen 36: 301–11.

Ikehata H, Yanase F, Mori T, Nikaido O, Tanaka K, Ono T (2007) Mutation spectrum in UVB-exposed skin epidermis of Xpa-knockout mice: frequent recovery of triplet mutations. Environ Mol Mutagen 48: 1–13.

Ikenishi K (1998) Germ plasm in Caenorabditis elegans, Drosophila and Xenopus. Dev Growth Differ 40: 1–10.

Ikenishi K, Tanaka TS, Komiya T (1996) Spatio-temporal distribution of the protein of Xenopus vasa homologue (Xenopus vasa-like gene 1, XVLG1) in embryos. Dev Growth Differ 38: 527–35.

Iliakis G (2009) Backup pathways of NHEJ in cells of higher eukaryotes: cell cycle dependence. Radiother Oncol 92: 310–5.

Iliescu R, Cucchiarelli VE, Yanes LL, Iles JW, Reckelhoff JF (2007) Impact of androgen-induced oxidative stress on hypertension in male SHR. Am J Physiol Regul Integr Comp Physiol 292: R731-5.

Iliopoulos I, Torok I, Mechler BM (1997) The DnaJ60 gene of Drosophila melanogaster encodes a new member of the DnaJ family of proteins. Biol Chem 378: 1177-81.

Illingworth RS, Bird AP (2009) CpG islands--'a rough guide'. FEBS Lett 583: 1713-20.

Illingworth RS, Gruenewald-Schneider U, Webb S, Kerr ARW, James KD, et al. (2010) Orphan CpG islands identify numerous conserved promoters in the mammalian genome. PLoS Genet 6: e1001134.

Illmensee K, Mahowald A (1974) Transplantation of posterior polar plasm in Drosophila: induction of germ cells at the anterior pole of the egg. Proc Natl Acad Sci USA 71: 1016-20.

Ilnytskyy Y, Kovalchuk O (2011) Non-targeted radiation effects-an epigenetic connection. Mutat Res 714: 113-25.

Ilves H (2006) Stress-induced transposition of Tn4652 in Pseudomonas putida. Tartu, Estonia: Tartu University Press.

Imai H, Nakagawa Y (2003) Biological significance of phospholipid hydroperoxide glutathione peroxidase (PHGPx, GPx4) in mammalian cells. Free Radic Biol Med 34: 145–69.

Imai M, Qin J, Yamakawa N, Miyado K, Umezawa A, Takahashi Y (2012) Molecular alterations during female reproductive aging: Can aged oocytes remind youth? In: Pereira LAV, ed. Embryology - Updates and highlights on classic topics. Rijeka, Croatia: InTech. pp 1-20.

Imaizumi K, Benito A, Kiryu-Seo S, Gonzalez V, Inohara N, et al. (2004) Critical role for DP5/Harakiri, a Bcl-2 homology domain 3-only Bcl-2 family member, in axotomy-induced neuronal cell death. J Neurosci 24: 3721-5.

Imasheva AG, Loeschcke V (2004) Environmental stress and quantitative genetic variation. In: Ferrière R, Dieckmann U, Couvet D, eds. Evolutionary Conservation Genetics. Cambridge Studies in Adaptive Dynamics. Cambridge, UK: Cambridge University Press. pp 136-150.

Imhof M, Schlotterer C (2001) Fitness effects of advantageous mutations in evolving Escherichia coli populations. Proc Natl Acad Sci USA 98: 1113–7.

Imlay JA, Linn S (1987) Mutagenesis and stress responses induced in Escherichia coli by hydrogen peroxide. J Bacteriol 169: 2967-76.

Imlay JA, Linn S (1988) DNA damage and oxygen radical toxicity. Science 240: 1302-9.

Imlay JA, Chin SM, Linn S (1988) Toxic DNA damage by hydrogen peroxide through the Fenton reaction in vivo and in vitro. Science 240: 640-2.

Immenschuh S, Baumgart-Vogt E (2005) Peroxiredoxins, oxidative stress, and cell proliferation. Antioxid Redox Signal 7: 768-77.

Immler S, Moore HDM, Breed WG, Birkhead TR (2007) By hook or by crook? Morphometry, competition and cooperation in rodent sperm. PLoS ONE 2: e170.

Immler S, Calhim S, Birkhead TR (2008) Increased postcopulatory sexual selection reduces the intramale variation in sperm design. Evolution 62: 1538–43.

Inagaki H, Ohye T, Kogo H, Kato T, Bolor H, et al. (2009) Chromosomal instability mediated by non-B DNA: cruciform conformation and not DNA sequence is responsible for recurrent translocation in humans. Genome Res 19: 191–8.

Ingman M, Kaessmann H, Paabo S, Gyllensten U (2000) Mitochondrial genome variation and the origin of modern humans. Nature 408: 708–13.

Ingolia TD, Slater MR, Craig EA (1982) Saccharomyces cerevisiae contains a complex multigene family related to the major heat shock-inducible gene of Drosophila. Mol Cell Biol 2: 1388-98.

Ingram CJ, Mulcare CA, Itan Y, Thomas MG, Swallow DM (2009) Lactose digestion and the evolutionary genetics of lactase persistence. Hum Genet 124: 579–91.

Inoue M, Sato EF, Nishikawa M, Park AM, Kira Y, et al. (2003) Mitochondrial generation of reactive oxygen species and its role in aerobic life. Curr Med Chem 10: 2495-505.

Intano GW, McMahan CA, Walter RB, McCarrey JR, Walter CA (2001) Mixed spermatogenic germ cell nuclear extracts exhibit high base excision repair activity. Nucleic Acids Res 29: 1366-72.

International Chicken Genome Sequencing Consortium (2004) Sequencing and comparative analysis of the chicken genome provide unique perspectives on vertebrate evolution. Nature 432: 695-716.

International Human Genome Consortium (2001) Initial sequencing and analysis of the human genome. Nature 409: 860–921.

Inuzuka N (2000) Primitive Late Oligocene desmostylians from Japan and phylogeny of the Desmostylia. Bull Ashoro Mus Paleont 1: 91–123.

Iorio MV, Piovan C, Croce CM (2010)Interplay between microRNAs and the epigenetic machinery: An intricate network. Biochim Biophys Acta 1799: 694–701.

Irani K, Xia Y, Zweier JL, Sollott SJ, Der CJ, et al. (1997) Mitogenic signaling mediated by oxidants in Ras-transformed fibroblasts. Science 275: 1649–52.

Irisawa C, Nakada T, Kubota Y, Sasagawa I, Adachi Y, Yaguchi H (1993) Transferrin concentration in seminal plasma with special reference to serum hormone levels in infertile men. Arch Androl 30: 13-21.

Irvine DS (1998) Epidemiology and aetiology of male infertility. Hum Reprod 13 Suppl 1: 33-44.

Irvine DS, Twigg JP, Gordon EL, Fulton N, Milne PA, Aitken RJ (2000) DNA integrity in human spermatozoa: relationships with semen quality. J Androl 21: 33–44.

Isaeva VV, Reunov AA (2001) Germ plasm and germ cell line determination: role of mitochondria. Russ J Mar Biol 27: 231–7.

Isaksson C, Sheldon BC, Uller T (2011) The challenges of integrating oxidative stress into life-history biology. BioScience 61: 194-202.

Ishibashi K (2006) Aquaporin subfamily with unusual NPA boxes. Biochim Biophys Acta 1758: 989–93.

Ishibashi K, Kuwahara M, Kageyama Y, Tohsaka A, Marumo F, Sasaki S (1997) Cloning and functional expression of a second new aquaporin abundantly expressed in testis. Biochem Biophys Res Commun 237: 714–8.

Ishibashi K, Hara S, Kondo S (2009) Aquaporin water channels in mammals. Clin Exp Nephrol 13: 107-17.

Ishibashi K, Kondo S, Hara S, Morishita Y (2011) The evolutionary aspects of aquaporin family. Am J Physiol Regul Integr Comp Physiol 300: R566-76.

Ishii T, Matsuki S, Iuchi Y, Okada F, Toyosaki S, Tomita Y, Ikeda Y, Fujii J (2005) Accelerated impairment of spermatogenic cells in SOD1-knockout mice under heat stress. Free Radic Res 39: 697–705.

Ishikawa T, Morris PL (2006) Interleukin-1beta signals through a c-Jun N-terminal kinase-dependent inducible nitric oxide synthase and nitric oxide production pathway in Sertoli epithelial cells. Endocrinology 147: 5424-30.

Ispolatov I, Ackermann M, Doebeli M (2012) Division of labour and the evolution of multicellularity. Proc Biol Sci 279: 1768-76.

Israel DA, Lou AS, Blaser MJ (2000) Characteristics of Helicobacter pylori natural transformation. FEMS Microbiol Lett 186: 275-80.

Ito H, Gaubert H, Bucher E, Mirouze M, Vaillant I, Paszkowski J (2011) An siRNA pathway prevents transgenerational retrotransposition in plants subjected to stress. Nature 472: 115–9.

Ito K, Hanazawa T, Tomita K, Barnes PJ, Adcock IM (2004) Oxidative stress reduces histone deacetylase 2 activity and enhances IL-8 gene expression: role of tyrosine nitration. Biochem Biophys Res Commun 315: 240-5.

Ito S, Shen L, Dai Q, Wu SC, Collins LB, et al. (2011) Tet proteins can convert 5-methylcytosine to 5-formylcytosine and 5-carboxylcytosine. Science 333: 1300–3.

Itoh T, Martin W, Nei M (2002) Acceleration of genomic evolution caused by enhanced mutation rate in endocellular symbionts. Proc Natl Acad Sci USA 99: 12944-8.

Itoh Y, Kampf K, Pigozzi MI, Arnold AP (2009) Molecular cloning and characterization of the germline-restricted chromosome sequence in the zebra finch. Chromosoma 118: 527-36.

Iuchi Y, Okada F, Tsunoda S, Kibe N, Shirasawa N, et al. (2009) Peroxiredoxin 4 knockout results in elevated spermatogenic cell death via oxidative stress. Biochem J 419: 149–58.

Ivan M, Harris AL, Martelli F, Kulshreshtha R (2008) Hypoxia response and microRNAs. No longer two separate worlds. J Cell Mol Med 12: 1426–31.

Ivanov PL, Wadhams MJ, Roby RK, Holland MM, Weedn VW, et al. (1996) Mitochondrial DNA sequence heteroplasmy in the Grand Duke of Russia Georgij Romanov establishes the authenticity of the remains of Tsar Nicholas II. Nat Genet 12: 417–20.

Ivell R (2007) Lifestyle impact and the biology of the human scrotum. Reprod Biol Endocrinol 5: 15.

Ives PT (1950) The importance of mutation rate genes in evolution. Evolution 4: 236-52.

Ivy TM, Sakaluk SK (2005) Polyandry promotes enhanced offspring survival in decorated crickets. Evolution 59: 152-9.

Iwata T, Fujita T, Hirao N, Matsuzaki Y, Okada T, et al. (2005) Frequent immune responses to a cancer/testis antigen, CAGE, in patients with microsatellite instability-positive endometrial cancer. Clin Cancer Res 11: 3949-57.

Iyer NV, Kotch LE, Agani F, Leung SW, Laughner E, et al. (1998) Cellular and developmental control of O2 homeostasis by hypoxia-inducible factor 1 alpha. Genes Dev 12: 149–62.

Izu H, Inouye S, Fujimoto M, Shiraishi K, Naito K, Nakai A (2004) Heat-shock transcription factor 1 is involved in quality control mechanisms in male germ cells. Biol Reprod 70: 18–24.

Izzo G, Francesco A, Ferrara D, Campitiello MR, Serino I, et al. (2010) Expression of melatonin (MT1, MT2) and melatonin-related receptors in the adult rat testes and during development. Zygote 18: 257-64.

Jablonka E (2004) Epigenetic epidemiology. Int J Epidemiol 33: 929–35.

Jablonka E, Lamb M (1995) Epigenetic inheritance and evolution: the Lamarckian dimension. Oxford, UK: Oxford University Press.

Jablonka E, Lamb MJ, Avital E (1998) 'Lamarckian' mechanisms in darwinian evolution. Trends Ecol Evol 13: 206-10.

Jablonka E, Lamb MJ (2005) Evolution in four dimensions: genetic, epigenetic, behavioral, and symbolic variation in the history of life. Cambridge, MA: MIT Press.

Jablonka E, Lamb MJ (2007) Précis of evolution in four dimensions. Behav Brain Sci 30: 353–92.

Jablonka E, Lamb MJ (2008a) The epigenome in evolution: beyond the modern synthesis. Proceedings of the Novosibirsk conference, Herald of Vavilov’s Genetic Society 12: 242–54.

Jablonka E, Lamb MJ (2008b) Soft inheritance: challenging the Modern Synthesis. Genet Mol Biol 31: 389–95.

Jablonka E, Raz G (2009) Transgenerational epigenetic inheritance: prevalence, mechanisms, and implications for the study of heredity and evolution. Q Rev Biol 84: 131-76.

Jablonski D (1993) The tropics as a source of evolutionary novelty through geological time. Nature 364: 142–4.

Jablonski D, Roy K, Valentine JW (2006) Out of the tropics: evolutionary dynamics of the latitudinal diversity gradient. Science 314: 102–6.

Jablonski EM, Mattocks MA, Sokolov E, Koniaris LG, Hughes FM Jr., et al. (2007) Decreased aquaporin expression leads to increased resistance to apoptosis in hepatocellular carcinoma. Cancer Lett 250: 36–46.

Jabs T (1999) Reactive oxygen intermediates as mediators of programmed cell death in plants and animals. Biochem Pharmacol 57: 231–45.

Jack CN, Ridgeway JG, Mehdiabadi NJ, Jones EI, Edwards TA, et al. (2008) Segregate or cooperate - a study of the interaction between two species of Dictyostelium. BMC Evol Biol 8: 293.

Jackson AL, Chen R, Loeb LA (1998) Induction of microsatellite instability by oxidative DNA damage. Proc Natl Acad Sci USA 95: 12468–73.

Jackson AL, Loeb LA (2000) Microsatellite instability induced by hydrogen peroxide in Escherichia coli. Mutat Res 447: 187-98.

Jackson AL, Loeb LA (2001) The contribution of endogenous sources of DNA damage to the multiple mutations in cancer. Mutat Res 477: 7-21.

Jackson AL, Bartz SR, Schelter J, Kobayashi SV, Burchard J, et al. (2003) Expression profiling reveals off-target gene regulation by RNAi. Nat Biotechnol 21: 635-7.

Jackson JBC, Cheetham AH (1999) Tempo and mode of speciation in the sea. Trends Ecol Evol 14: 72–7.

Jackson RB, Mooney HA, Schulze ED (1997) A global budget for fine root biomass, surface area, and nutrient contents. Proc Natl Acad Sci USA 94: 7362–6.

Jackson-Grusby L, Laird PW, Magge SN, Moeller BJ, Jaenisch R (1997) Mutagenicity of 5-aza-2’-deoxycytidine is mediated by the mammalian DNA methyltransferase. Proc Natl Acad Sci USA 94: 4681–5.

Jacob C, Giles GI, Giles NM, Sies H (2003) Sulfur and selenium: the role of oxidation state in protein structure and function. Angew Chem Int Ed Engl 42: 4742–58.

Jacobs HT, Lehtinen SK, Spelbrink JN (2000) No sex please, we're mitochondria: a hypothesis on the somatic unit of inheritance of mammalian mtDNA. BioEssays 22: 564-72.

Jacobson MD (1996) Reactive oxygen species and programmed cell death. Trends Biochem Sci 21: 83–6.

Jacobson MD, Weil M, Raff MC (1997) Programmed cell death in animal development. Cell 88: 347–54.

Jacquier-Sarlin MR, Jornot L, Polla BS (1995) Differential expression and regulation of hsp70 and hsp90 by phorbol esters and heat shock. J Biol Chem 270: 14094-9.

Jacquier-Sarlin MR, Polla BS (1996) Dual regulation of heat-shock transcription factor (Hsf) activation and DNA-binding activity by H2O2: role of thioredoxin. Biochem J 318: 187–93.

Jaenike J (1978) A hypothesis to account for the maintenance of sex within populations. Evol Theory 3: 191-4.

Jaffe K (2004) Sex promotes gamete selection: A quantitative comparative study of features favoring the evolution of sex. Complexity 9: 43-51.

Jaffe K, Issa S, Daniels E, Haile D (1997) Dynamics of the emergence of genetic resistance to biocides among asexual and sexual organisms. J Theor Biol 188: 289-99.

Jaffe R (1998) First trimester utero-placental circulation: maternal-fetal interaction. J Perinat Med 26: 168-74.

Jähner D, Stuhlmann H, Stewart CL, Harbers K, Löhler J, et al. (1982) De novo methylation and expression of retroviral genomes during mouse embryogenesis. Nature 298: 623-8.

Jahnukainen K, Chrysis D, Hou M, Parvinen M, Eksborg S, Söder O (2004) Increased apoptosis occurring during the first wave of spermatogenesis is stage-specific and primarily affects midpachytene spermatocytes in the rat testis. Biol Reprod 70: 290-6.

Jain M, Arvanitis C, Chu K, Dewey W, Leonhardt E, et al. (2002) Sustained loss of a neoplastic phenotype by brief inactivation of MYC. Science 297: 102–4.

Jain SK (1978) Inheritance of phenotypic plasticity in soft chess, Bromus mollis L. (Gramineae). Experientia 34: 835-6.

Jaiswal M, LaRusso NF, Nishioka N, Nakabeppu Y, Gores GJ (2001a) Human Ogg1, a protein involved in the repair of 8-oxoguanine, is inhibited by nitric oxide. Cancer Res 61: 6388-93.

Jaiswal M, LaRusso NF, Shapiro RA, Billiar TR, Gores GJ (2001b) Nitric oxide-mediated inhibition of DNA repair potentiates oxidative DNA damage in cholangiocytes. Gastroenterology 120: 190-9.

Jakobsson L, Franco CA, Bentley K, Collins RT, Ponsioen B, et al. (2010) Endothelial cells dynamically compete for the tip cell position during angiogenic sprouting. Nat Cell Biol 12: 943-53.

Jambhekar A, Amon A (2008) Control of meiosis by respiration. Curr Biol 18: 969–75.

Jang JY, Rhee JY, Chung GC, Kang H (2012) Aquaporin as a membrane transporter of hydrogen peroxide in plant response to stresses. Plant Signal Behav 7: 1180-1.

Jang YY, Sharkis SJ (2007) A low level of reactive oxygen species selects for primitive hematopoietic stem cells that may reside in the low-oxygenic niche. Blood 110: 3056–63.

Janic A, Mendizabal L, Llamazares S, Rossell D, Gonzalez C (2010) Ectopic expression of germline genes drives malignant brain tumor growth in Drosophila. Science 330: 1824–7.

Janko K, Bohlen J, Lamatsch D, Flajshans M, Epplen JT, et al. (2007) The gynogenetic reproduction of diploid and triploid hybrid spined loaches (Cobitis : Teleostei), and their ability to establish successful clonal lineages - on the evolution of polyploidy in asexual vertebrates. Genetica 131: 185-94.

Janko K, Drozd P, Eisner J (2011) Do clones degenerate over time? Explaining the genetic variability of asexuals through population genetic models. Biol Direct 6: 17.

Jann P, Ward PI (1999) Maternal effects and their consequences for offspring fitness in the yellow dung fly. Funct Ecol 13: 51-58.

Jannes P, Spiessens C, van der Auwera I, D’Hooghe T, Verhoeven G, Vanderschueren D (1998) Male subfertility induced by acute scrotal heating affects embryo quality in normal female mice. Hum Reprod 13: 372–5.

Jannini EA, Ulisse S, Piersanti D, Carosa E, Muzi P, et al. (1993) Early thyroid hormone treatment in rats increases testis size and germ cell number. Endocrinology 132: 2726–8.

Jansen RP (2000) Germline passage of mitochondria: quantitative considerations and possible embryological sequelae. Hum Reprod 15 Suppl 2: 112-28.

Jansen RP, de Boer K (1998) The bottleneck: mitochondrial imperatives in oogenesis and ovarian follicular fate. Mol Cell Endocrinol 145: 81-8.

Jansen RP, Burton GJ (2004) Mitochondrial dysfunction in reproduction. Mitochondrion 4: 577-600.

Jansen VA, Yoshimura J (1998) Populations can persist in an environment consisting of sink habitats only. Proc Natl Acad Sci USA 95: 3696-8.

Janssen-Heininger YMW, Mossman BT, Heintz NH, Forman HJ, Kalyanaraman B, et al. (2008) Redox-based regulation of signal transduction: principles, pitfalls, and promises. Free Radic Biol Med 45: 1–17.

Janssens PM, Van Haastert PJ (1987) Molecular basis of transmembrane signal transduction in Dictyostelium discoideum. Microbiol Rev 51: 396–418.

Janssens S, Burns K, Tschopp J, Beyaert R (2002)Regulation of interleukin-1- and lipopolysaccharide-induced NF-kappaB activation by alternative splicing of MyD88.Curr Biol 12: 467-71.

Jansson R, Davies TJ (2008) Global variation in diversification rates of flowering plants: energy vs. climate change. Ecol Lett 11: 173-83.

Jansson T, Powell TL (2007) Role of the placenta in fetal programming: underlying mechanisms and potential interventional approaches. Clin Sci (Lond) 113: 1-13.

Jaramillo N, Domingo E, Muñoz-Egea MC, Tabarés E, Gadea I (2013) Evidence of Muller's ratchet in herpes simplex virus type 1. J Gen Virol 94: 366-75.

Jaramillo-Lambert A, Ellefson M, Villeneuve AM, Engebrecht J (2007) Differential timing of S phases, X chromosome replication, and meiotic prophase in the C. elegans germ line. Dev Biol 308: 206–21.

Jaramillo-Lambert A, Harigaya Y, Vitt J, Villeneuve A, Engebrecht J (2010) Meiotic errors activate checkpoints that improve gamete quality without triggering apoptosis in male germ cells. Curr Biol 20: 2078–89.

Jarmer H, Berka R, Knudsen S, Saxild HH (2002) Transcriptome analysis documents induced competence of Bacillus subtilis during nitrogen limiting conditions. FEMS Microbiol Lett 206: 197–200.

Jarosz DF, Lindquist S (2010) Hsp90 and environmental stress transform the adaptive value of natural genetic variation. Science 330: 1820-4.

Jarosz DF, Taipale M, Lindquist S (2010) Protein homeostasis and the phenotypic manifestation of genetic diversity: Principles and mechanisms. Annu Rev Genet 44: 189-216.

Jaroudi S, SenGupta S (2007) DNA repair in mammalian embryos. Mutat Res 635: 53–77.

Jasinska A, Krzyzosiak WJ (2004) Repetitive sequences that shape the human transcriptome. FEBS Lett 567: 136-41.

Jaspers P, Kangasjärvi J (2010) Reactive oxygen species in abiotic stress signaling. Physiol Plant 138: 405-13.

Jayaraman L, Murthy KG, Zhu C, Curran T, Xanthoudakis S, Prives C (1997) Identification of redox/repair protein Ref-1 as a potent activator of p53. Genes Dev 11: 558–70.

Jayaraman R (2011) Hypermutation and stress adaptation in bacteria. J Genet 90: 383–91.

Jazin EE, Cavelier L, Eriksson I, Oreland L, Gyllensten U (1996) Human brain contains high levels of heteroplasmy in the noncoding regions of mitochondrial DNA. Proc Natl Acad Sci USA 93: 12382–7.

Jedlicka P, Mortin MA, Wu C (1997) Multiple functions of Drosophila heat shock transcription factor in vivo. EMBO J 16: 2452–62.

Jee HJ, Ko WH (1998) Phytophora cactorum can synthesize substances needed for sexual reproduction but requires a stress factor to trigger the process. Microbiology 144: 1071-5.

Jeffreys AJ, Murray J, Neumann R (1998) High-resolution mapping of crossovers in human sperm defines a minisatellite-associated recombination hotspot. Mol Cell 2: 267-73.

Jeffreys AJ, May CA (2004) Intense and highly localized gene conversion activity in human meiotic crossover hot spots. Nat Genet 36: 151–6.

Jeggo PA, Lobrich M (2005) Artemis links ATM to double strand end rejoining. Cell Cycle 4: 359-62.

Jegou B (1992) The Sertoli cell. Baillieres Clin Endocrinol Metab 6: 273–311.

Jenkins GM, Worobey M, Woelk CH, Holmes EC (2001) Evidence for the non-quasispecies evolution of RNA viruses. Mol Biol Evol 18: 987-94.

Jennions MD, Passmore NI (1993) Sperm competition in frogs: testis size and a `sterile male' experiment on Chiromantis xerampelina (Rhacophoridae). Biol J Linn Soc 50: 211-20.

Jennions MD, Petrie M (1997) Variation in mate choice and mating preferences: a review of causes and consequences. Biol Rev Camb Philos Soc 72: 283–327.

Jennions MD, Petrie M (2000) Why do females mate multiply? A review of the genetic benefits. Biol Rev Camb Philos Soc 75: 21–64.

Jennions MD, Møller AP, Petrie M (2001) Sexually selected traits and adult survival: a meta-analysis. Q Rev Biol 76: 3–36.

Jensen JD, Bachtrog D (2011) Characterizing the influence of effective population size on the rate of adaptation: Gillespie's Darwin Domain. Genome Biol Evol 3: 687-701.

Jensen JL, Pedersen AMK (2000) Probabilistic models of DNA sequence evolution with context dependent rates of substitution. Adv Appl Probab 32: 499–517.

Jensen TK, Bonde JP, Joffe M (2006) The influence of occupational exposure on male reproductive function. Occup Med 56: 544–53.

Jenuth JP, Peterson AC, Fu K, Shoubridge EA (1996) Random genetic drift in the female germline explains the rapid segregation of mammalian mitochondrial DNA. Nat Genet 14: 146–51.

Jenuwein T, Allis CD (2001) Translating the histone code. Science 293: 1074-80.

Jeon B, MuraokaW, Sahin O, Zhang Q (2008) Role of Cj1211 in natural transformation and transfer of antibiotic resistance determinants in Campylobacter jejuni. Antimicrob Agents Chemother 52: 2699–708.

Jerling L (1985) Are plants and animals alike? A note on evolutionary plant population ecology. Oikos 45: 150–3.

Jerne N (1955) The natural selection theory of antibody formation. Proc Natl Acad Sci USA 41: 849-57.

Jerzak G, Bernard KA, Kramer LD, Ebel GD (2005) Genetic variation in West Nile virus from naturally infected mosquitoes and birds suggests quasispecies structure and strong purifying selection. J Gen Virol 86: 2175-83.

Jerzak GV, Brown I, Shi PY, Kramer LD, Ebel GD (2008) Genetic diversity and purifying selection in West Nile virus populations are maintained during host switching. Virology 374: 256-60.

Jeschke JM, Kokko H (2009) The roles of body size and phylogeny in fast and slow life histories. Evol Ecol 23: 867–78.

Jeulin C, Soufir JC, Weber P, Laval-Martin D, Calvayrac R (1989) Catalase activity in human spermatozoa and seminal plasma. Gamete Res 24: 185-96.

Ježek P, Hlavatá L (2005) Mitochondria in homeostasis of reactive oxygen species in cell, tissues, and organism. Int J Biochem Cell Biol 37: 2478-503.

Ji L, Chen X (2012) Regulation of small RNA stability: methylation and beyond. Cell Res 22: 624-36.

Jia Y, Sinha-Hikim AP, Lue YH, Swerdloff RS, Vera Y, et al. (2007) Signaling pathways for germ cell death in adult cynomolgus monkeys (Macaca fascicularis) induced by mild testicular hyperthermia and exogenous testosterone treatment. Biol Reprod 77: 83-92.

Jiang C, Wright RJ, El-Zik KM, Paterson AH (1998) Polyploid formation created unique avenues for response to selection in Gossypium (cotton). Proc Natl Acad Sci USA 95: 4419-24.

Jiang J, Serinkan BF, Tyurina YY, Borisenko GG, Mi Z, et al. (2003) Peroxidation and externalization of phosphatidylserine associated with release of cytochrome c from mitochondria. Free Radic Biol Med 35: 814-25.

Jiang JY, Cheung CK, Wang Y, Tsang BK (2003) Regulation of cell death and cell survival gene expression during ovarian follicular development and atresia. Front Biosci 8: d222–d237.

Jiang N, Bao Z, Zhang X, Hirochika H, Eddy SR, et al. (2003) An active DNA transposon family in rice. Nature 421: 163-7.

Jiang N, Feschotte C, Zhang X, Wessler SR (2004) Using rice to understand the origin and amplification of miniature inverted repeat transposable elements (MITEs). Curr Opin Plant Biol 7: 115–9.

Jiang S, Zhang LF, Zhang HW, Hu S, Lu MH, et al. (2012) A novel miR-155/miR-143 cascade controls glycolysis by regulating hexokinase 2 in breast cancer cells. EMBO J 31: 1985–98.

Jiao Y, Wickett NJ, Ayyampalayam S, Chanderbali AS, Landherr L, et al. (2011) Ancestral polyploidy in seed plants and angiosperms. Nature 473: 97-100.

Jiménez JA, Hughes KA, Alaks G, Graham L, Lacy RC (1994) An experimental study of inbreeding depression in a natural habitat. Science 266: 271–3.

Jimenez-Chillaron JC, Isganaitis E, Charalambous M, Gesta S, Pentinat-Pelegrin T, et al. (2009) Intergenerational transmission of glucose intolerance and obesity by in utero undernutrition in mice. Diabetes 58: 460–8.

Jin DY, Chae HZ, Rhee SG, Jeang KT (1997) Regulatory role for a novel human thioredoxin peroxidase in NF-κB activation. J Biol Chem 272: 30952–61.

Jin S, White E (2007) Role of autophagy in cancer: management of metabolic stress. Autophagy 3: 28-31.

Jin Z, Kirilly D, Weng C, Kawase E, Song X, et al. (2008) Differentiation-defective stem cells outcompete normal stem cells for niche occupancy in the Drosophila ovary. Cell Stem Cell 2: 39-49.

Jinks-Robertson S, Petes TD (1985) High-frequency meiotic gene conversion between repeated genes on nonhomologous chromosomes in yeast. Proc Natl Acad Sci USA 82: 3350–4.

Jiricny J (1994) Colon cancer and DNA repair: have mismatches met their match? Trends Genet 10: 164-8.

Jiricny J (1998) Replication errors: cha(lle)nging the genome. EMBO J 17: 6427-36.

Jirtle RL, Skinner MK (2007) Environmental epigenomics and disease susceptibility. Nat Rev Genet 8: 253–62.

Joanisse DR, Michaud S, Inaguma Y, Tanguay RM (1998) Small heat shock proteins of Drosophila: developmental expression and functions. J Biosci 23: 369–76.

Joffe M (2007) What harms the developing male reproductive system? In: Anderson D, Brinkworth MH, eds. Male-mediated Developmental Toxicology. Cambridge, UK: RSC Publishing. pp 28-50.

Joffe M (2010) What has happened to human fertility? Hum Reprod 25: 295-307.

Johansen HP, Cross AJ (1980) Effects of sexual maturation and sex steroid hormone treatment on the temperature preference of the guppy, Poecilia reticulata (Peters). Can J Zool 58: 586-8.

Johansson J (2008) Evolutionary responses to environmental changes: how does competition affect adaptation? Evolution 62: 421-35.

John B, Miklos GL (1988) The Eukaryotic Genome in Development and Evolution. London, UK: Allen & Unwin.

John JL (1997) The Hamilton-Zuk theory and initial test: an examination of some parasitological criticisms. Int J Parasitol 27: 1269-88.

Johnsborg O, Eldholm V, Havarstein LS (2007) Natural genetic transformation: prevalence, mechanisms and function. Res Microbiol 158: 767–78.

Johnsborg O, Eldholm V, Bjornstad ML, Havarstein LS (2008) A predatory mechanism dramatically increases the efficiency of lateral gene transfer in Streptococcus pneumoniae and related commensal species. Mol Microbiol 69: 245-53.

Johnsborg O, Håvarstein LS (2009) Regulation of natural genetic transformation and acquisition of transforming DNA in Streptococcus pneumoniae. FEMS Microbiol Rev 33: 627–42.

Johnsen A, Andersen V, Sunding C, Lifjeld JT (2000) Female bluethroats enhance offspring immunocompetence through extra-pair copulation. Nature 406: 296–9.

Johnson A, O’Donnell M (2005) Cellular DNA replicases: components and dynamics at the replication fork. Annu Rev Biochem 74: 283–315.

Johnson AB, Barton MC (2007) Hypoxia-induced and stress-specific changes in chromatin structure and function. Mutat Res 618: 149-62.

Johnson AB, Denko N, Barton MC (2008) Hypoxia induces a novel signature of chromatin modifications and global repression of transcription. Mutat Res 640: 174-9.

Johnson AD, Bachvarova RF, Drum M, Masi T (2001) Expression of axolotl DAZL RNA, a marker of germ plasm: widespread maternal RNA and onset of expression in germ cells approaching the gonad. Dev Biol 234: 402-15.

Johnson AD, Drum M, Bachvarova RF, Masi T, White ME, Crother BI (2003) Evolution of predetermined germ cells in vertebrate embryos: implications for macroevolution. Evol Dev 5: 414-31.

Johnson AD, Richardson E, Bachvarova RF, Crother BI (2011) Evolution of the germ line-soma relationship in vertebrate embryos. Reproduction 141: 291-300.

Johnson DW, Christie MR, Moye J (2010) Quantifying evolutionary potential of marine fish larvae: heritability, selection, and evolutionary constraints. Evolution 64: 2614-28.

Johnson JA, Toepfer JE, Dunn PO (2003) Contrasting patterns of mitochondrial and microsatellite population structure in fragmented populations of greater prairie-chickens. Mol Ecol 12: 3335-47.

Johnson KP, Seger J (2001) Elevated rates of nonsynonymous substitution in island birds. Mol Biol Evol 18: 874–81.

Johnson L (1986) Review article: spermatogenesis and aging in the human. J Androl 7: 331-54.

Johnson L (1991) Spermatogenesis. In: Cupps PT, ed. Reproduction in Domestic Animals, fourth edn. San Diego, CA: Academic Press Inc. pp 173–219.

Johnson L (1995) Efficiency of spermatogenesis. Microsc Res Tech 32: 385-422.

Johnson L, Petty CS, Neaves WB (1980) A comparative study of daily sperm production and testicular composition in humans and rats. Biol Reprod 22: 1233-43.

Johnson L, Petty CS, Neaves WB (1983a) Further quantification of human spermatogenesis: germ cell loss during postprophase of meiosis and its relationship to daily sperm production. Biol Reprod 29: 207-15.

Johnson L, Thompson DL Jr (1983b) Age-related and seasonal variation in the Sertoli cell population, daily sperm production and serum concentrations of follicle-stimulating hormone, luteinizing hormone and testosterone in stallions. Biol Reprod 29: 777–89.

Johnson L, Petty CS, Porter JC, Neaves WB (1984a) Influence of age on sperm production and testicular weights in men. J Reprod Fertil 70: 211–8.

Johnson L, Zane RS, Petty CS, Neaves WB (1984b) Quantification of the human Sertoli cell population: its distribution, relation to germ cell numbers, and age-related decline. Biol Reprod 31: 785–95.

Johnson L, Nguyen HB (1986) Annual cycle of the Sertoli cell population in adult stallions. J Reprod Fertil 76: 311–6.

Johnson L, Varner DD, Tatum ME, Scrutchfield WL (1991) Season but not age affects Sertoli cell number in stallions. Biol Reprod 45: 404–10.

Johnson L, Chaturvedi PK, Williams JD (1992) Missing generations of spermatocytes and spermatids contribute to the low efficiency of spermatogenesis in humans. Biol Reprod 47: 1091–8.

Johnson L, Varner DD, Roberts ME, Smith TL, Keillor GE, Scrutchfield WL (2000) Efficiency of spermatogenesis: a comparative approach. Anim Reprod Sci 60-61: 471-80.

Johnson L, Thompson DL Jr, Varner DD (2008) Role of Sertoli cell number and function on regulation of spermatogenesis. Anim Reprod Sci 105: 23-51.

Johnson LJ, Tricker PJ (2010) Epigenomic plasticity within populations: its evolutionary significance and potential. Heredity 105: 113–21.

Johnson MTJ, Lajeunesse MJ, Agrawal AA (2006) Additive and interactive effects of plant genotypic diversity on arthropod communities and plant fitness. Ecol Lett 9: 24-34.

Johnson MT, Smith SD, Rausher MD (2009) Plant sex and the evolution of plant defenses against herbivores. Proc Natl Acad Sci USA 106: 18079-84.

Johnson MT, Fitzjohn RG, Smith SD, Rausher MD, Otto SP (2011) Loss of sexual recombination and segregation is associated with increased diversification in evening primroses. Evolution 65: 3230-40.

Johnson RE,Washington MT, Prakash S, Prakash L (2000) Fidelity of human DNA polymerase et al. J Biol Chem 275: 7447–50.

Johnson RD, Jasin M (2001) Double-strand-break-induced homologous recombination in mammalian cells. Biochem Soc Trans 29: 196-201.

Johnson SG, Hopkins R, Goddard K (1999) Constraints on elevated ploidy in hybrid and nonhybrid parthenogenetic snails. J Hered 90: 659–62.

Johnson T (1999) Beneficial mutations, hitchhiking and the evolution of mutation rates in sexual populations. Genetics 151: 1621–31.

Johnson T, Barton N (2005) Theoretical models of selection and mutation on quantitative traits. Philos Trans R Soc Lond B 360: 1411–25.

Johnson TM, Yu ZX, Ferrans VJ, Lowenstein RA, Finkel T (1996) Reactive oxygen species are downstream mediators of p53-dependent apoptosis. Proc Natl Acad Sci USA 93: 11848–52.

Johnston H, Baker PJ, Abel M, Charlton HM, Jackson G, et al. (2004) Regulation of Sertoli cell number and activity by follicle-stimulating hormone and androgen during postnatal development in the mouse. Endocrinology 145: 318-29.

Johnston LA (2009) Competitive interactions between cells: death, growth and geography. Science 324: 1679–82.

Johnston M (1999) Feasting, fasting and fermenting. Glucose sensing in yeast and other cells. Trends Genet 15: 29–33.

Johnston TD (1982) Selective costs and benefits in the evolution of learning. Adv Stud Behav 12: 65–106.

Johnstone CP, Reina RD, Lill A (2012) Interpreting indices of physiological stress in free-living vertebrates. J Comp Physiol B182: 861-79.

Johnstone O, Lasko P (2001) Translational regulation and RNA localization in Drosophila oocytes and embryos. Annu Rev Genet 5: 365-406.

Jokela J, Lively CM (1995) Spatial variation for infection by digenetic trematodes in a population of freshwater snails (Potamopyrgus antipodarum). Oecologia 103: 509–17.

Jokela J, Lively CM, Dybdahl MF, Fox JA (1997) Evidence for a cost of sex in the freshwater snail Potamopyrgus antipodarum. Ecology 78: 452-60.

Jokela J, Schmid-Hempel P, Rigby MC (2000) Dr. Pangloss restrained by the Red Queen – steps towards a unified defence theory. Oikos 89: 267–74.

Jokela J, Lively CM, Dybdahl MF, Fox JA (2003) Genetic variation in sexual and clonal lineages of a freshwater snail. Biol J Linn Soc 79: 165–81.

Jokela J, Dybdahl MF, Lively CM (2009) The maintenance of sex, clonal dynamics, and host–parasite coevolution in a mixed population of sexual and asexual snails. Am Nat 174 Suppl 1: S43–S53.

Jolley KA, Wilson DJ, Kriz P, McVean G, Maiden MCJ (2005) The influence of mutation, recombination, population history, and selection on patterns of genetic diversity in Neisseria meningitidis. Mol Biol Evol 22: 562-9.

Jolly SE, Blackshaw AW (1988) Testicular migration, spermatogenesis, temperature regulation and environment of the sheath-tail bat, Taphozous georgianus. J Reprod Fertil 84: 447–55.

Joly-Tonetti N, Lamartine J (2012) The role of microRNAs in the cellular response to ionizing radiations. In: Nenoi M, ed. Current Topics in Ionizing Radiation Research. Rijeka, Croatia: InTech. pp 149-174.

Jones AM (2001) Programmed cell death in development and defense. Plant Physiol 125: 94–7.

Jones AG, Arnold SJ, Bürger R (2007) The mutation matrix and the evolution of evolvability. Evolution 61: 727–45.

Jones B (2012) Chromosome biology: mixing it up. Nat Rev Mol Cell Biol 13: 750.

Jones DT, Taylor WR, Thornton JM (1992) The rapid generation of mutation data matrices from protein sequences. Comput Appl Biosci 8: 275–82.

Jones EI, Bronstein JL, Ferrière R (2012) The fundamental role of competition in the ecology and evolution of mutualisms. Ann NY Acad Sci 1256: 66-88.

Jones KL, Smith DW, Harvey MA, Hall BD, Quan L (1975) Older paternal age and fresh gene mutation: data on additional disorders. J Pediatr 86: 84–8.

Jones KN (1994) Nonrandom mating in Clarkia gracilis (Onagraceae): a case of cryptic self-incompatibility. Am J Bot 81: 195–8.

Jones KT (2008) Meiosis in oocytes: Predisposition to aneuploidy and its increased incidence with age. Hum Reprod Update 14: 143–58.

Jones L, Hamilton AJ, Voinnet O, Thomas CL, Maule AJ, Baulcombe DC (1999) RNA–DNA interactions and DNA methylation in post-transcriptional gene silencing. Plant Cell 11: 2291–302.

Jones LP, Frankham R, Barker JS (1968) The effects of population size and selection intensity in selection for a quantitative character in Drosophila. II. Long-term response to selection. Genet Res 12: 249-66.

Jones ME, Cockburn A, Hamede R, Hawkins C, Hesterman H, et al. (2008) Life-history change in disease-ravaged Tasmanian devil populations. Proc Natl Acad Sci USA 105: 10023-7.

Jones OR, Crawley MJ, Pilkington JG, Pemberton JM (2005) Predictors of early survival in Soay sheep: cohort-, maternal- and individual-level variation. Proc R Soc Lond B 272: 2619-25.

Jones PA, Rideout WM 3rd, Shen JC, Spruck CH, Tsai YC (1992) Methylation, mutation and cancer. BioEssays 14: 33-6.

Jones PL, Wolffe AP (1999) Relationships between chromatin organization and DNA methylation in determining gene expression. Semin Cancer Biol 9: 339–47.

Jones R, Mann T, Sherins RJ (1979) Peroxidative breakdown of phospholipids in human spermatozoa: spermicidal effects of fatty acid peroxides and protective action of seminal plasma. Fertil Steril 31: 531–7.

Jones RG, Plas DR, Kubek S, Buzzai M, Mu J, et al. (2005) AMP-activated protein kinase induces a p53-dependent metabolic checkpoint. Mol Cell 18: 283–93.

Jonsson KI, Rabbow E, Schill RO, Harms-Ringdahl M, Rettberg P (2008) Tardigrades survive exposure to space in low Earth orbit. Curr Biol 18: 729–31.

Jopling CL, Schütz S, Sarnow P (2008) Position-dependent function for a tandem microRNA miR-122-binding site located in the hepatitis C virus RNA genome. Cell Host Microbe 4: 77-85.

Jordan IK, Rogozin IB, Wolf YI, Koonin EV (2002) Essential genes are more evolutionarily conserved than are nonessential genes in bacteria. Genome Res 12: 962-8.

Jørgensen KT, Sørensen JG, Bundgaard J (2006) Heat tolerance and the effect of mild heat stress on reproductive characters in Drosophila buzzatii males. J Therm Biol 31: 280–6.

Jørgensen MM, Bross P, Gregersen N (2003) Protein quality control in the endoplasmic reticulum. APMIS Suppl 2003: 86-91.

Jorgensen RA (2004) Restructuring the genome in response to adaptive challenge: McClintock’s bold conjecture revisited. Cold Spring Harb Symp Quant Biol 69: 349-54.

Jose AM, Hunter CP (2007) Transport of sequence-specific RNA interference information between cells. Annu Rev Genet 41: 305–30.

Jose AM, Garcia GA, Hunter CP (2011) Two classes of silencing RNAs move between Caenorhabditis elegans tissues. Nat Struct Mol Biol 18: 1184–8.

Jose C, Dufresne F (2010) Differential survival among genotypes of Daphnia pulex differing in reproductive mode, ploidy level, and geographic origin. Evol Ecol 24: 413–21.

José HJ, Berenice SGA, Cecilia VR (1997) Induction of antioxidant enzymes by dexamethasone in the adult rat lung. Life Sci 60: 2059-67.

Joseph SB, Kirkpatrick M (2004) Haploid selection in animals. Trends Ecol Evol 19: 592–7.

Joshi A, Moody ME (1998) The cost of sex revisited: effects of male gamete output of hermaphrodites that are asexual in their female capacity. J Theor Biol 195: 533-42.

Joshi A, Prasad NG, Shakarad M (2001) K-selection, α-selection, effectiveness, and tolerance in competition: density-dependent selection revisited. J Genet 80: 63–75.

Jovanovic SV, Clements D, MacLeod K (1998) Biomarkers of oxidative stress are significantly elevated in Down syndrome. Free Radic Biol Med 25: 1044-8.

Joyner Matos J (2007) Magnitude of the oxidative stress response influences species distributions. Thesis. Gainesville, FL: University of Florida.

Joyner-Matos J, Chapman LJ, Downs CA, Hofer T, Leeuwenburgh C, Julian D (2007) Stress response of a freshwater clam along an abiotic gradient: too much oxygen may limit distribution. Funct Ecol 21: 344–55.

Juan ME, González-Pons E, Munuera T, Ballester J, Rodríguez-Gil JE, Planas JM (2005) trans-Resveratrol, a natural antioxidant from grapes, increases sperm output in healthy rats. J Nutr 135: 757-60.

Judson HF (1979) The Eighth Day of Creation. New York, NY: Simon and Schuster.

Judson OP (1997) A model of asexuality and clonal diversity: cloning the Red Queen. J Theor Biol 186: 33-40.

Judson OP, Normark BB (1996) Ancient asexual scandals. Trends Ecol Evol 11: 41–6.

Juedes MJ, Wogan GN (1996) Peroxynitrite-induced mutation spectra of pSP189 following replication in bacteria and in human cells. Mutat Res 349: 51-61.

Juliano CE, Swartz SZ, Wessel GM (2010) A conserved germline multipotency program. Development 137: 4113–26.

Jullien PE, Berger F (2010) DNA methylation reprogramming during plant sexual reproduction? Trends Genet 26: 394-9.

Jullien PE, Susaki D, Yelagandula R, Higashiyama T, Berger F (2012) DNA methylation dynamics during sexual reproduction in Arabidopsis thaliana. Curr Biol 22: 1825–30.

Jung Y, Isaacs JS, Lee S, Trepel J, Liu ZG, Neckers L (2003) Hypoxia-inducible factor induction by tumour necrosis factor in normoxic cells requires receptor-interacting protein-dependent nuclear factor κB activation. Biochem J 370: 1011–7.

Jungbluth A, Busam K, Kolb D, Iversen K, Coplan K, et al. (2000a) Expression of MAGE-antigens in normal tissues and cancer. Int J Cancer 85: 460-5.

Jungbluth AA, Stockert E, Chen YT, Kolb D, Iversen K, et al. (2000b) Monoclonal antibody MA454 reveals a heterogeneous expression pattern of MAGE-1 antigen in formalin-fixed paraffin embedded lung tumours. Br J Cancer 83: 493-7.

Jungbluth AA, Chen YT, Stockert E, Busam KJ, Kolb D, et al. (2001a) Immunohistochemical analysis of NY-ESO-1 antigen expression in normal and malignant human tissues. Int J Cancer 92: 856-60.

Jungbluth A, Busam K, Iversen K, Kolb D, Coplan K, et al. (2001b) Cancer-Testis (CT) antigens MAGE-1, MAGE-3, NY-ESO-1, and CT7 are expressed in female germ cells. Mod Pathol 14: 211A.

Jungbluth AA, Chen YT, Busam KJ, Coplan K, Kolb D, et al. (2002) CT7 (MAGE-C1) antigen expression in normal and neoplastic tissues. Int J Cancer 99: 839-45.

Junttila MR, Karnezis AN, Garcia D, Madriles F, Kortlever RM, et al. (2010) Selective activation of p53-mediated tumour suppression in high-grade tumours. Nature 468: 567-71.

Jutte NH, Jansen R, Grootegoed JA, Rommerts FF, Clausen OP, van der Molen HJ (1982) Regulation of survival of rat pachytene spermatocytes by lactate supply from Sertoli cells. J Reprod Fertil 65: 431–8.

Jutte NH, Jansen R, Grootegoed JA, RommertsFF, van der Molen HJ (1983) FSH stimulation of the production of pyruvate and lactate by rat Sertoli cells may be involved in hormonal regulation of spermatogenesis. J Reprod Fertil 68: 219–26.

Kaati G, Bygren LO, Edvinsson S (2002) Cardiovascular and diabetes mortality determined by nutrition during parents’ and grandparents’ slow growth period. Eur J Hum Genet 10: 682–8.

Kaati G, Bygren LO, Pembrey M, Sjostrom M (2007) Transgenerational response to nutrition, early life circumstances and longevity. Eur J Hum Genet 15: 784–90.

Kadonaga JT, Carner KR, Masiarz FR, Tjian R (1987) Isolation of cDNA encoding transcription factor Sp1 and functional analysis of the DNA binding domain. Cell 51: 1079–90.

Kadyk LC, Hartwell LH (1992) Sister chromatids are preferred over homologs as substrates for recombinational repair in Saccharomyces cerevisiae. Genetics 132: 387–402.

Kaelin WG Jr, Ratcliffe PJ (2008) Oxygen sensing by metazoans: the central role of the HIF hydroxylase pathway. Mol Cell 30: 393–402.

Kærn M, Elston TC, Blake WJ, Collins JJ (2005) Stochasticity in gene expression: from theories to phenotypes. Nat Rev Genet 6: 451–64.

Kafri T, Ariel M, Brandeis M, Shemer R, Urven L, et al. (1992) Developmental pattern of gene-specific DNA methylation in the mouse embryo and germ line. Genes Dev 6: 705–14.

Kagan VE, Fabisiak JP, Shvedova AA, Tyurina YY, Tyurin VA, et al. (2000) Oxidative signaling pathway for externalization of plasma membrane phosphatidylserine during apoptosis. FEBS Lett 477: 1-7.

Kagan VE, Gleiss B, Tyurina YY, Tyurin VA, Elenström-Magnusson C, et al. (2002) A role for oxidative stress in apoptosis: oxidation and externalization of phosphatidylserine is required for macrophage clearance of cells undergoing Fas-mediated apoptosis. J Immunol 169: 487-99.

Kageyama Y, Ishibashi K, Hayashi T, Xia G, Sasaki S, Kihara K (2001) Expression of aquaporins 7 and 8 in the developing rat testis. Andrologia 33: 165-9.

Kahn NW, Quinn TW (1999) Male-driven evolution among Eoaves? A test of the replicative division hypothesis in a heterogametic female (ZW) system. J Mol Evol 49: 750-9.

Kaipia A, Hsueh AJ (1997) Regulation of ovarian follicle atresia. Annu Rev Physiol 59: 349–63.

Kaiser D (1996) Bacteria also vote. Science 272: 1598-9.

Kajiwara C, Kondo S, Uda S, Dai L, Ichiyanagi T, et al. (2012) Spermatogenesis arrest caused by conditional deletion of Hsp90α in adult mice. Biol Open 1: 977-82.

Kalejs M, Erenpreisa J (2005) Cancer/testis antigens and gametogenesis: a review and "brain-storming" session. Cancer Cell Int 5: 4.

Kalia S, Bansal MP (2009) Regulation of apoptosis by Caspases under oxidative stress conditions in mice testicular cells: in vitro molecular mechanism. Mol Cell Biochem 322: 43-52.

Kalidas S, Sanders C, Ye X, Strauss T, Kuhn M, et al. (2008) Drosophila R2D2 mediates follicle formation in somatic tissues through interactions with Dicer-1. Mech Dev 125: 475–85.

Kallio M, Chang Y, Manuel M, Alastalo TP, Rallu M, et al. (2002) Brain abnormalities, defective meiotic chromosome synapsis and female subfertility in HSF2 null mice. EMBO J 21: 2591–601.

Kalliomäki TM, McCallum G, Lunt SJ, Wells PG, Hill RP (2008) Analysis of the effects of exposure to acute hypoxia on oxidative lesions and tumour progression in a transgenic mouse breast cancer model. BMC Cancer 8: 151.

Kalmar B, Greensmith L (2009) Induction of heat shock proteins for protection against oxidative stress. Adv Drug Deliv Rev 61: 310-8.

Kaltschmidt B, Kaltschmidt C, Hofmann TG, Hehner SP, Dröge W, Schmitz ML (2000) The pro- or anti-apoptotic function of NF-kappaB is determined by the nature of the apoptotic stimulus. Eur J Biochem 267: 3828-35.

Kaltz O, Bell G (2002) The ecology and genetics of fitness in Chlamydomonas. XII. Repeated sexual episodes increase rates of adaptation to novel environments. Evolution 56: 1743–53.

Kaluz S, Kaluzová M, Stanbridge EJ (2008) Regulation of gene expression by hypoxia: integration of the HIF-transduced hypoxic signal at the hypoxia-responsive element. Clin Chim Acta 395: 6-13.

Kamaruddin M, Kroetsch T, Basrur PK, Hansen PJ, King WA (2004) Immunolocalization of heat shock protein 70 in bovine spermatozoa. Andrologia 36: 327–34.

Kambhampati S, Rai KS, Verleye DM (1992) Frequencies of mitochondrial DNA haplotypes in laboratory cage populations of the mosquito, Aedes albopictus. Genetics 132: 205-9.

Kamer G, Argos P (1984) Primary structural comparison of RNA-dependent polymerases from plant, animal and bacterial viruses. Nucleic Acids Res 12: 7269–82.

Kaminker J, Bergman C, Kronmiller B, Carlson J, Svirskas R, et al. (2002) The transposable elements of the Drosophila melanogaster euchromatin: A genomics perspective. Genome Biol 3: RESEARCH0084.

Kamminga LM, Luteijn MJ, den Broeder MJ, Redl S, Kaaij LJ, et al. (2010) Hen1 is required for oocyte development and piRNA stability in zebrafish. EMBO J 29: 3688-700.

Kamminga LM, van Wolfswinkel JC, Luteijn MJ, Kaaij LJT, Bagijn MP, et al. (2012) Differential impact of the HEN1 homolog HENN-1 on 21U and 26G RNAs in the germline of Caenorhabditis elegans. PLoS Genet 8: e1002702.

Kanai Y (2010) Genome-wide DNA methylation profiles in precancerous conditions and cancers. Cancer Sci 101: 36-45.

Kanatsu-Shinohara M, Inoue K, Lee J, Yoshimoto M, Ogonuki N, Miki H, et al. (2004) Generation of pluripotent stem cells from neonatal mouse testis. Cell 119: 1001-12.

Kanazawa A, Liu B, Kong F, Arase S, Abe J (2009) Adaptive evolution involving gene duplication and insertion of a novel Ty1/copia-like retrotransposon in soybean. J Mol Evol 69: 164–75.

Kanazawa H, Kurihara N, Hirata K, Terakawa K, Fujiwara H, et al. (1992) The effect of thyroid hormone on generation of free radicals by neutrophils and alveolar macrophages. Jpn J Allergol 41: 135–9.

Kaneko G, Yoshinaga T, Yanagawa Y, Ozaki Y, Tsukamoto K, Watabe S (2011) Calorie restriction-induced maternal longevity is transmitted to their daughters in a rotifer. Funct Ecol 25: 209-16.

Kaneko K (2007) Evolution of robustness to noise and mutation in gene expression dynamics. PLoS ONE 2: e434.

Kaneko K (2009) Relationship among phenotypic plasticity, phenotypic fluctuations, robustness, and evolvability; Waddington’s legacy revisited under the spirit of Einstein. J Biosci 34: 529–42.

Kaneko K, Furusawa C (2006) An evolutionary relationship between genetic variation and phenotypic fluctuation. J Theor Biol 240: 78-86.

Kaneko Y, Nishiyama H, Nonoguchi K, Higashitsuji H, Kishishita M, Fujita J (1997a) A novel hsp110-related gene, apg-1, that is abundantly expressed in the testis responds to a low temperature heat shock rather than the traditional elevated temperatures. J Biol Chem 272: 2640–5.

Kaneko Y, Kimura T, Nishiyama H, Noda Y, Fujita J (1997b) Developmentally regulated expression of APG-1, a member of heat shock protein 110 family in murine male germ cells. Biochem Biophys Res Commun 233: 113–6.

Kang HC, Yoon SH, Lee CM, Roh KH (2013) DNA repair of eukaryotes associated with non-coding small RNAs. J Appl Biol Chem 56: 37-42.

Kang MY, Kim HB, Piao C, Lee KH, Hyun JW, et al. (2013) The critical role of catalase in prooxidant and antioxidant function of p53. Cell Death Differ 20: 117-29.

Kang SW, Thayananuphat A, Bakken T, El Halawani ME (2007) Dopamine-melatonin neurons in the avian hypothalamus controlling seasonal reproduction. Neuroscience 150: 223–33.

Kanki T, Wang K, Cao Y, Baba M, Klionsky DJ (2009) Atg32 is a mitochondrial protein that confers selectivity during mitophagy. Dev Cell 17: 98-109.

Kannan K, Jain SK (2000) Oxidative stress and apoptosis. Pathophysiology 7: 153–163.

Kanno T, Mette MF, Kreil DP, Aufsatz W, Matzke M, Matzke AJ (2004) Involvement of putative SNF2 chromatin remodeling protein DRD1 in RNA-directed DNA methylation. Curr Biol 14: 801–5.

Kant GJ, Lenox RH, Bunnell BN, Mougey EH, Pennington LL, Meyerhoff JL (1983) Comparison of stress response in male and female rats: pituitary cyclic AMP and plasma prolactin, growth hormone and corticosterone. Psychoneuroendocrinology 8: 421-8.

Kao KC, Sherlock G (2008) Molecular characterization of clonal interference during adaptive evolution in asexual populations of Saccharomyces cerevisiae. Nat Genet 40: 1499-504.

Kao S, Chao HT, Wei YH (1995) Mitochondrial deoxyribonucleic acid 4977-bp deletion is associated with diminished fertility and motility of human sperm. Biol Reprod 52: 729–36.

Kapitonov VV, Jurka J (2005) RAG1 core and V(D)J recombination signal sequences were derived from Transib transposons. PLoS Biol 3: 998–1011.

Kapoor A, Agius F, Zhu JK (2005) Preventing transcriptional gene silencing by active DNA demethylation. FEBS Lett 579: 5889-98.

Kapranov P, Willingham AT, Gingeras TR (2007) Genome-wide transcription and the implications for genomic organization. Nat Rev Genet 8: 413-23.

Kara H, Cevik A, Konar V, Dayangac A, Yilmaz M (2007) Protective effects of antioxidants against cadmium-induced oxidative damage in rat testes. Biol Trace Elem Res 120: 205-11.

Karahalil B, Hogue BA, de Souza-Pinto NC, Bohr VA (2002) Base excision repair capacity in mitochondria and nuclei: tissue-specific variations. FASEB J 16: 1895–902.

Karanjawala ZE, Murphy N, Hinton DR, Hsieh CL, Lieber MR (2002) Oxygen metabolism causes chromosome breaks and is associated with the neuronal apoptosis observed in DNA double-strand break repair mutants. Curr Biol 12: 397-402.

Karbownik M, Lewinski A (2003) The role of oxidative stress in physiological and pathological processes in the thyroid gland; possible involvement in pineal-thyroid interactions. Neuroendocrinol Lett 24: 293–303.

Kärkkäinen K, Savolainen O, Koski V (1999) Why do plants abort so many developing seeds: bad offspring or bad maternal genotypes? Evol Ecol 13: 305–17.

Karlberg M, Ekoff M, Labi V, Strasser A, Huang D, Nilsson G (2010) Pro-apoptotic Bax is the major and Bak an auxiliary effector in cytokine deprivation-induced mast cell apoptosis. Cell Death Dis 1: e43.

Karlin S, McGregor J (1974) Towards a theory of the evolution of modifier genes. Theor Popul Biol 5: 59–103

Karlsson A, Deb-Basu D, Cherry A, Turner S, Ford J, Felsher DW (2003) Defective double-strand DNA break repair and chromosomal translocations by MYC overexpression. Proc Natl Acad Sci USA 100: 9974–9.

Karp X, Ambros V (2005) Developmental biology. Encountering microRNAs in cell fate signaling. Science 310: 1288–9.

Kartal Ö, Ebenhöh O (2009) Ground state robustness as an evolutionary design principle in signaling networks. PLoS ONE 4: e8001.

Kashi YD, King D, Soller M (1997) Simple sequence repeats as a source of quantitative genetic variation. Trends Genet 13: 74–8.

Kashi Y, King DG (2006) Simple sequence repeats as advantageous mutators in evolution. Trends Genet 22: 253-9.

Kashkush K, Feldman M, Levy AA (2002) Gene loss, silencing and activation in a newly synthesized wheat allotetraploid. Genetics 160: 1651–9.

Kashtan N, Alon U (2005) Spontaneous evolution of modularity and network motifs. Proc Natl Acad Sci USA 102: 13773–8.

Kashtan N, Noor E, Alon U (2007) Varying environments can speed up evolution. Proc Natl Acad Sci USA 104: 13711–6.

Kassen R (2002) The experimental evolution of specialists, generalists, and the maintenance of diversity. J Evol Biol 15: 173–90.

Kassir Y, Granot D, Simchen G (1988) IME1, a positive regulator gene of meiosis in S. cerevisiae. Cell 52: 853–62.

Kato H, Fu Z, Kotera N, Sugahara K, Kubo T (1999) Regulation of the expression of serotonin N-acetyltransferase gene in Japanese quail (Coturnix japonica): I. Rhythmic pattern and effect of light. J Pineal Res 27: 24–33.

Kato H, Tamamizu-Kato S, Shibasaki F (2004) Histone deacetylase 7 associates with hypoxia-inducible factor 1alpha and increases transcriptional activity. J Biol Chem 279: 41966-74.

Kato MV, Ishizaki K, Shimizu T, Ejima Y, Tanooka H, et al. (1994) Parental origin of germ-line and somatic mutations in the retinoblastoma gene. Hum Genet 94:31-8.

Kato T, Watanabe M, Ohta T (1994) Induction of the SOS response and mutations by reactive oxygen-regenerating compounds in various Escherichia coli mutants defective in the mutM, mutY or soxRS loci. Mutagenesis 9: 245-51.

Kato T, Okazaki K, Murakami H, Stettler S, Fantes PA, Okayama H (1996) Stress signal, mediated by a Hog1-like MAP kinase, controls sexual development in fission yeast. FEBS Lett 378: 207-12.

Kato T, Yamada K, Inagaki H, Kogo H, Ohye T, et al. (2007) Age has no effect on de novo constitutional t(11;22) translocation frequency in sperm. Fertil Steril 88: 1446–8.

Katz DJ, Edwards TM, Reinke V, Kelly WG (2009) A C. elegans LSD1 demethylase contributes to germline immortality by reprogramming epigenetic memory. Cell 137: 308–20.

Katz JE, Dlakic M, Clarke S (2003) Automated identification of putative methyltransferases from genomic open reading frames. Mol Cell Proteomics 2: 525–40.

Kaur A, Van PT, Busch CR, Robinson CK, Pan M, et al. (2010) Coordination of frontline defense mechanisms under severe oxidative stress. Mol Syst Biol 6: 393.

Kawabe Y, Alvarez-Curto E, Ritchie AV, Schaap P (2009) The evolution of morphogenetic signalling in social amoebae. In: Pontarotti P, ed. Evolutionary biology. Concept, modeling and application. Berlin: Springer-Verlag. pp 91-108.

Kawaji H, Hayashizaki Y (2008) Exploration of small RNAs. PLoS Genet 4: 3–8.

Kawamura K, Fujiwara S (2000) Advantage or disadvantage: Is asexual reproduction beneficial to survival of the tunicate, Polyandrocarpa misakiensis? Zool Sci 17: 281-91.

Kawamura T, Suzuki J, Wang YV, Menendez S, Morera LB, et al. (2009) Linking the p53 tumour suppressor pathway to somatic cell reprogramming. Nature 460: 1140–4.

Kawano KM, Yamaguchi N, Kasuya E, Yahara T (2009) Extra-pair mate choice in the female great tit, Parus major: good males or compatible males. J Ethol 27: 349–59.

Kawasaki H, Taira K (2004) Induction of DNA methylation and gene silencing by short interfering RNAs in human cells. Nature 431: 211–7.

Kawasaki Y, Nakagawa A, Nagaosa K, Shiratsuchi A, Nakanishi Y (2002) Phosphatidylserine binding of class B scavenger receptor type I, a phagocytosis receptor of testicular sertoli cells. J Biol Chem 277: 27559-66.

Kawecki TJ (1988) Unisexual-bisexual breeding complexes in Poeciliidae: why do males copulate with unisexual females? Evolution 42: 1018–23.

Kawecki TJ (1997) Sympatric speciation via habitat specialization driven by deleterious mutations. Evolution 51: 1751–63.

Kawecki TJ (2000) The evolution of genetic canalization under fluctuating selection. Evolution 54: 1-12.

Kawecki TJ, Barton NH, Fry JD (1997) Mutational collapse of fitness in marginal habitats and the evolution of ecological specialisation. J Evol Biol 10: 407-29.

Kaya M, Herniou EA, Barraclough TG, Fontaneto D (2009) Inconsistent estimates of diversity between traditional and DNA taxonomy in bdelloid rotifers. Org Divers Evol 9: 3–12.

Kazakov SA, Astashkina TG, Mamaev SV, Vlassov VV (1988) Site-specific cleavage of single-stranded DNAs at unique sites by a copper-dependent redox reaction. Nature 335: 186-8.

Kazazian HH Jr (2004) Mobile elements: Drivers of genome evolution. Science 303: 1626-32.

Ke Q, Costa M (2006) Hypoxia-inducible factor-1 (HIF-1). Mol Pharmacol 70: 1469–80.

Kearney MR (2003) Why is sex so unpopular in the Australian desert? Trends Ecol Evol 18: 605-7.

Kearney M (2005) Hybridization, glaciation and geographical parthenogenesis. Trends Ecol Evol 20: 495–502.

Kearney M, Moussalli A, Strasburg J, Lindenmayer D, Moritz C (2003) Geographic parthenogenesis in the Australian arid zone: I. A climatic anaysis of Heteronotia binoei complex (Gekkonidae). Evol Ecol Res 5: 953–76.

Kearney M, Shine R (2005) Lower fecundity in parthenogenetic geckos than sexual relatives in the Australian arid zone. J Evol Biol 18: 609–18.

Kearney M, Wahl R, Autumn K (2005) Increased capacity for sustained locomotion at low temperature in parthenogenetic geckos of hybrid origin. Physiol Biochem Zool 78: 316–24.

Kearney MR, Blacket MJ, Strasburg JL, Moritz C (2006) Waves of parthenogenesis in the desert: evidence for the parallel loss of sex in a grasshopper and a gecko from Australia. Mol Ecol 15: 1743-8.

Kearns DB, Chu F, Rudner R, Losick R (2004) Genes governing swarming in Bacillus subtilis and evidence for a phase variation mechanism controlling surface motility. Mol Microbiol 52: 357-69.

Kearns M, Preis J, McDonald M, Morris C, Whitelaw E (2000) Complex patterns of inheritance of an imprinted murine transgene suggest incomplete germline erasure. Nucleic Acids Res 28: 3301–9.

Kedersha N, Anderson P (2002) Stress granules: sites of mRNA triage that regulate mRNA stability and translatability. Biochem Soc Trans 30: 963-9.

Kedersha N, Stoecklin G, Ayodele M, Yacono P, Lykke-Andersen J, et al. (2005) Stress granules and processing bodies are dynamically linked sites of mRNP remodeling. J Cell Biol 169: 871-84.

Kedersha N, Anderson P (2007) Mammalian stress granules and processing bodies. Methods Enzymol 431: 61–81.

Kedersha N, Tisdale S, Hickman T, Anderson P (2008) Real-time and quantitative imaging of mammalian stress granules and processing bodies. Methods Enzymol 448: 521–52.

Keeling PJ, Palmer JD (2008) Horizontal gene transfer in eukaryotic evolution. Nat Rev Genet 9: 605–18.

Keeney S (2007) Spo11 and the formation of DNA double-strand breaks in meiosis. In: Egel R, Lankenau D-H, eds. Recombination and Meiosis: Crossing-Over and Disjunction. Heidelberg, Germany: Springer-Verlag. pp 81-123.

Keeney S, Giroux CN, Kleckner N (1997) Meiosis-specific DNA double-strand breaks are catalyzed by Spo11, a member of a widely conserved protein family. Cell 88: 375–84.

Keeney S, Baudat F, Angeles M, Zhou ZH, Copeland NG, et al. (1999) A mouse homolog of the Saccharomyces cerevisiae meiotic recombination DNA transesterase Spo11p. Genomics 61: 170-82.

Kehler J, Tolkunova E, Koschorz B, Pesce M, Gentile L, et al. (2004) Oct4 is required for primordial germ cell survival. EMBO Rep 5: 1078–83.

Keightley PD (1996) Metabolic models of selection response. J Theor Biol 182: 311–6.

Keightley PD (2004) Mutational variation and long-term selection response. Plant Breeding Rev 24: 227–247.

Keightley PD, Caballero A (1997) Genomic mutation rates for lifetime reproductive output and lifespan in Caenorhabditis elegans. Proc Natl Acad Sci USA 94: 3823-7.

Keightley PD, Ohnishi O(1998) EMS-induced polygenic mutation rates for nine quantitative characters in Drosophila melanogaster. Genetics 148: 753-66.

Keightley PD, Eyre-Walker A (1999) Terumi Mukai and the riddle of deleterious mutation rates. Genetics 153: 515–23.

Keightley PD, Eyre-Walker A (2000) Deleterious mutations and the evolution of sex. Science 290: 331-3.

Keightley PD, Davies EK, Peters AD, Shaw RG (2000) Properties of EMS-induced mutations affecting life history traits in Caenorhabditis elegans and inferences about bivariate distributions of mutation effects. Genetics 156: 143-54.

Keightley PD, Lynch M (2003) Towards a realistic model of mutations affecting fitness. Evolution 57: 683–5.

Keightley PD, Otto SP (2006) Interference among deleterious mutations favours sex and recombination in finite populations. Nature 443: 89-92.

Keightley PD, Trivedi U, Thomson M, Oliver F, Kumar S, et al. (2009) Analysis of the genome sequences of three Drosophila melanogaster spontaneous mutation accumulation lines. Genome Res 19: 1195–201.

Keil RL, Roeder GS (1984) Cis-acting, recombination-stimulating activity in a fragment of the ribosomal DNA of S. cerevisiae. Cell 39: 377–86.

Kelkar YD, Tyekucheva S, Chiaromonte F, Makova KD (2008) The genome-wide determinants of human and chimpanzee microsatellite evolution. Genome Res 18: 30–8.

Keller L, ed. (1999) Levels of selection in evolution. Princeton, NJ: Princeton University Press.

Keller LF (1998) Inbreeding and its fitness effects in an insular population of sparrows (Melospiza melodia). Evolution 52: 240–50.

Keller LF, Arcese P, Smith JNM, Hochachka WM, Stearns SC (1994) Selection against inbred song sparrows during a natural population bottleneck. Nature 372: 356–7.

Keller LF, Waller DM (2002) Inbreeding effects in wild populations. Trends Ecol Evol 17: 230–41.

Kellermann VM, Hoffmann AA, Sgro CM (2007) Hsp90 inhibition and the expression of phenotypic variability in the rainforest species Drosophila birchii. Biol J Linn Soc 92: 457–65.

Kelley MR, Parsons SH (2001) Redox regulation of the DNA repair function of the human AP endonuclease Ape1/ref-1. Antioxid Redox Signal 3: 671–83.

Kelley MR, Georgiadis MM, Fishel ML (2012) APE1/Ref-1 role in redox signaling: translational applications of targeting the redox function of the DNA repair/redox protein APE1/Ref-1. Curr Mol Pharmacol 5: 36-53.

Kelly K, Cochran BH, Stiles CD, Leder P (1983) Cell-specific regulation of the c-myc gene by lymphocyte mitogens and platelet-derived growth factor. Cell 35: 603-610.

Kelner MJ, Bagnell R, Montoya M, Estes L, Uglick SF, Cerutti P (1995) Transfection with human copper-zinc superoxide dismutase induces bidirectional alterations in other antioxidant enzymes, proteins, growth factor response, and paraquat resistance. Free Radic Biol Med 18: 497-506.

Kelus A, Steinberg CM (1991) Is there a high rate of mitotic recombination between the loci encoding immunoglobulin VH and CH regions in gonial cells? Immunogenetics 33: 255-9.

Kemal Duru N, Morshedi M, Oehninger S (2000) Effects of hydrogen peroxide on DNA and plasma membrane integrity of human spermatozoa. Fertil Steril 74: 1200–7.

Kempenaers B (2007) Mate choice and genetic quality: A review of the heterozygosity theory. Adv Stud Behav 37: 189–278.

Kempenaers B, Verheyen GR, Dhont A (1997) Extra-pair paternity in the blue tit Parus caeruleus: Female choice, male characteristics, and offspring quality. Behav Ecol 8: 481–92.

Kenneth NS, Rocha S (2008) Regulation of gene expression by hypoxia. Biochem J 414: 19-29.

Kennison JA, Ripoll P (1981) Spontaneous mitotic recombination and evidence for an X-ray-inducible system for the repair of DNA damage in Drosophila melanogaster. Genetics 98: 91-103.

Kent WJ, Baertsch R, Hinrichs A, Miller W, Haussler D (2003) Evolution’s cauldron: duplication, deletion, and rearrangement in the mouse and human genomes. Proc Natl Acad Sci USA 100: 11484–9.

Kerfoot WC, Weider LJ (2004) Experimental paleoecology (resurrection ecology): chasing Van Valen’s Red Queen hypothesis. Limnol Oceanogr 49: 1300–16.

Kern JC, Kehrer JP (2005) Free radicals and apoptosis: relationships with glutathione, thioredoxin, and the BCL family of proteins. Front Biosci 10: 1727-38.

Kern S, Robertson SA, Mau VJ, Maddocks S (1995) Cytokine secretion by macrophages in the rat testis. Biol Reprod 53: 1407–16.

Kerr B, Neuhauser C, Bohannan BJ, Dean AM (2006) Local migration promoted competitive restraint in a host-pathogen "tragedy of the commons". Nature 442: 75-8.

Kerr JB (1992) Spontaneous degeneration of germ cells in normal rat testis: assessment of cell types and frequency during the spermatogenic cycle. J Reprod Fertil 95: 825–30.

Kerszberg M (2000) The survival of slow reproducers. J Theor Biol 206: 81–9.

Kervancioglu ME, Djahanbakhch O, Aitken RJ (1994) Epithelial cell coculture and the induction of sperm capacitation. Fertil Steril 61: 1103-8.

Keskintepe L, Brackett BG (1996) In vitro developmental competence of in vitro-matured bovine oocytes fertilized and cultured in completely defined media. Biol Reprod 55: 333-9.

Kessin RH (2010) Dictyostelium: Evolution, cell biology, and the development of multicellularity. Cambridge, UK: Cambridge University Press.

Kessler DA, Levine H, Ridgway D, Tsimring L (1997) Evolution on a smooth landscape. J Stat Phys 87: 519–44.

Keszenman DJ, Carmen Candreva E, Nunes E (2000) Cellular and molecular effects of bleomycin are modulated by heat shock in Saccharomyces cerevisiae. Mutat Res 459: 29-41.

Ketterer B, Meyer DJ (1989) Glutathione transferases: a possible role in the detoxication and repair of DNA and lipid hydroperoxides. Mutat Res 214: 33-40.

Keyes ME (1999) The prion challenge to the “Central Dogma” of molecular biology, 1965–1991. Part I: Prelude to prions. Stud Hist Philos Biol Sci 30: 1–19.

Keyse SM, Tyrrell RM (1989) Heme oxygenase is the major 32-kDa stress protein induced in human skin fibroblasts by UVA radiation, hydrogen peroxide, and sodium arsenite. Proc Natl Acad Sci USA 86: 99-103.

Keyte AL, Percifield R, Liu B, Wendel JF (2006) Intraspecific DNA methylation polymorphism in cotton (Gossypium hirsutum L.). J Hered 97: 444–50.

Kha DT, Wang G, Natrajan N, Harrison L, Vasquez KM (2010) Pathways for double-strand break repair in genetically unstable Z-DNA-forming sequences. J Mol Biol 398: 471–80.

Khachane AN, Harrison PM (2010) Mining mammalian transcript data for functional long non-coding RNAs. PLoS ONE 5: e10316.

Khakhlova O, Bock R (2006) Elimination of deleterious mutations in plastid genomes by gene conversion. Plant J 46: 85–94.

Khan AI, Dinh DM, Schneider D, Lenski RE, Cooper TF (2011) Negative epistasis between beneficial mutations in an evolving bacterial population. Science 332:1193-6.

Khan MA, Antonovics J, Bradshaw AD (1976) Adaptation to heterogeneous environments. III. The inheritance of response to spacing in flax and linseed (Linum usitatissimum). Aust J Agric Res 27: 649-59.

Khan SA, Söder O, Syed V, Gustafsson K, Lindh M, Ritzén EM (1987) The rat testis produces large amounts of an interleukin-1-like factor. Int J Androl 10: 495-503.

Khan SA, Khan SJ, Dorrington JH (1992) Interleukin-1 stimulates deoxyribonucleic acid synthesis in immature rat Leydig cells in vitro. Endocrinology 131: 1853-7.

Khanna KK, Jackson SP (2001) DNA double-strand breaks: signaling, repair and the cancer connection. Nat Genet 27: 247–54.

Khare A, Shaulsky G (2006) First among equals: competition between genetically identical cells. Nat Rev Genet 7: 577-83.

Khole V (2003) Epididymis as a target for contraception. Indian J Exp Biol 41: 764-72.

Kholodenko B (2006) Cell signalling dynamics in time and space. Nat Rev Mol Cell Biol 7: 165–76.

Khor TO, Huang MT, Kwon KH, Chan JY, Reddy BS, Kong AN (2006) Nrf2-deficient mice have an increased susceptibility to dextran sulfate sodium-induced colitis. Cancer Res 66: 11580–4.

Khurana JS, Theurkauf W (2010) piRNAs, transposon silencing, and Drosophila germline development. J Cell Biol 191: 905-13.

Khurana NK, Niemann H (2000) Effects of oocyte quality, oxygen tension, embryo density, cumulus cells and energy substrates on cleavage and morula/blastocyst formation of bovine embryos. Theriogenology 54: 742–56.

Kibota TT, Lynch M (1996) Estimate of the genomic mutation rate deleterious to overall fitness in E. coli. Nature 381: 694-6.

Kidwell MG (2005) Transposable elements. In: Gregory TR, ed. The Evolution of the Genome. San Diego, CA: Elsevier. pp 165–221.

Kidwell MG, Lisch DR (1997) Transposable elements as sources of variation in animals and plants. Proc Natl Acad Sci USA 94: 7704–11.

Kidwell MG, Lisch DR (2000) Transposable elements and host genome evolution. Trends Ecol Evol 15: 95-9.

Kidwell MG, Lisch DR (2001) Perspective: transposable elements, parasitic DNA, and genome evolution. Evolution 55: 1–24.

Kierszenbaum AL (2001) Apoptosis during spermatogenesis: the thrill of being alive. Mol Reprod Dev 58: 1–3.

Kierszenbaum AL, Tres LL (1978) RNA transcription and chromatin structure during meiotic and postmeiotic stages of spermatogenesis. Fed Proc 37: 2512-6.

Kietzmann T, Görlach A (2005) Reactive oxygen species in the control of hypoxia-inducible factor-mediated gene expression. Semin Cell Dev Biol 16: 474-86.

Kikuchi K, Terauchi K, Wada M, Hirano HY (2003) The plant MITE mPing is mobilized in anther culture. Nature 421: 167-70.

Kikuchi Y, Hosokawa T, Nikoh N, Meng XY, Kamagata Y, Fukatsu T (2009) Host-symbiont co-speciation and reductive genome evolution in gut symbiotic bacteria of acanthosomatid stinkbugs. BMC Biol 7: 2.

Kilic M, Kasperczyk H, Fulda S, Debatin KM (2007) Role of hypoxia inducible factor-1 alpha in modulation of apoptosis resistance. Oncogene 26: 2027–38.

Killick SC, Obbard DJ, West SA, Little TJ (2008) Parasitism and breeding system variation in North American populations of Daphnia pulex. Ecol Res 23: 235-40.

Killingback T, Bieri J, Flatt T (2006) Evolution in group-structured populations can resolve the tragedy of the commons. Proc Biol Sci 273: 1477–81.

Kilpatrick ST, Rand DM (1995) Conditional hitchhiking of mitochondrial DNA: frequency shifts of Drosophila melanogaster mtDNA variants depend on nuclear genetic background. Genetics 141: 1113-24.

Kim A, Zhong W, Oberley TD (2004) Reversible modulation of cell cycle kinetics in NIH/3T3 mouse fibroblasts by inducible overexpression of mitochondrial manganese superoxide dismutase. Antioxid Redox Signal 6: 489–500.

Kim A, Oberley LW, Oberley TD (2005) Induction of apoptosis by adenovirus-mediated manganese superoxide dismutase overexpression in SV-40-transformed human fibroblasts. Free Radic Biol Med 39: 1128–41.

Kim A, Joseph S, Khan A, Epstein CJ, Sobel R, Huang TT (2010) Enhanced expression of mitochondrial superoxide dismutase leads to prolonged in vivo cell cycle progression and up-regulation of mitochondrial thioredoxin. Free Radic Biol Med 48: 1501-12.

Kim GJ, Georg I, Scherthan H, Merkenschlager M, Guillou F, et al. (2010) Dicer is required for Sertoli cell function and survival. Int J Dev Biol 54: 867–75.

Kim H, Lee B-S, Tomita M, Kanai A (2010) Transcription-associated mutagenesis increases protein sequence diversity more effectively than does random mutagenesis in Escherichia coli. PLoS ONE 5: e10567.

Kim HC, Byun JS, Lee TK, Jeong CW, Ahn M, Shin T (2007) Expression of nitric oxide synthase isoforms in the testes of pigs. Anat Histol Embryol 36: 135-8.

Kim JW, Tchernyshyov I, Semenza GL, Dang CV (2006) HIF-1-mediated expression of pyruvate dehydrogenase kinase: a metabolic switch required for cellular adaptation to hypoxia. Cell Metab 3:177–85.

Kim KH, Oh DS, Jeong JH, Shin BS, Joo BS, Lee KS (2004) Follicular blood flow is a better predictor of the outcome of in vitro fertilization-embryo transfer than follicular fluid vascular endothelial growth factor and nitric oxide concentrations. Fertil Steril 82: 586–92.

Kim M, Trinh BN, Long TI, Oghamian S, Laird PW (2004) Dnmt1 deficiency leads to enhanced microsatellite instability in mouse embryonic stem cells. Nucleic Acids Res 32: 5742-9.

Kim MH, Kim HB, Yoon SP, Lim SC, Cha MJ, et al. (2013) Colon cancer progression is driven by APEX1-mediated upregulation of Jagged. J Clin Invest doi:10.1172/JCI65521.[Epub ahead of print]

Kim MS, Baek JH, Bae MK, Kim KW (2001) Human rad21 gene, hHR21(SP), is downregulated by hypoxia in human tumor cells. Biochem Biophys Res Commun 281: 1106–12.

Kim N, Abdulovic AL, Gealy R, Lippert MJ, Jinks-Robertson S (2007) Transcription-associated mutagenesis in yeast is directly proportional to the level of gene expression and influenced by the direction of DNA replication. DNA Repair (Amst) 6: 1285–96.

Kim N, Jinks-Robertson S (2012) Transcription as a source of genome instability. Nat Rev Genet 13: 204-14.

Kim NK, Lee SH, Sohn TJ, Roy R, Mitra S, et al. (2000) Spatial expression of a DNA repair gene, Nmethylpurine-DNA glycosylase (MPG) during development in mice. Anticancer Res 20: 3037–43.

Kim PS, Lee PP, Levy D (2011) A theory of immunodominance and adaptive regulation. Bull Math Biol 73: 1645-65.

Kim VN (2006) Small RNAs just got bigger: Piwi-interacting RNAs (piRNAs) in mammalian testes. Genes Dev 20: 1993-7.

Kim VN, Nam JW (2006) Genomics of microRNA. Trends Genet 22: 165–73.

Kim Y, Orr HA (2005) Adaptation in sexuals vs. asexuals: clonal interference and the Fisher-Muller model. Genetics 171: 1377–86.

Kimball SR, Horetsky RL, Ron D, Jefferson LS, Harding HP (2003) Mammalian stress granules represent sites of accumulation of stalled translation initiation complexes. Am J Physiol Cell Physiol 284: C273-84.

Kimmerer RW (1991) Reproductive ecology of Tetraphis pellucida. I. Population density and reproductive mode. Bryologist 94: 255-60.

Kimmins S, Sassone-Corsi P (2005) Chromatin remodelling and epigenetic features of germ cells. Nature 434: 583-9.

Kimura M (1960) Optimum mutation rate and degree of dominance as determined by the principle of minimum genetic load. J Genet 57: 21–34.

Kimura M (1962) On the probability of fixation of mutant genes in a population. Genetics 47: 713–9.

Kimura M (1965) A stochastic model concerning the maintenance of genetic variability in quantitative characters. Proc Natl Acad Sci USA 54: 731–6.

Kimura M (1967) On the evolutionary adjustment of spontaneous mutation rates. Genet Res 9: 23–34.

Kimura M (1968) Evolutionary rate at the molecular level. Nature 217: 624-6.

Kimura M (1983) The neutral theory of molecular evolution. Cambridge, UK: Cambridge University Press.

Kimura M (1991) The neutral theory of molecular evolution: A review of recent evidence. Jpn J Genet 66: 367–86.

Kimura M, Crow JF (1964) The number of alleles that can be maintained in a finite population. Genetics 49: 725–38.

Kimura M, Itoh N, Takagi S, Sasao T, Takahashi A, Masumori N, Tsukamoto T (2003) Balance of apoptosis and proliferation of germ cells related to spermatogenesis in aged men. J Androl 24: 185–91.

Kimura S, Maekawa T, Hirakawa K, Murakami A, Abe T (1995) Alterations of c-Myc expression by antisense oligodeoxynucleotides enhance the induction of apoptosis in HL-60 cells. Cancer Res 55: 1379–84.

Kimura S, Zhang GX, Nishiyama A, Shokoji T, Yao L, et al. (2005) Role of NAD(P)H oxidase- and mitochondria-derived reactive oxygen species in cardioprotection of ischemic reperfusion injury by angiotensin II. Hypertension 45: 860–866.

King DG (2012a) Indirect selection of implicit mutation protocols. Ann NY Acad Sci 1267: 45-52.

King DG (2012b) Evolution of simple sequence repeats as mutable sites. In: Hannan AJ, ed. Tandem Repeat Polymorphisms: Genetic Plasticity, Neural Diversity and Disease. New York, NY: Springer Verlag.

King DG, Soller M, Kashi Y (1997) Evolutionary tuning knobs. Endeavour 21: 36–40.

King DG, Kashi Y (2007) Mutability and evolvability: indirect selection for mutability. Heredity 99: 123-4.

King JL (1967) Continuously distributed factors affecting fitness. Genetics 55: 483-92.

King JL, Jukes TH (1969) Non-Darwinian evolution. Science 164: 788–98.

King KC, Lively CM (2009) Geographic variation in sterilizing parasite species and the Red Queen. Oikos 118: 1416–20.

King KC, Delph LF, Jokela J, Lively CM (2009) The geographic mosaic of sex and the Red Queen. Curr Biol 19: 1438–41.

King KC, Seppälä O, Neiman M (2012) Is more better? Polyploidy and parasite resistance. Biol Lett 8: 598-600.

King LM, Schaal BA (1990) Genotypic variation within asexual lineages of Taraxacum officinale. Proc Natl Acad Sci USA 87: 998–1002.

King OD, Masel J (2007) The evolution of bet-hedging adaptations to rare scenarios. Theor Popul Biol 72: 560-75.

King RC (1970) Ovarian development in Drosophila melanogaster. New York, NY: Academic Press.

Kingsolver JG, Hoekstra HE, Hoekstra JM, Berrigan D, Vignieri SN, et al. (2001) The strength of phenotypic selection in natural populations. Am Nat 157: 245–61.

Kinnunen L, Pöyry T, Hovi T (1992) Genetic diversity and rapid evolution of poliovirus in human hosts. Curr Top Microbiol Immunol 176: 49-61.

Kirino Y, Mourelatos Z (2007) The mouse homolog of HEN1 is a potential methylase for Piwi-interacting RNAs. RNA 13: 1397-401.

Kirk DL (1998) Volvox: molecular-genetic origins of multicellularity and cellular differentiation. Cambridge, UK: Cambridge University Press.

Kirk DL (2003) Seeking the ultimate and proximate causes of Volvox multicellularity and cellular differentiation. Integr Comp Biol 43: 247-53.

Kirk DL, Kirk MM (1986) Heat shock elicits production of sexual inducer in Volvox. Science 231: 51–4.

Kirkpatrick M (1982) Sexual selection and the evolution of female choice. Evolution 36: 1–12.

Kirkpatrick M (1996) Good genes and direct selection in the evolution of mating preferences. Evolution 50: 2125–40.

Kirkpatrick M, Jenkins CD (1989) Genetic segregation and the maintenance of sexual reproduction. Nature 339: 300-1.

Kirkpatrick M, Ryan MJ (1991) The evolution of mating preferences and the paradox of the lek. Nature 350: 33–8.

Kirkpatrick M, Barton NH (1997) Evolution of a species range. Am Nat 150: 1–23.

Kirkwood TBL (1977) Evolution of ageing. Nature 270: 301-4.

Kirkwood TB, Holliday R (1979) The evolution of ageing and longevity. Proc R Soc Lond B Biol Sci 205: 531–46.

Kirkwood TB, Austad SN (2000) Why do we age? Nature 408: 233–8.

Kirschner M, Gerhart J (1998) Evolvability. Proc Natl Acad Sci USA 95: 8420-7.

Kirschner M, Gerhart JC (2005) The Plausibility of Life. London, UK: Yale University Press.

Kirschvink JL, Raub TD (2003) A methane fuse for the Cambrian explosion: true polar wander. CR Geosci 335: 71–83.

Kirzhner VM, Korol AB, Nevo E (1998) Complex limiting behavior of multilocus systems in cyclical environments. J Theor Biol 190: 215-25.

Kisdi É, Jacobs FJA, Geritz SAH (2001) Red queen evolution by cycles of evolutionary branching and extinction. Selection 2: 161–76.

Kishony R, Leibler S (2003) Environmental stresses can alleviate the average deleterious effect of mutations. J Biol 2: 14.

Kisielow P, von Boehmer H (1965) Development and selection of T cells: facts and puzzles. Adv Immunol 58: 87-209.

Kis-Papo T, Kirzhner V, Wasser SP, Nevo E (2003) Evolution of genomic diversity and sex at extreme environments: fungal life under hypersaline Dead Sea stress. Proc Natl Acad Sci USA 100: 14970–5.

Kisseljova NP, Kisseljov FL (2005) DNA demethylation and carcinogenesis. Biochemistry (Mosc) 70: 743-52.

Kitagawa H, Yamaoka I, Akimoto C, Kase I, Mezaki Y, Shimizu T, et al. (2007) A reduction state potentiates the glucocorticoid response through receptor protein stabilization. Genes Cells 12: 1281–7.

Kitamura Y, Scavarda N, Wells SA Jr, Jackson CE, Goodfellow PJ (1995) Two maternally derived missense mutations in the tyrosine kinase domain of the RET protooncogene in a patient with de novo MEN2B. Hum Mol Genet 4: 1987–8.

Kitano H (2004) Biological robustness. Nat Rev Genet 5: 826–37.

Kitay JI (1961) Sex differences in adrenal cortical secretion in the rat. Endocrinology 68: 818–24.

Kitcher P (1981) Explanatory unification. Philos Sci 48: 507-31.

Kitcher P (1984) Species. Philos Sci 51: 308-33.

Kivisaar M (2003) Stationary phase mutagenesis: mechanisms that accelerate adaptation of microbial populations under environmental stress. Environ Microbiol 5: 814-27.

Kivisild T, Shen P, Wall DP, Do B, Sung R, Davis K, et al. (2006) The role of selection in the evolution of human mitochondrial genomes. Genetics 172: 373–87.

Klapacz J, Bhagwat AS (2002) Transcription-dependent increase in multiple classes of base substitution mutations in Escherichia coli. J Bacteriol 184: 6866–72.

Klapacz J, Bhagwat AS (2005) Transcription promotes guanine to thymine mutations in the non-transcribed strand of an Escherichia coli gene. DNA Repair 4: 806–13.

Klase Z, Kale P, Winograd R, Gupta MV, Heydarian M, et al. (2007) HIV-1 TAR element is processed by Dicer to yield a viral micro-RNA involved in chromatin remodeling of the viral LTR. BMC Mol Biol 8: 63.

Klattenhoff C, Theurkauf W (2008) Biogenesis and germline functions of piRNAs. Development 135: 3-9.

Klaunig JE, Kamendulis LM (2004) The role of oxidative stress in carcinogenesis. Annu Rev Pharmacol Toxicol 44: 239-67.

Klaunig JE, Kamendulis LM, Hocevar BA (2010) Oxidative stress and oxidative damage in carcinogenesis. Toxicol Pathol 38: 96-109.

Klein NA, Battaglia DE, Fujimoto VY, Davis GS, Bremner WJ, Soules MR (1996) Reproductive aging: accelerated ovarian follicular development associated with a monotropic follicle-stimulating hormone rise in normal older women. J Clin Endocrinol Metab 81: 1038-45.

Kleisner K, Ivell R, Flegr J (2010) The evolutionary history of testicular externalization and the origin of the scrotum. J Biosci 35: 27–37.

Kleiven OT, Larsson P, Hobæk A (1992) Sexual reproduction in Daphnia magna requires three stimuli. Oikos 65: 197–206.

Kleizen B, Braakman I (2004) Protein folding and quality control in the endoplasmic reticulum. Curr Opin Cell Biol 16: 343-9.

Klekowski EJ Jr (1988) Mutation, Developmental Selection, and Plant Evolution. New York, NY: Columbia University Press.

Klekowski EJ (2003) Plant clonality, mutation, diplontic selection and mutational meltdown. Bot J Linn Soc 79: 61– 67.

Klekowski EJ Jr, Kazarinova-Fukshansky N (1984) Shoot apical meristems and mutation: selective loss of disadvantageous cell genotypes. Am J Bot 71: 28-34.

Klekowski EJ Jr, Godfrey PJ (1989) Ageing and mutation in plants. Nature 340: 389–91.

Klempner MS, Dinarello CA, Henderson WR, Gallin JI (1979) Stimulation of neutrophil oxygen-dependent metabolism by human leukocytic pyrogen. J Clin Invest 64: 996-1002.

Klenov MS, Lavrov SA, Stolyarenko AD, Ryazansky SS, Aravin AA, et al. (2007) Repeat-associated siRNAs cause chromatin silencing of retrotransposons in the Drosophila melanogaster germline. Nucleic Acids Res 35: 5430-8.

Klenova EM, Morse HC 3rd, Ohlsson R, Lobanenkov VV (2002) The novel BORIS + CTCF gene family is uniquely involved in the epigenetics of normal biology and cancer. Semin Cancer Biol 12: 399-414.

Kleven O, Fossøy F, Laskemoen T, Robertson RJ, Rudolfsen G, et al. (2009) Comparative evidence for the evolution of sperm swimming speed by sperm competition and female sperm storage duration in passerine birds. Evolution 63: 2466–73.

Klimes L, Klimesova J, Hendriks R, Van Groenendael J (1997) Clonal plant architecture: a comparative analysis of form and function. In: de Kroon H, Van Groenendael J, eds. The Ecology and Evolution of Clonal Plants. Leiden, The Netherlands: Backhuys Publishers. pp 1-29.

Klimesova J, Klimes L (2007) Bud banks and their role in vegetative regeneration – a literature review and proposal for simple classification and assessment. Perspect Plant Ecol Evol Syst 8: 115–29.

Klimova T, Chandel NS (2008) Mitochondrial complex III regulates hypoxic activation of HIF. Cell Death Differ 15: 660-6.

Kloc A, Martienssen R (2008) RNAi, heterochromatin and the cell cycle. Trends Genet 24: 511-7.

Kloc M, Zagrodzinska B (2001) Chromatin elimination: an oddity or a common mechanism in differentiation and development? Differentiation 68: 84–91.

Kloc M, Bilinski S, Etkin LD (2004a) The Balbiani body and germ cell determinants: 150 years later. Curr Top Dev Biol 59: 1–36.

Kloc M, Bilinski S, Dougherty MT, Brey EM, Etkin LD (2004b) Formation, architecture and polarity of female germline cyst in Xenopus. Dev Biol 266: 43–61.

Klomp H (1964) Intraspecific competition and the regulation of insect numbers. Annu Rev Entomol 9: 17-40.

Klomp H (1970) The determination of clutch size in birds. A review. Ardea 58: 1-124.

Kloosterman WP, Plasterk RH (2006) The diverse functions of microRNAs in animal development and disease. Dev Cell 11: 441–50.

Klose RJ, Zhang Y (2007) Regulation of histone methylation by demethylimination and demethylation. Nat Rev Mol Cell Biol 8: 307-18.

Klose SM, Smith CL, Denzel AJ, Kalko EK (2006) Reproduction elevates the corticosterone stress response in common fruit bats. J Comp Physiol A Neuroethol Sens Neural Behav Physiol 192: 341-50.

Kloss RJ, Nesse RM (1992) Trisomy: chromosome competition or maternal strategy? Ethol Sociobiol 13: 283–7.

Klotz T, Vorreuther R, Heidenreich A, Zumbe J, Engelmann U (1996) Testicular tissue oxygen pressure. J Urol 155: 1488–91.

Kluge J, Kessler M, Dunn RR (2006) What drives elevational patterns in diversity? A test of geometric constraints, climate and species pool effects for pteridophytes on an elevational gradient in Costa Rica. Global Ecol Biogeogr 15:358–71.

Klungland A, Rosewell I, Hollenbach S, Larsen E, Daly G, et al. (1999) Accumulation of premutagenic DNA lesions in mice defective in removal of oxidative base damage. Proc Natl Acad Sci USA 96: 13300–5.

Klütsch CF, Seppälä EH, Uhlén M, Lohi H, Savolainen P (2011) Segregation of point mutation heteroplasmy in the control region of dog mtDNA studied systematically in deep generation pedigrees. Int J Legal Med 125: 527-35.

Kmiec B, Woloszynska M, Janska H (2006) Heteroplasmy as a common state of mitochondrial genetic information in plants and animals. Curr Genet 50: 149-59.

Knapen MFCM, Zusterzeel PLM, Peters WHM, Steegers EAP (1999) Glutathione and glutathione-related enzymes in reproduction - A review. Eur J Obstet Gynecol Reprod Biol 82: 171-84.

Knapp R, Moore MC (1997) Male morphs in tree lizards have different testosterone responses to elevated levels of corticosterone. Gen Comp Endocrinol 107: 273-9.

Knaut H, Pelegri F, Bohmann K, Schwarz H, Nüsslein-Volhard C (2000) Zebrafish vasa RNA but not its protein is a component of the germ plasm and segregates asymmetrically before germline specification. J Cell Biol 149: 875–88.

Kniewel R, Keeney S (2009) Histone methylation sets the stage for meiotic DNA breaks. EMBO J 28: 81-3.

Knight TM, Steets JA, Vamosi JC, Mazer SJ, Burd M, et al. (2005) Pollen limitation of plant reproduction: pattern and process. Annu Rev Ecol Evol Syst 36: 467–97.

Knobil E, Neill JD, eds. (1995) The Physiology of Reproduction. New York, NY: Raven Press.

Knouf EC, Garg K, Arroyo JD, Correa Y, Sarkar D, et al. (2012) An integrative genomic approach identifies p73 and p63 as activators of miR-200 microRNA family transcription. Nucleic Acids Res 40: 499-510.

Knowlton N, Jackson JBC (1993) Inbreeding and outbreeding in marine invertebrates. In: Thornhill NW, ed. The natural history of inbreeding and outbreeding—theoretical and empirical perspectives. Chicago, IL: University of Chicago Press. pp 200–249.

Knudson CM, Tung KS, Tourtellotte WG, Brown GA, Korsmeyer SJ (1995) Bax-deficient mice with lymphoid hyperplasia and male germ cell death. Science 270: 96–9.

Koch AL (1971) The adaptive responses of Escherichia coli to a feast and famine existence. Adv Microb Physiol 6: 147-217.

Kocher TD (2004) Adaptive evolution and explosive speciation: the cichlid fish model. Nat Rev Genet 5: 288-98.

Kodaira K, Takahashi R, Hirabayashi M, Suzuki T, Obinata M, Ueda M (1996) Overexpression of c-myc induces apoptosis at the prophase of meiosis of rat primary spermatocytes. Mol Reprod Dev 45: 403-10.

Kodaman PH, Behrman HR (2001) Endocrine-regulated and protein kinase C-dependent generation of superoxide by rat preovulatory follicles. Endocrinology 142: 687–93.

Koehler CM, Lindberg GL, Brown DR, Beitz DC, Freeman AE, et al. (1991) Replacement of bovine mitochondrial DNA by a sequence variant within one generation. Genetics 129: 247–55.

Koella JC (1988) The Tangled Bank: The maintenance of sexual reproduction through competitive interactions. J Evol Biol 1: 95-116.

Kogo N, Tazaki A, Kashino Y, Morichika K, Orii H, Mochii M, Watanabe K (2011) Germ-line mitochondria exhibit suppressed respiratory activity to support their accurate transmission to the next generation. Dev Biol 349: 462–9.

Kohle C, Bock KW (2007) Coordinate regulation of Phase I and II xenobiotic metabolisms by the Ah receptor and Nrf2. Biochem Pharmacol 73: 1853–62.

Kohler SW, Provost GS, Fieck A, Kretz PL, Bullock WO, et al. (1991) Spectra of spontaneous and mutagen-induced mutations in the lacI gene in transgenic mice. Proc Natl Acad Sci USA 88: 7958–62.

Kohn MH, Fang S, Wu C-I (2004) Inference of positive and negative selection on the 5' regulatory regions of Drosophila genes. Mol Biol Evol 21: 374–83.

Kohno S, Kubota S, Nakai Y (1998) Chromatin diminution and chromosome elimination in hagfish species. In: Jorgensen JM, Lomholt JP, Weber RE, Malte H, eds. The Biology of Hagfishes. London, UK: Chapman and Hall. pp 81–100.

Koji T, Izumi S, Tanno M, Moriuchi T, Nakane PK (1988) Localization in situ of c-myc mRNA and c-myc protein in adult mouse testis. Histochem J 20: 551-7.

Kojima NF, Kojima KK, Kobayakawa S, Higashide N, Hamanaka C, et al. (2010) Whole chromosome elimination and chromosome terminus elimination both contribute to somatic differentiation in Taiwanese hagfish Paramyxine sheni. Chromosome Res 18: 383-400.

Kokko H (2001) Fisherian and ‘‘good genes’’ benefits of mate choice: How (not) to distinguish them. Ecol Lett 4: 322–6.

Kokko H, Brooks R, Jennions MD, Morley J (2003) The evolution of mate choice and mating biases. Proc R Soc Lond Ser B 270: 653–64.

Kokko H, Rankin DJ (2006) Lonely hearts or sex in the city? Density-dependent effects in mating systems. Philos Trans R Soc B 361: 319–34.

Kokko H, López-Sepulcre A (2007) The ecogenetic link between demography and evolution: can we bridge the gap between theory and data? Ecol Lett 10: 773-82.

Kolluru GR, Grether GF, South SH, Dunlop E, Cardinali A, et al. (2006) The effects of carotenoid and food availability on resistance to a naturally occurring parasite (Gyrodactylus turnbulli) in guppies (Poecilia reticulata). Biol J Linn Soc 89: 301–9.

Kolodner RD, Marsischky GT (1999) Eukaryotic DNA mismatch repair. Curr Opin Genet Dev 9: 89–96.

Kolowrat C, Partensky F, Mella-Flores D, Le Corguillé G, Boutte C, et al. (2010) Ultraviolet stress delays chromosome replication inlight/darksynchronizedcellsofthemarine cyanobacterium Prochlorococcus marinus PCC9511. BMC Microbiol 10: 204.

Kolter R, Siegele DA, Tormo AA (1993) The stationary phase of the bacterial life cycle. Annu Rev Microbiol 47: 855-74.

Komarova EA, Diatchenko L, Rokhlin OW, Hill JE, Wang ZJ, et al. (1998) Stress-induced secretion of growth inhibitors: a novel tumor suppressor function of p53. Oncogene 17: 1089-96.

Komp Lindgren P, Marcusson LL, Sandvang D, Frimodt-Moller N, Hughes D (2005) Biological cost of single and multiple norfloxacin resistance mutations in Escherichia coli implicated in urinary tract infections. Antimicrob Agents Chemother 49: 2343–51.

Kon Y, Endoh D, Iwanaga T (1999) Expression of protein gene product 9.5, a neuronal ubiquitin C-terminal hydrolase, and its developing change in sertoli cells of mouse testis. Mol Reprod Dev 54: 333–41.

Kondo A, Safaei R, Mishima M, Niedner H, Lin X, Howell SB (2001) Hypoxia-induced enrichment and mutagenesis of cells that have lost DNA mismatch repair. Cancer Res 61: 7603-7.

Kondo M, Senoo-Matsuda N, Yanase S, Ishii T, Hartman PS, Ishii N (2005) Effect of oxidative stress on translocation of DAF-16 in oxygen-sensitive mutants, mev-1 and gas-1 of Caenorhabditis elegans. Mech Ageing Dev 126: 637–41.

Kondo Y (2009) Epigenetic cross-talk between DNA methylation and histone modifications in human cancers. Yonsei Med J 50: 455-63.

Kondoh H, Lleonart ME, Gil J, Wang J, Degan P, et al. (2005) Glycolytic enzymes can modulate cellular life span. Cancer Res 65: 177–85.

Kondoh H, Lleonart ME, Bernard D, Gil J (2007) Protection from oxidative stress by enhanced glycolysis; a possible mechanism of cellular immortalization. Histol Histopathol 22: 85–90.

Kondrashov AS (1988) Deleterious mutations and the evolution of sexual reproduction. Nature 336: 435-40.

Kondrashov AS (1993) Classification of hypotheses on the advantage of amphimixis. J Hered 84: 372–87.

Kondrashov AS (1994) Muller’s ratchet under epistatic selection. Genetics 136: 1469–73.

Kondrashov AS (1995a) Contamination of the genome by very slightly deleterious mutations: why have we not died 100 times over? J Theor Biol 175: 583-94.

Kondrashov AS (1995b) Modifiers of mutation-selection balance: general approach and the evolution of mutation rates. Genet Res 66: 53–69.

Kondrashov AS (2003) Direct estimates of human per nucleotide mutation rates at 20 loci causing Mendelian diseases. Hum Mutat 21: 12–27.

Kondrashov AS, Mina MV (1986) Sympatric speciation: when is it possible? Biol J Linn Soc 27: 201–223.

Kondrashov AS, Turelli M (1992) Deleterious mutations, apparent stabilizing selection and the maintenance of quantitative variation. Genetics 132: 603-18.

Kondrashov AS, Crow JF (1993) A molecular approach to estimating the human deleterious mutation rate. Hum Mutat 2: 229–34.

Kondrashov AS, Yampolsky LYu (1996) High genetic variability under under the balance between symmetric mutation and fluctuating stabilizing selection. Genet Res 68: 157-64.

Kondrashov AS, Shpak M (1998) On the origin of species by means of assortative mating. Proc R Soc Lond B 265: 2273–8.

Kondrashov AS, Kondrashov FA (1999) Interactions among quantitative traits in the course of sympatric speciation. Nature 400: 351–4.

Kondrashov FA, Kondrashov AS (2001) Multidimensional epistasis and the disadvantage of sex. Proc Natl Acad Sci USA 98: 12089–92.

Kong A, Frigge ML, Masson G, Besenbacher S, Sulem P, et al. (2012) Rate of de novo mutations and the importance of father's age to disease risk. Nature 488: 471-5.

Kongruttanachok N, Phuangphairoj C, Thongnak A, Ponyeam W, Rattanatanyong P, et al. (2010) Replication independent DNA double-strand break retention may prevent genomic instability. Mol Cancer 9: 70.

Konkel MK, Batzer MA (2010) A mobile threat to genome stability: The impact of non-LTR retrotransposons upon the human genome. Semin Cancer Biol 20: 211-21.

Kono Y, Fridovich I (1982) Superoxide radical inhibits catalase. J Biol Chem 257: 5751-4.

Konopka JB, Fields S (1992) The pheromone signal pathway in Saccharomyces cerevisiae. Antonie Van Leeuwenhoek 62: 95-108.

Koochmeshgi J, Ladonni S, Hosseini-Mazinani SM (2004) Investigations on the nature of the cost of reproduction: susceptibility to heat stress in fruitflies. Ann NY Acad Sci 1019: 368-9.

Koolhaas JM, Bartolomucci A, Buwalda B, de Boer SF, Flügge G, et al. (2011) Stress revisited: a critical evaluation of the stress concept. Neurosci Biobehav Rev 35: 1291-301.

Koonin EV (2012) A half-century after the molecular clock: new dimensions of molecular evolution. EMBO Rep 13: 664-6.

Koonin EV, Dolja VV (1993) Evolution and taxonomy of positive-strand RNA viruses: implications of comparative analysis of amino acid sequences. Crit Rev Biochem Mol Biol 28: 375–430.

Koonin EV, Wolf YI (2006) Evolutionary systems biology: links between gene evolution and function. Curr Opin Biotechnol 17: 481–7.

Koonin EV, Wolf YI (2009) Is evolution Darwinian or/and Lamarckian? Biol Direct 4: 42.

Koonin EV, Wolf YI (2010) Constraints and plasticity in genome and molecular-phenome evolution. Nat Rev Genet 11: 487-98.

Koonin EV, Wolf YI (2012) Evolution of microbes and viruses: a paradigm shift in evolutionary biology? Front Cell Infect Microbiol 2: 119.

Kopp M, Hermisson J (2006) The evolution of genetic architecture under frequency-dependent disruptive selection. Evolution 60: 1537-50.

Koppers AJ, De Iuliis GN, Finnie JM, McLaughlin EA, Aitken RJ (2008) Significance of mitochondrial reactive oxygen species in the generation of oxidative stress in spermatozoa. J Clin Endocrinol Metab 93: 3199–207.

Koppers AJ, Garg ML, Aitken RJ (2010) Stimulation of mitochondrial reactive oxygen species production by unesterified, unsaturated fatty acids in defective human spermatozoa. Free Radic Biol Med 48: 112-9.

Korbecka G, Klinkhamer PGL, Vrieling K (2002) Selective embryo abortion hypothesis revisited: a molecular approach. Plant Biol 4:298–310.

Kordiš D, Gubenšek F (1998) Unusual horizontal transfer of a long interspersed nuclear element between distant vertebrate classes. Proc Natl Acad Sci USA 95: 10704–9.

Korfiatis KJ, Stamou GP (1999) Habitat templets and the changing worldview of ecology. Biol Philos 14: 375–93.

Korhonen HM, Meikar O, Yadav RP, Papaioannou MD, Romero Y, et al. (2011) Dicer is required for haploid male germ cell differentiation in mice. PLoS One 6: e24821.

Koritzinsky M, Wouters BG (2007) Hypoxia and regulation of messenger RNA translation. Methods Enzymol 435: 247-73.

Kornberg A, Baker TA (1992) DNA Replication. New York, NY: W. H. Freeman.

Kornberg RD, Lorch Y (1999) Twenty-five years of the nucleosome, fundamental particle of the eukaryote chromosome. Cell 98: 285-94.

Korol AB, Preygel IA, Preygel SI (1994) Recombination Variability and Evolution. London, UK: Chapman and Hall.

Korol AB, Kirzhner VM, Ronin YI, Nevo E (1996) Cyclical environmental changes as a factor maintaining genetic polymorphism. 2. Diploid selection for an additive trait. Evolution 50: 1432-41.

Korotkova GP (1970) Regeneration and somatic embryogenesis in sponges. In: Fry WG, ed. The Biology of the Porifera. London, UK: Academic Press. pp 423-436.

Korpelainen EI, Karkkainen MJ, Tenhunen A, Lakso M, Rauvala H, Vierula M, et al. (1998) Overexpression of VEGF in testis and epididymis causes infertility in transgenic mice: Evidence for nonendothelial targets for VEGF. J Cell Biol 143: 1705–12.

Korpelainen H (1998) Labile sex expression in plants. Biol Rev Camb Philos Soc 73: 157-80.

Kortner TM, Skugor S, Penn MH, Mydland LT, Djordjevic B, et al. (2012) Dietary soya saponin supplementation to pea protein concentrate reveals nutrigenomic interactions underlying enteropathy in Atlantic salmon (Salmo salar). BMC Vet Res 8: 101.

Korzets A, Chagnac A, Weinstein T, Ori Y, Malachi T, Gafter U (1999) H2O2 induces DNA repair in mononuclear cells: evidence for association with cytosolic Ca2+ fluxes. J Lab Clin Med 133: 362-9.

Kosaka N, Iguchi H, Yoshioka Y, Hagiwara K, Takeshita F, Ochiya T (2012) Competitive interactions of cancer cells and normal cells via secretory microRNAs. J Biol Chem 287: 1397-405.

Koshi JM, Goldstein RA (1995) Context-dependent optimal substitution matrices. Protein Eng 8: 641–645.

Koshi JM, Goldstein RA (1998) Models of natural mutations including site heterogeneity. Proteins 32: 289–95.

Koshiji M, Kageyama Y, Pete EA, Horikawa I, Barrett JC, Huang LE (2004) HIF-1alpha induces cell cycle arrest by functionally counteracting Myc. Embo J 23: 1949–56.

Koshiji M, To KK, Hammer S, Kumamoto K, Harris AL, Modrich P, Huang LE (2005) HIF-1alpha induces genetic instability by transcriptionally downregulating MutSalpha expression. Mol Cell 17: 793-803.

Koshimizu U, Sawada K, Tajima Y, Watanabe D, Nishimune Y (1991) White-spotting mutations affect the regenerative differentiation of testicular germ cells: demonstration by experimental cryptorchidism and its surgical reversal. Biol Reprod 45: 642-8.

Kosiol C, Vinar T, da Fonseca RR, Hubisz MJ, Bustamante CD, et al. (2008) Patterns of positive selection in six mammalian genomes. PLoS Genet 4: e1000144.

Koskela E (1998) Offspring growth, survival and reproductive success in the bank vole: a litter size manipulation experiment. Oecologia 115: 379-84.

Koskella B, Lively CM (2007) Advice of the Rose: experimental coevolution of a trematode parasite and its snail host. Evolution 62: 152–9.

Koskella B, Lively CM (2009) Evidence for negative frequency-dependent selection during experimental coevolution of a freshwater snail and a sterilizing trematode. Evolution 63: 2213-21.

Koster R, Timmer-Bosscha H, Bischoff R, Gietema JA, de Jong S (2011) Disruption of the MDM2-p53 interaction strongly potentiates p53-dependent apoptosis in cisplatin-resistant human testicular carcinoma cells via the Fas/FasL pathway. Cell Death Dis 2: e148.

Kostic T, Andric S, Kovacevic R, Maric D (1998) The involvement of nitric oxide synthase in stress-impaired testicular steroidogenesis. Eur J Pharmacol 346: 267-73.

Kostova Z, Wolf DH (2003) For whom the bell tolls: protein quality control of the endoplasmic reticulum and the ubiquitin-proteasome connection. EMBO J 22: 2309–17.

Kotaja N, Bhattacharyya SN, Jaskiewicz L, Kimmins S, Parvinen M, et al. (2006) The chromatoid body of male germ cells: similarity with processing bodies and presence of Dicer and microRNA pathway components. Proc Natl Acad Sci USA 103: 2647–52.

Kotaja N, Sassone-Corsi P (2007) The chromatoid body: a germ-cell-specific RNA-processing centre. Nat Rev Mol Cell Biol 8: 85–90.

Kotani T, Ozaki M, Matsuoka K, Snell TW, Hagiwara A (2001) Reproductive isolation among geographically and temporally isolated marine Brachionus strains. Hydrobiologia 446/447: 283–90.

Kotch LE, Iyer NV, Laughner E, Semenza GL (1999) Defective vascularization of HIF-1alpha-null embryos is not associated with VEGF deficiency but with mesenchymal cell death. Dev Biol 209: 254–67.

Kothari S, Thompson A, Agarwal A, du Plessis SS (2010) Free radicals: their beneficial and detrimental effects on sperm function. Indian J Exp Biol 48: 425-35.

Kotiaho JS, Simmons LW, Tomkins JL (2001) Towards a resolution of the lek paradox. Nature 410: 684–6.

Kotiaho JS, Simmons LW, Hunt J, Tomkins JL (2003) Males influence maternal effects that promote sexual selection: a quantitative genetic experiment with dung beetles Onthophagus taurus. Am Nat 161: 852–9.

Koturbash I, Kutanzi K, Hendrickson K, Rodriguez-Juarez R, Kogosov D, Kovalchuk O (2008) Radiation-induced bystander effects in vivo are sex specific. Mutat Res 642: 28–36.

Koudele KA, Napolitano RJ, Aulerich RJ (1986) Inability to perceive photoperiod affects testes size and testosterone secretion in mink. Biol Reprod 34 Suppl. 1: 66.

Kourmouli N, Jeppesen P, Mahadevaiah S, Burgoyne P, Wu R, et al. (2004) Heterochromatin and tri-methylated lysine 20 of histone H4 in animals. J Cell Sci 117: 2491–501.

Kouyos RD, Silander OK, Bonhoeffer S (2007) Epistasis between deleterious mutations and the evolution of recombination. Trends Ecol Evol 22: 308–15

Kovalchuk I, Kovalchuk O, Kalck V, Boyko V, Filkowski J, et al. (2003) Pathogen-induced systemic plant signal triggers DNA rearrangements. Nature 423: 760–2.

Kovalchuk O, Baulch JE (2008) Epigenetic changes and nontargeted radiation effects–is there a link? Environ Mol Mutagen 49: 16–25.

Kowaltowski AJ, Vercesi AE (1999) Mitochondrial damage induced by conditions of oxidative stress. Free Radic Biol Med 26: 463–71.

Kowaltowski AJ, de Souza-Pinto NC, Castilho RF, Vercesi AE (2009) Mitochondria and reactive oxygen species. Free Radic Biol Med 47: 333-43.

Koyama N, Ishibashi K, Kuwahara M, Inase N, Ichioka M, et al. (1998) Cloning and functional expression of human aquaporin 8 cDNA and analysis of its gene. Genomics 54: 169–72.

Koziol MJ, Rinn JL (2010) RNA traffic control of chromatin complexes. Curr Opin Genet Dev 20: 142–8.

Kozlov AV, Staniek K, Nohl H (1999) Nitrite reductase activity is a novel function of mammalian mitochondria. FEBS Lett 454: 127–30.

Kozlowski J, Stearns SC (1989) Hypotheses for the production of excess zygotes: models of bet-hedging and selective abortion. Evolution 43: 1369–77.

Kozlowski P, de Mezer M, Krzyzosiak WJ (2010) Trinucleotide repeats in human genome and exome. Nucleic Acids Res 38: 4027-39.

Kraaijeveld AR, Godfray HCJ (1997) Trade-off between parasitoid resistance and larval competitor ability in Drosophila melanogaster. Nature 389: 278–80.

Krakauer DC, Mira A (1999) Mitochondria and germ cell death. Nature 400: 125-6.

Krakauer DC, Plotkin JB (2002) Redundancy, antiredundancy, and the robustness of genomes. Proc Natl Acad Sci USA 99: 1405–9.

Krakauer T (1986) Human interleukin 1. Crit Rev Immunol 6: 213-44.

Krämer, A, Ho AD (2001) Centrosome aberrations and cancer. Onkologie 24: 538–44.

Kramer MG, Templeton AR (2001) Life-history changes that accompany the transition from sexual to parthenogenetic reproduction in Drosophila mercatorum. Evolution 55: 748–61.

Krawczyk Z, Mali P, Parvinen M (1988) Expression of a testis-specific hsp70 gene-related RNA in defined stages of rat seminiferous epithelium. J Cell Biol 107: 1317-23.

Kraytsberg Y, Schwartz M, Brown TA, Ebralidse K, Kunz WS, et al. (2004) Recombination of human mitochondrial DNA. Science 304: 981.

Krebs CJ, Jarvis ED, Pfaff DW (1999) The 70-kDa heat shock cognate protein (Hsc73) gene is enhanced by ovarian hormones in the ventromedial hypothalamus. Proc Natl Acad Sci USA 96: 1686-91.

Krebs RA, Loeschcke V (1994) Costs and benefits of activation of the heat-shock response in Drosophila melanogaster. Funct Ecol 8: 730–7.

Kreft H, Jetz W (2007) Global patterns and determinants of vascular plant diversity. Proc Natl Acad Sci USA 104: 5925-30.

Kremer N, Voronin D, Charif D, Mavingui P, Mollereau B, Vavre F (2009) Wolbachia interferes with ferritin expression and iron metabolism in insects. PLoS Pathog 5: e1000630.

Kreuzer J, Viedt C, Brandes RP, Seeger F, Rosenkranz AS, et al. (2003) Platelet-derived growth factor activates production of reactive oxygen species by NAD(P)H oxidase in smooth muscle cells through Gi1, 2. FASEB J 17: 38–40.

Kreuzer KN (2005) Interplay between DNA replication and recombination in prokaryotes. Annu Rev Microbiol 59: 43-67.

Kriaucionis S, Heintz N (2009) The nuclear DNA base 5-hydroxymethylcytosine is present in Purkinje neurons and the brain. Science 324: 929–30.

Krieg AJ, Rankin EB, Chan D, Razorenova O, Fernandez S, Giaccia AJ (2010) Regulation of the histone demethylase JMJD1A by hypoxia-inducible factor 1 alpha enhances hypoxic gene expression and tumor growth. Mol Cell Biol 30: 344-53.

Krisher RL (2004) The effect of oocyte quality on development. J Anim Sci 82: E14–E23.

Krisher RL, Bavister BD (1998) Responses of oocytes and embryos to the culture environment. Theriogenology 59: 103–14.

Krisher RL, Bavister BD (1999) Enhanced glycolysis after maturation of bovine oocytes in vitro is associated with increased developmental competence. Mol Reprod Dev 53: 19–26.

Krisher RL, Prather RS (2012) A role for the Warburg effect in preimplantation embryo development: metabolic modification to support rapid cell proliferation. Mol Reprod Dev 79: 311-20.

Krishnamurthy H, Weinbauer GF, Aslam H, Yeung CH, Neischlag E (1998) Quantification of apoptotic testicular germ cells in normal and methoxyacetic acid-treated mice as determined by flow cytometry. J Androl 19: 710-7.

Krol J, Loedige I, Filipowicz W (2010) The widespread regulation of microRNA biogenesis, function and decay. Nat Rev Genet 11: 597-610.

Krömer T, Kessler M, Gradstein SR, Acebey A (2005) Diversity patterns of vascular epiphytes along an elevational gradient in the Andes. J Biogeogr 32: 1799–809.

Kroon FJ, Munday PL, Westcott DA, Hobbs JP, Liley NR (2005) Aromatase pathway mediates sex change in each direction. Proc Biol Sci 272: 1399-405.

Krueger J, Liu D, Scholz K, Zimmer A, Shi Y, et al. (2011) Flt1 acts as a negative regulator of tip cell formation and branching morphogenesis in the zebrafish embryo. Development 138: 2111–20.

Krug AZ, Jablonski D, Valentine JW (2007) Contrarian clade confirms the ubiquity of spatial origination patterns in the production of latitudinal diversity gradients. Proc Natl Acad Sci USA 104: 18129–34.

Kruszewski M (2003) Labile iron pool: the main determinant of cellular response to oxidative stress. Mutat Res 531: 81–92.

Krutzsch PH (1955) Observations on the Mexican free-tailed bat, Tadarida mexicana. J Mammal 36: 236-42.

Krutzsch PH, Crichton EG (1987) Reproductive biology of the male little mastiff bat, Mormopterus planiceps (Chiroptera: Molossidae), in southeast Australia. Am J Anat 178: 353-68.

Kryazhimskiy S, Tkacik G, Plotkin JB (2009) The dynamics of adaptation on correlated fitness landscapes. Proc Natl Acad Sci USA 106: 18638-43.

Kryazhimskiy S, Draghi JA, Plotkin JB (2011) Evolution. In evolution, the sum is less than its parts. Science 332: 1160-1.

Krylov DM, Wolf YI, Rogozin IB, Koonin EV (2003) Gene loss, protein sequence divergence, gene dispensability, expression level, and interactivity are correlated in eukaryotic evolution. Genome Res 13: 2229–35.

Kryston TB, Georgiev AB, Pissis P, Georgakilas AG (2011) Role of oxidative stress and DNA damage in human carcinogenesis. Mutat Res 711: 193-201.

Ku WW, Wine RN, Chae BY, Ghanayem BI, Chapin RE (1995) Spermatocyte toxicity of 2-methoxyethanol (ME) in rats and guinea pigs: evidence for the induction of apoptosis. Toxicol Appl Pharmacol 134: 100-10.

Kuang H, Padmanabhan C, Li F, Kamei A, Bhaskar PB, et al. (2009) Identification of miniature inverted-repeat transposable elements (MITEs) and biogenesis of their siRNAs in the Solanaceae: new functional implications for MITEs. Genome Res 19: 42-56.

Kuang Y, Fagan WF, Loladze I (2003) Biodiversity, habitat area, resource growth rate and interference competition. Bull Math Biol 65: 497–518.

Kubanek J, Snell TW (2008) Quorum sensing in rotifers. In: Winans SC, Bassler BL, eds. Chemical Communication Among Bacteria. Washington DC: ASM Press. pp 453–461.

Kubo R, Toda M, Hashitsume N (1985) Statistical physics II. Berlin, Germany: Springer.

Kubohara Y (1997) DIF-1, putative morphogen of D-discoideum, suppresses cell growth and promotes retinoic acid-induced cell differentiation in HL-60. Biochem Biophys Res Commun 236: 418-22.

Kubohara Y (1999) Effects of differentiation-inducing factors of Dictyostelium discoideum on human leukemia K562 cells: DIF-3 is the most potent anti-leukemic agent. Eur J Pharmacol 381: 57-62.

Kubohara Y, Kimura C, Tatemoto K (1995a) Putative morphogen, DIF, of Dictyostelium discoideum induces apoptosis in rat pancreatic AR42J cells. Dev Growth Differ 37: 711-6.

Kubohara Y, Saito Y, Tatemoto K (1995b) Differentiation-inducing factor of D-discoideum raises intracellular calcium concentration and suppresses cell growth in rat pancreatic AR42J cells. FEBS Lett 359: 122.

Kubohara Y, Kabeya K, Matsui H, Ishikawa K (1998) Effects of DIF-1, an anti-tumor agent isolated from Dictyostelium discoideum, on rat gastric mucosal RGM-1 and leptomeningeal cells. Zool Sci 15: 713-21.

Kubota S, Ishibashi T, Kohno S (1997) A germline restricted, highly repetitive DNA sequence in Paramyxine atami: An interspecifically conserved, but somatically eliminated, element. Mol Gen Genet 256: 252–6.

Kubota S, Takano J, Tsuneishi R, Kobayakawa S, Fujikawa N, et al. (2001) Highly repetitive DNA families restricted to germ cells in a Japanese hagfish (Eptatretus burgeri): A hierarchical and mosaic structure in eliminated chromosomes. Genetica 111: 319–28.

Kuchino Y, Mori F, Kasai H, Inoue H, Iwai S, Miura K, Ohtsuka E, Nishimura S (1987) Misreading of DNA templates containing 8-hydroxydeoxyguanosine at the modified base and at adjacent residues. Nature 327: 77–9.

Kudoh H, Shibaike H, Takasu H, Whigham DF, Kawano S (1999) Genet structure and determinants of clonal structure in a temperate deciduous woodland herb Uvularia perfoliata. J Ecol 87: 244–57.

Kugelberg E,Kofoid E, Reams AB, Andersson DI, Roth JR (2006) Multiple pathways of selected gene amplification during adaptive mutation. Proc Natl Acad Sci USA 103: 17319–24.

Kugu K, Ratts VS, Piquette GN, Tilly KI, Tao XJ, et al. (1998) Analysis of apoptosis and expression of bcl-2 gene family members in the human and baboon ovary. Cell Death Differ 5: 67-76.

Kuhn G, Hijri M, Sanders IR (2001) Evidence for the evolution of multiple genomes in arbuscular mycorrhizal fungi. Nature 414: 745–8.

Kukucka MA, Misra HP (1993) The antioxidant defense system of isolated guinea pig Leydig cells. Mol Cell Biochem 126: 1-7.

Kula K, Walczak-Jedrzejowska R, Slowikowska-Hilczer J, Oszukowska E (2001) Estradiol enhances the stimulatory effect of FSH on testicular maturation and contributes to precocious initiation of spermatogenesis. Mol Cell Endocrinol 178: 89–97.

Kulkarni AC, Kuppusamy P, Parinandi N (2007) Oxygen, the lead actor in the pathophysiologic drama: enactment of the trinity of normoxia, hypoxia, and hyperoxia in disease and therapy. Antioxid Redox Signal 9: 1717–30.

Kulshreshtha R, Ferracin M, Wojcik SE, Garzon R, Alder H, et al. (2007) A microRNA signature of hypoxia. Mol Cell Biol 27: 1859–67.

Kulshreshtha R, DavuluriRV, Calin GA, Ivan M (2008) A microRNA component of the hypoxic response. Cell Death Differ 15: 667–71.

Kültz D (2003) Evolution of the cellular stress proteome: from monophyletic origin to ubiquitous function. J Exp Biol 206: 3119–24.

Kültz D (2005) Molecular and evolutionary basis of the cellular stress response. Annu Rev Physiol 67: 225–57.

Kuma K, Iwabe N, Miyata T (1995) Functional constraints against variations on molecules from the tissue-level – slowly evolving brain-specific genes demonstrated by protein-kinase and immunoglobulin supergene families. Mol Ecol Evol 12: 123-30.

Kumar G, Patel D, Naz RK (1993) c-MYC mRNA is present in human sperm cells. Cell Mol Biol Res 39: 111-7.

Kumar S, Lata K, Mukhopadhyay S, Mukherjee TK (2010) Role of estrogen receptors in pro-oxidative and anti-oxidative actions of estrogens: a perspective. Biochim Biophys Acta 1800: 1127-35.

Kumar S, Subramanian S (2002) Mutation rates in mammalian genomes. Proc Natl Acad Sci USA 99: 803–8.

Kümmerli R, van den Berg P, Griffin AS, West SA, Gardner A (2010) Repression of competition favours cooperation: experimental evidence from bacteria. J Evol Biol 23: 699-706.

Kumpulainen T, Grapputo A, Mappes J (2004) Parasites and sexual reproduction in psychid moths. Evolution 58: 1511–20.

Kunkel TA (2004) DNA replication fidelity. J Biol Chem 279: 16895-8.

Kuno E (1981) Dispersal and the persistence of populations in unstable habitats: a theoretical note. Oecologia 49: 123-6.

Kuphal S, Winklmeier A, Warnecke C, Bosserhoff AK (2010) Constitutive HIF-1 activity in malignant melanoma. Eur J Cancer 46: 1159-69.

Kupiec JJ (1986) A probabilist theory for cell differentiation: the extension of Darwinian principles to embryogenesis. Specul Sci Technol 9: 19-22.

Kupiec JJ (1996) A chance-selection model for cell differentiation. Cell Death Differ 3: 385-90.

Kupiec JJ (1997) A Darwinian theory for the origin of cellular differentiation. Mol Gen Genet 255: 201-8.

Kupiec JJ, Gandrillon O, Kolesnik D, Beslon G (2012) Special Issue: Chance at the heart of the cell. Progress in Biophysics and Molecular Biology. Vol. 110.

Kupwaide VA, Saidapur SK (1987) Effect of dexamethasone and ACTH on oocyte growth and recruitment in the frog Rana cyanophlyctis during the prebreeding vitellogenic phase. Gen Comp Endocrinol 65: 394-8.

Kuramochi-Miyagawa S, Kimura T, Ijiri TW, Isobe T, Asada N, et al. (2004) Mili, a mammalian member of piwi family gene, isessential for spermatogenesis. Development 131: 839–49.

Kuramochi-Miyagawa S, Watanabe T, Gotoh K, Totoki Y, Toyoda A, et al. (2008) DNA methylation of retrotransposon genes is regulated by Piwi family members MILI and MIWI2 in murine fetal testes. Genes Dev 22: 908–17.

Kuramochi-Miyagawa S, Watanabe T, Gotoh K, Takamatsu K, Chuma S, et al. (2010) MVH in piRNA processing and gene silencing of retrotransposons. Genes Dev 24: 887-92.

Kurata S, Matsumoto M, Hoshi M, Nakajima H (1993) Transcriptional control of the heme oxygenase gene during mouse spermatogenesis. Eur J Biochem 217: 633-8.

Kurath G, Palukaitis P (1990) Serial passage of infectious transcripts of a cucumber mosaic virus satellite RNA clone results in sequence heterogeneity. Virology 176: 8-15.

Kurjan J (1992) Pheromone response in yeast. Annu Rev Biochem 61: 1097-129.

Kuroki M, Voest EE, Amano S, Beerepoot LV, Takashima S, et al. (1996) Reactive oxygen intermediates increase vascular endothelial growth factor expression in vitro and in vivo. J Clin Invest 98: 1667-75.

Kurtén B (1959) Rates of evolution in fossil mammals. Cold Spring Harb Symp Quant Biol 24: 205–15.

Kurtz J, Kalbe M, Langefors S, Mayer I, Milinski M, Hasselquist D (2007) An experimental test of the immunocompetence handicap hypothesis in a teleost fish: 11-ketotestosterone suppresses innate immunity in three-spined sticklebacks. Am Nat 170: 509–19.

Kussell E, Leibler S (2005) Phenotypic diversity, population growth, and information in fluctuating environments. Science 309: 2075-8.

Kussell E, Kishony R, Balaban NQ, Leibler S (2005) Bacterial persistence: a model of survival in changing environments. Genetics 169: 1807–14.

Kuttler F, Mai S (2006) c-Myc, genomic instability and disease. Genome Dyn 1: 171-90.

Kuznetsov S, Lyanguzowa M, Bosch TCG (2001) Role of epithelial cells and programmed cell death in Hydra spermatogenesis. Zoology 104: 25-31.

Kvitek DJ, Sherlock G (2011) Reciprocal sign epistasis between frequently experimentally evolved adaptive mutations causes a rugged fitness landscape. PLoS Genet 7: e1002056.

Kwon J (2007) The new function of two ubiquitin C-terminal hydrolase isozymes as reciprocal modulators of germ cell apoptosis. Exp Anim 56: 71-7.

Kwon J, Lee SR, Yang KS, Ahn Y, Kim YJ, et al. (2004) Reversible oxidation and inactivation of the tumor suppressor PTEN in cells stimulated with peptide growth factors. Proc Natl Acad Sci USA 101: 16419-24.

Kwon J, Kikuchi T, Setsuie R, Ishii Y, Kyuwa S, Yoshikawa Y (2003) Characterization of the testis in congenitally ubiquitin carboxy-terminal hydrolase-1 (Uch-L1) defective (gad) mice. Exp Anim 52: 1–9.

Kwon J, Wang YL, Setsuie R, Sekiguchi S, Sato Y, et al. (2004) Two closely related ubiquitin C-terminal hydrolase isozymes function as reciprocal modulators of germ cell apoptosis in cryptorchid testis. Am J Pathol 165: 1367–74.

Kwon J, Mochida K, Wang Y, Sekiguchi S, Sankai T, et al. (2005) Ubiquitin C-terminal hydrolase L-1 is essential for the early apoptotic wave of germinal cells and for sperm quality control during spermatogenesis. Biol Reprod 73: 29–35.

Kwon OJ, Lee SM, Floyd RA, Park JW (1998) Thiol-dependent metal-catalyzed oxidation of copper, zinc superoxide dismutase. Biochim Biophys Acta 1387: 249–56.

Labat F, Pradillon O, Garry L, Peuchmaur M, Fantin B, et al. (2005) Mutator phenotype confers advantage in Escherichia coli chronic urinary tract infection pathogenesis. FEMS Immunol Med Microbiol 44: 317–21.

Laberge RM, Boissonneault G (2005) On the nature and origin of DNA strand breaks in elongating spermatids. Biol Reprod 73: 289–96.

Labra M, Ghiani A, Citterio S, Sgorbati S, Sala F, et al. (2002) Analysis of cytosine methylation pattern in response to water deficit in pea root tips. Plant Biol 4: 694–9.

Lacey SW, Naim HY, Magness RR, Gething M-J, Sambrook JF (1994) Expression of lactase-phlorizin hydrolase in sheep is regulated at the RNA level. Biochem J 302: 929–35.

Lachance C, Fortier M, Thimon V, Sullivan R, Bailey JL, Leclerc P (2010) Localization of Hsp60 and Grp78 in the human testis, epididymis and mature spermatozoa. Int J Androl 33: 33–44.

Lachapelle J, Bell J (2012) Evolutionary rescue of sexual and asexual populations in a deteriorating environment. Evolution 66: 3508–18.

Lachke SA, Joly S, Daniels K, Soll DR (2002) Phenotypic switching and filamentation in Candida glabrata. Microbiology 148: 2661-74.

Lachmann M, Blackston NW, Haig D, Kowald A, Michod RE, et al. (2003) Cooperation and conflict in the evolution of genomes, cells and multicellular organisms. In: Hammerstein P, ed. Genetic and cultural evolution of cooperation. Boston, MA: MIT Press. pp 327–356.

Lack D (1947a) The significance of clutch-size. Ibis 89: 302–52.

Lack D (1947b) Darwin’s Finches. Cambridge, UK: Cambridge University Press.

Lack D (1966) Population Studies of Birds. Oxford, UK: Clarendon Press.

Ladle RJ, Johnstone RA, Judson OP (1993) Coevolutionary dynamics of sex in a metapopulation: escaping the Red Queen. Proc R Soc Lond B 253: 155–60.

Ladoukakis E, Zouros E (2001) Recombination in animal mitochondrial DNA: evidence from published sequences. Mol Biol Evol 18: 2127–31.

Laforsch C, Tollrian R (2004) Inducible defenses in multipredator environments: cyclomorphosis in Daphnia cucullata. Ecology 85: 2302–11.

Lafyatis R, Denhez F, Williams T, Sporn M, Roberts A (1991) Sequence specific protein binding to and activation of the TGF-beta3 promoter through a repeated TCCC motif. Nucleic Acids Res 19: 6419-25.

Lagos-Quintana M, Rauhut R, Lendeckel W, Tuschl T (2001) Identification of novel genes coding for small expressed RNAs. Science 294: 853–8.

Laipis PJ, Van de Walle MJ, Hauswirth WW (1988) Unequal partitioning of bovine mitochondrial genotypes among siblings. Proc Natl Acad Sci USA 85: 8107–10.

Laird CD, Pleasant ND, Clark AD, Sneeden JLS, Hassan KMA, et al. (2004) Hairpin-bisulfite PCR: assessing epigenetic methylation patterns on complementary strands of individual DNA molecules. Proc Natl Acad Sci USA 101: 204–9.

Laird PW, Jaenisch R (1994) DNA methylation and cancer. Hum Mol Genet 3: 1487–95.

Laird PW, Jaenisch R (1996) The role of DNA methylation in cancer genetics and epigenetics. Annu Rev Genet 30: 441–64.

Lal A, Pan Y, Navarro F, Dykxhoorn DM, Moreau L, et al. (2009) miR-24-mediated downregulation of H2AX suppresses DNA repair in terminally differentiated blood cells. Nat Struct Mol Biol 16: 492-8.

Lamatsch DK, Stöck M (2009) Sperm-dependent parthenogenesis and hybridogenesis in teleost fishes. In: Schön I, Martens K, van Dijk P, eds. Lost sex. Dordrecht, The Netherlands: Springer. pp 399–432.

Lamb BC, Saleem M, Scott W, Thapa N, Nevo E (1998) Inherited and environmentally-induced differences in mutation frequencies between wild strains of Sordaria fimicola from “Evolution Canyon”. Genetics 149: 87–99.

Lamb BC, Mandaokar S, Bahsoun B, Grishkan I, Nevo E (2008) Differences in spontaneous mutation frequencies as a function of environmental stress in soil fungi at "Evolution Canyon," Israel. Proc Natl Acad Sci USA 105: 5792-6.

Lamb RY, Willey RB (1979) Are parthenogenetic and related bisexual insects equal in fertility? Evolution 33: 774-5.

Lamb RY, Wiley RB (1987) Cytological mechanisms of thelytokous parthenogenesis in insects. Genome 29: 367-9.

Lambeck K, Chappell J (2001) Sea level change through the last glacial cycle. Science 292: 679–86.

Lambert AJ, Merry BJ (2004) Effect of caloric restriction on mitochondrial reactive oxygen species production and bioenergetics: reversal by insulin. Am J Physiol Regul Integr Comp Physiol 286: R71-9.

Lambert S, Watson A, Sheedy DM, Martin B, Carr AM (2005) Gross chromosomal rearrangements and elevated recombination at an inducible site-specific replication fork barrier. Cell 121: 689–702.

Lamm E, Jablonka E (2008) The nurture of nature: hereditary plasticity in evolution. Philos Psychol 21: 305–19.

Lampe DJ, Witherspoon DJ, Soto-Adames FN, Robertson HM (2003) Recent horizontal transfer of mellifera subfamily mariner transposons into insect lineages representing four different orders shows that selection acts only during horizontal transfer. Mol Biol Evol 20: 554–62.

Lampert KP, Lamatsch DK, Epplen JT, Schartl M (2005) Evidence for a monophyletic origin of triploid clones of the Amazon molly, Poecilia formosa. Evol Int J Org Evol 59: 881–9.

Lampert KP, Schartl M (2008) The origin and evolution of a unisexual hybrid: Poecilia formosa. Philos Trans R Soc Lond B Biol Sci 363: 2901–9.

Lampert KP, Schartl M (2010) A little bit is better than nothing: the incomplete parthenogenesis of salamanders, frogs and fish. BMC Biol 8: 78.

LaMunyon CW, Ward S (1998) Larger sperm outcompete smaller sperm in the nematode Caenorhabditis elegans. Proc R SocLond Ser B 265: 1997–2002.

LaMunyon CW, Ward S (1999) Evolution of sperm size in nematodes: Sperm competition favours larger sperm. Proc Biol Sci 266: 263–7.

LaMunyon CW, Ward S (2002) Evolution of larger sperm in response to experimentally increased sperm competition in Caenorhabditis elegans. Proc Biol Sci 269: 1125–8.

Lancaster AK, Bardill JP, True HL, Masel J (2010) The spontaneous appearance rate of the yeast prion [PSI+] and its implications for the evolution of the evolvability properties of the [PSI+] system. Genetics 184: 393-400.

Lance VA (1997) Sex determination in reptiles: an update. Am Zool 37: 504–13.

Lance VA (2009) Is regulation of aromatase expression in reptiles the key to understanding temperature-dependent sex determination? J Exp Zool A Ecol Genet Physiol 311A: 314-22.

Lance VA, Elsey RM (1986) Stress-induced suppression of testosterone secretion in male alligators. J Exp Zool 239: 241-6.

Lanciotti RS, Ebel GD, Deubel V, Kerst AJ, Murri S, et al. (2002) Complete genome sequences and phylogenetic analysis of West Nile virus strains isolated from the United States, Europe, and the Middle East. Virology 298: 96–105.

Lande R (1975) The maintenance of genetic variability by mutation in a polygenic character with linked loci. Genet Res 26: 221–35.

Lande R (1980) Sexual dimorphism, sexual selection, and adaptation in polygenic characters. Evolution 34: 292–305.

Lande R (1981) Models of speciation by sexual selection on polygenic traits. Proc Natl Acad Sci USA 78: 3721–5.

Lande R (1994) Risk of population extinction from fixation of new deleterious mutations. Evolution 48: 1460–9.

Lande R (1995) Mutation and conservation. Conserv Biol 9: 782–91.

Lande R, Schemske DW, Schultze ST (1994) High inbreeding depression, selective interference among loci, and the threshold selfing rate for purging recessive lethal mutations. Evolution 48: 965–78.

Lande R, Shannon S (1996) The role of genetic variation in adaptation and population persistence in a changing environment. Evolution 50: 434-7.

Lander ES (2011) Initial impact of the sequencing of the human genome. Nature 470: 187–97.

Lander HM (1997) An essential role for free radicals and derived species in signal transduction. FASEB J 11: 118-24.

Landgren BM, Collins A, Csemiczky G, Burger HG, Baksheev L, Robertson DM (2004) Menopause transition: annual changes in serum hormonal patterns over the menstrual cycle in women during a nine-year period prior to menopause. J Clin Endocrinol Metab 89: 2763–9.

Lando D, Pongratz I, Poellinger L, Whitelaw ML (2000) A redox mechanism controls differential DNA binding activities of hypoxia-inducible factor (HIF) 1alpha and the HIF-like factor. J Biol Chem 275: 4618–27.

Landry C, Garant D, Duchesne P, Bernatchez L (2001) ‘Good genes as heterozygosity’: the major histocompatibility complex and mate choice in Atlantic salmon (Salmo salar). Proc R Soc Lond B 268: 1279-85.

Landry CA, Steele SL, Manning S, Cheek AO (2007) Long term hypoxia suppresses reproductive capacity in the estuarine fish, Fundulus grandis. Comp Biochem Phys A 148: 317–23.

Landry CR, Wittkopp PJ, Taubes CH, Ranz JM, Clark AG, et al. (2005) Compensatory cis-trans evolution and the dysregulation of gene expression in interspecific hybrids of Drosophila. Genetics 171: 1813–22.

Lane DP (1992) Cancer. p53, guardian of the genome. Nature 358: 15–6.

Lane JE, Boutin S, Speakman JR, Humphries MM (2010) Energetic costs of male reproduction in a scramble competition mating system. J Anim Ecol 79: 27-34.

Lane N (2002) Oxygen: the molecule that made the world. Oxford, UK: Oxford University Press.

Lane N (2005) Power, sex, suicide: mitochondria and the meaning of life. Oxford, UK: Oxford University Press.

Lane N (2009) Biodiversity: On the origin of barcodes. Nature 462: 272-4.

Lane N (2011a) The evolution of oxidative stress. In: Pantopoulos K, Schipper HM, eds. Principles of Free Radical Biomedicine. Vol. 1. Hauppauge, NY: Nova Science Publishers, Inc. pp 1-18.

Lane N (2011b) Mitonuclear match: optimizing fitness and fertility over generations drives ageing within generations. Bioessays 33: 860-9.

Lane N (2012) The problem with mixing mitochondria. Cell 151: 246-8.

Lane N, Dean W, Erhardt S, Hajkova P, Surani A, et al. (2003) Resistance of IAPs to methylation reprogramming may provide a mechanism for epigenetic inheritance in the mouse. Genesis 35: 88-93.

Lane TF, Sage EH (1994) The biology of SPARC, a protein that modulates cell-matrix interactions. FASEB J 8: 163-73.

Lang GI, Botstein D, Desai MM (2011) Genetic variation and the fate of beneficial mutations in asexual populations. Genetics 188: 647–61.

Langand J, Jourdane J, Coustau C, Delay B, Morand S (1998) Cost of resistance, expressed as a delayed maturity, detected in the host-parasite system Biomphalaria glabrata/Echinostoma caproni. Heredity 80: 320-5.

Lange C, Zerulla K, Breuert S, Soppa J (2011) Gene conversion results in the equalization of genome copies in the polyploid haloarchaeon Haloferax volcanii. Mol Microbiol 80: 666-77.

Lange IG, Hartel A, Meyer HH (2002) Evolution of oestrogen functions in vertebrates. J Steroid Biochem Mol Biol 83: 219–26.

Langley-Evans SC (2009) Nutritional programming of disease: unravelling the mechanism. J Anat 215: 36-51.

Lankau RA, Strauss SY (2007) Mutual feedbacks maintain both genetic and species diversity in a plant community. Science 317: 1561–3.

Lankenau D-H (2007) Germline double-strand break repair and gene targeting in Drosophila: A trajectory system throughout evolution. Genome Dyn Stab 1: 153-97.

Lankenau D-H, Volff J-N (2009) Transposons and dynamic genome. Berlin, Germany: Springer.

Lanz RB, Wieland S, Hug M, Rusconi S (1995) A transcriptional repressor obtained by alternative translation of a trinucleotide repeat. Nucleic Acids Res 23: 138-45.

Lapinski J, Tunnacliffe A (2003) Anhydrobiosis without trehalose in bdelloid rotifers. FEBS Lett 553: 387-90.

LaPolt PS, Hong LS (1995) Inhibitory effects of superoxide dismutase and cyclic guanosine 3’,5’-monophosphate on estrogen production in cultured rat granulosa cells. Endocrinology 136: 5533-9.

La Porta CA, Gena P, Gritti A, Fascio U, Svelto M, Calamita G (2006) Adult murine CNS stem cells express aquaporin channels. Biol Cell 98: 89-94.

Lara-Ortíz T, Riveros-Rosas H, Aguirre J (2003) Reactive oxygen species generated by microbial NADPH oxidase NoxA regulate sexual development in Aspergillus nidulans. Mol Microbiol 50: 1241-55.

Lardies MA, Bozinovic F (2008) Genetic variation for plasticity in physiological and life-history traits among populations of an invasive species, the terrestrial isopod Porcellio laevis. Evol Ecol Res 10: 747–62.

Larracuente AM, Sackton TB, Greenberg AJ, Wong A, Singh ND, et al. (2008) Evolution of protein-coding genes in Drosophila. Trends Genet 24: 114-23.

Larsson NG, Tulinius MH, Holme E, Oldfors A, Andersen O, et al. (1992) Segregation and manifestations of the mtDNA tRNA(Lys) ARG(8344) mutation of myoclonus epilepsy and ragged-red fibers (MERRF) syndrome. Am J Hum Genet 51: 1201–12.

Larsson P (1991) Intraspecific variability in response to stimuli for male and ephippia formation in Daphnia pulex. Hydrobiologia 225: 281–90.

Lartillot N, Philippe H (2004) A Bayesian mixture model for across-site heterogeneities in the amino-acid replacement process. Mol Biol Evol 21:1095–109.

Laskemoen T, Kleven O, Fossøy F, Robertson RJ, Rudolfsen G, et al. (2010) Sperm quantity and quality effects on fertilization success in a highly promiscuous passerine, the tree swallow Tachycineta bicolor. Behav Ecol Sociobiol 64: 1473–84.

La Spada AR, Wilson EM, Lubahn DB, Harding AE, Fischbeck KH (1991) Androgen receptor gene mutations in X-linked spinal and bulbar muscular atrophy. Nature 352: 77-9.

La Spada AR, Taylor JP (2010) Repeat expansion disease: progress and puzzles in disease pathogenesis. Nat Rev Genet 11: 247-58.

Lass S, Ebert D (2006) Apparent seasonality of parasite dynamics: analysis of cyclic prevalence patterns. Proc R Soc B 273: 199–206.

Latta LC 4th, Bakelar JW, Knapp RA, Pfrender ME (2007) Rapid evolution in response to introduced predators II: the contribution of adaptive plasticity. BMC Evol Biol 7: 21.

Lau NC, Lim LP, Weinstein EG, Bartel, DP (2001) An abundant class of tiny RNAs with probable regulatory roles in Caenorhabditis elegans. Science 294: 858–62.

Lau NC, Seto AG, Kim J, Kuramochi-Miyagawa S, Nakano T, et al. (2006) Characterization of the piRNA complex from rat testes. Science 313: 363–7.

Lau NC, Robine N, Martin R, Chung WJ, Niki Y, et al. (2009) Abundant primary piRNAs, endo-siRNAs, and micro-RNAs in a Drosophila ovary cell line. Genome Res 19: 1776–85.

Laurent LC, Ulitsky I, Slavin I, Tran H, Schork A, et al. (2011) Dynamic changes in the copy number of pluripotency and cell proliferation genes in human ESCs and iPSCs during reprogramming and time in culture. Cell Stem Cell 8: 106–18.

Laurie CC, Chasalow SD, LeDeaux JR, McCarroll R, Bush D, et al. (2004) The genetic architecture of response to long-term artificial selection for oil concentration in the maize kernel. Genetics 168: 2141-55.

Laurila A, Karttunen S, Merilä J (2002) Adaptive phenotypic plasticity and genetics of larval life histories in two Rana temporaria populations. Evolution 56: 617-27.

Laurin M, Gussekloo SWS, Marjanovic D, Legendre L, Cubo J (2012) Testing gradual and speciational models of evolution in extant taxa: the example of ratites. J Evol Biol 25: 293–303.

Lauring AS, Andino R (2010) Quasispecies theory and the behavior of RNA viruses. PLoS Pathog 6: e1001005.

Laval F, Wink DA (1994) Inhibition by nitric oxide of the repair protein, O6-methylguanine-DNA-methyltransferase. Carcinogenesis 15: 443-7.

Laval F, Wink DA, Laval J (1997) A discussion of mechanisms of NO genotoxicity: implication of inhibition of DNA repair proteins. Rev Physiol Biochem Pharmacol 131: 175-91.

Lavoie L, Tremblay A, Mirault ME (1992) Distinct oxidoresistance phenotype of human T47D cells transfected by rat glutathione S-transferase Yc expression vectors. J Biol Chem 267: 3632-6.

Lavon I, Fuchs D, Zrihan D, Efroni G, Zelikovitch B, et al. (2007) Novel mechanism whereby nuclear factor kappaB mediates DNA damage repair through regulation of O(6)-methylguanine-DNA-methyltransferase. Cancer Res 67: 8952–9.

Lavut A, Raveh D (2012) Sequestration of highly expressed mRNAs in cytoplasmic granules, P-bodies, and stress granules enhances cell viability. PLoS Genet 8: e1002527.

Law JA, Jacobsen SE (2010) Establishing, maintaining and modifying DNA methylation patterns in plants and animals. Nat Rev Genet 11: 204-20.

Lawlor LR, Maynard Smith J (1976) The coevolution and stability of competing species. Am Nat 110: 79-99.

Lawrence CW (1982) Mutagenesis in Saccharomyces cerevisiae. Adv Genet 21: 173-254.

Layton JC, Foster PL (2003) Error-prone DNA polymerase IV is controlled by the stress-response sigma factor, RpoS, in Escherichia coli. Mol Microbiol 50: 549–61.

Layton JC, Foster PL (2005) Error-prone DNA polymerase IV is regulated by the heat shock chaperone GroE in Escherichia coli. J Bacteriol 187: 449–57.

Lázár L (2012) The role of oxidative stress in female reproduction and pregnancy. In: Lushchak V, ed. Oxidative stress and diseases. Rijeka, Croatia: InTech. pp 314-336.

Lázaro E, Escarmís C, Pérez-Mercader J, Manrubia SC, Domingo E (2003) Resistance of virus to extinction on bottleneck passages: study of a decaying and fluctuating pattern of fitness loss. Proc Natl Acad Sci USA 100: 10830-5.

Lea DE, Coulson CA (1949) The distribution of the numbers of mutants in bacterial populations. J Genet 49: 264–85.

Leadon SA, Cooper PK (1993) Preferential repair of ionizing radiation-induced damage in the transcribed strand of an active human gene is defective in Cockayne syndrome. Proc Natl Acad Sci USA 90: 10499–503.

Leal MC, Cardoso ER, Nóbrega RH, Batlouni SR, Bogerd J, et al. (2009) Histological and stereological evaluation of zebrafish (Danio rerio) spermatogenesis with an emphasis on spermatogonial generations. Biol Reprod 81: 177–87.

Lebel EG, Masson J, Bogucki A, Paszkowski J (1993) Stress-induced intrachromosomal recombination in plant somatic cells. Proc Natl Acad Sci USA 90: 422-6.

Lebenthal I, Unger R (2010) Computational evidence for functionality of noncoding mouse transcripts. Genomics 96: 10-6.

Le Bras M, Clément MV, Pervaiz S, Brenner C (2005) Reactive oxygen species and the mitochondrial signaling pathway of cell death. Histol Histopathol 20: 205-19.

LeClerc JE, Li B, Payne WL, Cebula TA (1996) High mutation frequencies among Escherichia coli and Salmonella pathogens. Science 274: 1208–11.

LeClerc JE, Payne WL, Kupchella E, Cebula TA (1998) Detection of mutator subpopulations in Salmonella typhimurium LT2 by reversion of his alleles. Mutat Res 400: 89–97.

Leclère L, Jager M, Barreau C, Chang P, Le Guyader H, et al. (2012) Maternally localized germ plasm mRNAs and germ cell/stem cell formation in the cnidarian Clytia. Dev Biol 364: 236-48.

Lécureuil C, Saleh MC, Fontaine I, Baron B, Zakin MM, Guillou F (2004) Transgenic mice as a model to study the regulation of human transferrin expression in Sertoli cells. Hum Reprod 19: 1300-7.

Leduc F, Maquennehan V, Nkoma GB, Boissonneault G (2008) DNA damage response during chromatin remodeling in elongating spermatids of mice. Biol Reprod 78: 324–32.

Leduc F, Faucher D, Bikond Nkoma G, Grégoire M-C, et al. (2011a) Genome-wide mapping of DNA strand breaks. PLoS ONE 6: e17353.

Leduc F, Acteau G, Grégoire MC, Simard O, Leroux J, et al. (2011b) Post-meiotic DNA damage and response in male germ cells. In: Kruman I, ed. DNA repair. New York, NY: InTech North America.

Lee CE, Remfert JL, Gelembiuk GW (2003) Evolution of physiological tolerance and performance during freshwater invasions. Integr Comp Biol 43: 439–49.

Lee FS, Kim AH, Khursigara G, Chao MV (2001) The uniqueness of being a neurotrophin receptor. Curr Opin Neurobiol 11: 281-6.

Lee FY, Li Y, Zhu H, Yang S, Lin HZ, et al. (1999) Tumor necrosis factor increases mitochondrial oxidant production and induces expression of uncoupling protein-2 in the regenerating mice [correction of rat] liver. Hepatology 29: 677-87.

Lee HC, Gu W, Shirayama M, Youngman E, Conte D Jr, Mello CC (2012) C. elegans piRNAs mediate the genome-wide surveillance of germline transcripts. Cell 150: 78–87.

Lee HS, Chen ZJ (2001) Protein-coding genes are epigenetically regulated in Arabidopsis polyploids. Proc Natl Acad Sci USA 98: 6753–8.

Lee J, Richburg JH, Younkin SC, Boekelheide K (1997) The Fas system is a key regulator of germ cell apoptosis in the testis. Endocrinology 138: 2081–8.

Lee J, Bruce-Keller AJ, Kruman Y, Chan SL, Mattson MP (1999a) 2-Deoxy-D-glucose protects hippocampal neurons against excytotoxic and oxidative injury: Evidence for the involvement of stress proteins. J Neurosci Res 57: 48-61.

Lee J, Richburg JH, Shipp EB, Meistrich ML, Boekelheide K (1999b) The Fas system, a regulator of testicular germ cell apoptosis, is differentially up-regulated in Sertoli cell versus germ cell injury of the testis. Endocrinology 140: 852-8.

Lee J, Inoue K, Ono R, Ogonuki N, Kohda T, et al. (2002) Erasing genomic imprinting memory in mouse clone embryos produced from day 11.5 primordial germ cells. Development 129: 1807–17.

Lee JJ, Kim BC, Park MJ, Lee YS, Kim YN, et al. (2011) PTEN status switches cell fate between premature senescence and apoptosis in glioma exposed to ionizing radiation. Cell Death Differ 18: 666-77.

Lee JY, Yao TP (2010) Quality control autophagy: A joint effort of ubiquitin, protein deacetylase and actin cytoskeleton. Autophagy 6: 555-7.

Lee JY, Koga H, Kawaguchi Y, Tang W, Wong E, et al. (2010) HDAC6 controls autophagosome maturation essential for ubiquitin-selective quality-control autophagy. EMBO J 29: 969-80.

Lee JT (2012) Epigenetic regulation by long noncoding RNAs. Science 338: 1435-9.

Lee K, Park JS, Kim YJ, Soo Lee YS, Sook Hwang TS, et al. (2002) Differential expression of Prx I and II in mouse testis and their up-regulation by radiation. Biochem Biophys Res Commun 296: 337–42.

Lee MC, Marx CJ (2013) Synchronous waves of failed soft sweeps in the laboratory: remarkably rampant clonal interference of alleles at a single locus. Genetics 193: 943-52.

Lee NP, Cheng CY (2004) Nitric oxide/nitric oxide synthase, spermatogenesis, and tight junction dynamics. Biol Reprod 70: 267–76.

Lee RC, Feinbaum RL, Ambros V (1993) The C. elegans heterochronic gene lin-4 encodes small RNAs with antisense complementarity to lin-14. Cell 75: 843-854.

Lee RC, Ambros V (2001) An extensive class of small RNAs in Caenorhabditis elegans. Science 294: 862–4.

Lee SB, Bae IH, Bae YS, Um HD (2006) Link between mitochondria and NADPH oxidase 1 isozyme for the sustained production of reactive oxygen species and cell death. J Biol Chem 281: 36228–35.

Lee SC, Ni M, Li W, Shertz C, Heitman J (2010) The evolution of sex: a perspective from the fungal kingdom. Microbiol Mol Biol Rev 74: 298-340.

Lee SH, Chang DK, Goel A, Boland CR, Bugbee W, et al. (2003) Microsatellite instability and suppressed DNA repair enzyme expression in rheumatoid arthritis. J Immunol 170: 2214-20.

Lee SM, Huh TL, Park JW (2001) Inactivation of NADP(+)−dependent isocitrate dehydrogenase by reactive oxygen species. Biochimie 83: 1057–65.

Lee SR, Kwon KS, Kim SR, Rhee SG (1998) Reversible inactivation of protein-tyrosine phosphatase 1B in A431 cells stimulated with epidermal growth factor. J Biol Chem 273: 15366-72.

Lee SS, Lee RY, Fraser AG, Kamath RS, Ahringer J, Ruvkun G (2003) A systematic RNAi screen identifies a critical role for mitochondria in C. elegans longevity. Nat Genet 33: 40–8.

Lee TD (1984) Patterns of fruit maturation: a gametophyte competition hypothesis. Am Nat 123: 427–8.

Lee WK, Bork U, Gholamrezaei F, Thévenod (2005) Cd2+-induced cytochrome c release in apoptotic proximal tubule cells: role of mitochondrial permeability transition pore and Ca2+ uniporter. Am J Physiol Renal Physiol 288: F27–F39.

Lee WK, Thévenod F (2006) A role for mitochondrial aquaporins in cellular life-and-death decisions? Am J Physiol Cell Physiol 291: C195–C202.

Lee YH, Ota T, Vacquier VD (1995) Positive selection is a general phenomenon in the evolution of abalone sperm lysin. Mol Biol Evol 12: 231–8.

Lee YM, Lee JY, Kim MJ, Bae HI, Park JY, et al. (2006) Hypomethylation of the protein gene product 9.5 promoter region in gallbladder cancer and its relationship with clinicopathological features. Cancer Sci 97: 1205–10.

Leedom L, Lewis C, Garcia-Segura LM, Naftolin F (1994) Regulation of arcuate nucleus synaptology by estrogen. Ann NY Acad Sci 743: 61-71.

Leeflang EP, Tavare S, Marjoram P, Neal CO, Srinidhi J, et al. (1999) Analysis of germline mutation spectra at the Huntington's disease locus supports a mitotic mutation mechanism. Hum Mol Genet 8: 173–83.

Leem YE, Ripmaster TL, Kelly FD, Ebina H, Heincelman ME, et al. (2008) Retrotransposon Tf1 is targeted to Pol II promoters by transcription activators. Mol Cell 30: 98-107.

Lees-Murdock DJ, Walsh CP (2008) DNA methylation reprogramming in the germ line. Epigenetics 3: 5–13.

Leese HJ (2002) Quiet please, do not disturb: a hypothesis of embryo metabolism and viability. Bioessays 24: 845–9.

Leeton PR, Smyth DR (1993) An abundant LINE-like element amplified in the genome of Lilium speciosum. Mol Gen Genet 237: 97-104.

Leffler EM, Bullaughey K, Matute DR, Meyer WK, Ségurel L, et al. (2012) Revisiting an old riddle: What determines genetic diversity levels within species? PLoS Biol 10: e1001388.

Legan SJ, Callahan WH (1999) Suppression of tonic luteinizing hormone secretion and norepinephrine release near the GnRH neurons by estradiol in ovariectomized rats. Neuroendocrinology 70: 237-45.

Legeai F, Rizk G, Walsh T, Edwards O, Gordon K, et al. (2010) Bioinformatic prediction, deep sequencing of microRNAs and their role in phenotypic plasticity in the pea aphid, Acyrthosiphon pisum. BMC Genomics 11: 281.

Legendre M, Pochet N, Pak T, Verstrepen KJ (2007) Sequence-based estimation of minisatellite and microsatellite repeat variability. Genome Res 17: 1787-96.

Legendre-Guillemin V, Wasiak S, Hussain NK, Angers A, McPherson PS (2004) ENTH/ANTH proteins and clathrin-mediated membrane budding. J Cell Sci 117: 9–18.

Le Goff P, Michel D (1999) HSF1 activation occurs at different temperatures in somatic and male germ cells in the poikilotherm rainbow trout. Biochem Biophys Res Comm 259: 15-20.

Le Grande CE (1970) Radioprotectors and anticarcinogens. In: Johnson AD, Gomes WR, Vandemark NL, eds. The Testis, vol. 3. New York, NY: Academic Press. pp 333-365.

Le Hir M, Bluethmann H, Kosco-Vilbois MH, Müller M, di Padova F, et al. (1996) Differentiation of follicular dendritic cells and full antibody responses require tumor necrosis factor receptor-1 signaling. J Exp Med 183: 2367-72.

Lehner B (2010) Genes confer similar robustness to environmental, stochastic, and genetic perturbations in yeast. PLoS ONE 5: e9035.

Lehner B (2011) Molecular mechanisms of epistasis within and between genes. Trends Genet 27: 323–31.

Lehner B, Kaneko K (2011) Fluctuation and response in biology. Cell Mol Life Sci 68: 1005-10.

Lehman N (2003) A case for the extreme antiquity of recombination. J Mol Evol 56: 770–7.

Lei L, Jin S, Gonzalez G, Behringer RR, Woodruff TK (2010) The regulatory role of Dicer in folliculogenesis in mice. Mol Cell Endocrinol 315: 63-73.

Leichtmann-Bardoogo Y, Cohen LA, Weiss A, Marohn B, Schubert S, et al. (2012) Compartmentalization and regulation of iron metabolism proteins protect male germ cells from iron overload. Am J Physiol Endocrinol Metab 302: E1519-30.

Leiderman B, Mancini RE (1968) Aerobic and anaerobic lactate production in the prepuberal and adult rat testis. Proc Soc Exp Biol Med 128: 818-21.

Leigh EG (1970) Natural selection and mutability. Am Nat 104: 301–5.

Leigh EG (1973) The evolution of mutation rates. Genetics 73(Suppl.): 1-18.

Leigh EG Jr (1977) How does selection reconcile individual advantage with the good of the group? Proc Natl Acad Sci USA 74: 4542–6.

Leigh EG Jr (1999) Levels of selection, potential conflicts, and their resolution – the role of the ‘‘Common Good’’. In: Keller L, ed. Levels of selection in evolution. Princeton, NJ: Princeton University Press. pp 15–30.

Leigh EG Jr (2010) The evolution of mutualism. J Evol Biol 23: 2507-28.

Leigh EG Jr (2010) The group selection controversy. J Evol Biol 23: 6–19.

Leigh EG Jr, Rowell TE (1995) The evolution of mutualism and other forms of harmony at various levels of biological organization. Ecologie 26: 131–58.

Leigh Simpson J (1999) Genetic causes of spontaneous abortion. In: Queenan JT, ed. Management of high-risk pregnancy. Malden, MA: Blackwell Science. pp 351-6.

Leimu R, Mutikainen P, Koricheva J, Fischer M (2006) How general are positive relationships between plant population size, fitness and genetic variation? J Ecol 94: 942–52.

Leitch AR, Leitch IJ (2008) Genome plasticity and the diversity of polyploid plants. Science 320: 481–3.

Le Mevel JC, Abitibol S, Beraud G, Maniey J (1978) Dynamic changes in plasma adrenocorticotrophin after neurotropic stress in male and female rats. J Endocrinol 76: 359-60.

Le Mevel JC, Abitibol S, Beraud G, Maniey J (1979) Temporal changes in plasma adrenocorticotropin concentration after repeated neurotropic stress in male and female rats. Endocrinology 105: 812-7.

Lengner CJ, Camargo FD, Hochedlinger K, Welstead GG, Zaidi S, et al. (2007) Oct4 expression is not required for mouse somatic stem cell self-renewal. Cell Stem Cell 1: 403–15.

Lennon JT, Jones SE (2011) Microbial seed banks: the ecological and evolutionary implications of dormancy. Nat Rev Microbiol 9: 119-30.

Lenormand T, Dutheil J (2005) Recombination difference between sexes: A role for haploid selection. PLoS Biol 3: e63.

Lenormand T, Roze D, Rousset F (2009) Stochasticity in evolution. Trends Ecol Evol 24: 157-65.

Lenski RE, Mittler JE (1993) The directed mutation controversy and neo-Darwinism. Science 259: 188–194.

Lenski RE, Travisano M (1994) Dynamics of adaptation and diversification: a 10,000-generation experiment with bacterial populations. Proc Natl Acad Sci USA 91: 6808–14.

Lenski RE, Ofria C, Pennock RT, Adami C (2003) The evolutionary origin of complex features. Nature 423: 139–44.

Lenski RE, Barrick JE, Ofria C (2006) Balancing robustness and evolvability. PLoS Biol 4: e428.

Lenzi A, Gandini L, Picardo M, Tramer F, Sandri G, Panfili E (2000) Lipoperoxidation damage of spermatozoa polyunsaturated fatty acids (PUFA): scavenger mechanisms and possible scavenger therapies. Front Biosci 5: E1-E15.

Leon J, Acuna-Castroviejo D, Escames G, Tan DX, Reiter RJ (2005) Melatonin mitigates mitochondrial malfunction. J Pineal Res 38: 1–9.

Lepikhov K, Walter J (2004) Differential dynamics of histone H3 methylation at positions K4 and K9 in the mouse zygote. BMC Dev Biol 4: 12–6.

Lequarré AS, Grisart B, Moreau B, Schuurbiers N, Massip A, Dessy F (1997) Glucose metabolism during bovine preimplantation development: analysis of gene expression in single oocytes and embryos. Mol Reprod Dev 48: 216–26.

Lercher MJ, Hurst LD (2002) Human SNP variability and mutation rate are higher in regions of high recombination. Trends Genet 18: 337–40.

Le Roux JJ, Wieczorek AM, Wright MG, Tran CT (2007) Super-genotype: global monoclonality defies the odds of nature. PLoS ONE 2: e590.

Le Rouzic A, Carlborg Ö (2008) Evolutionary potential of hidden genetic variation. Trends Ecol Evol 23: 33–7.

Lertratanangkoon K, Wu CJ, Savaraj N, Thomas ML (1997) Alterations of DNA methylation by glutathione depletion. Cancer Lett 120: 149-56.

Lescoat G, Jego P, Beraud B, Maniey J (1970) Influence de sexe sur les modalités de réponse de l’axe hypothalamo-hypophyso-surrénalien aux agressions emotionelles et somatiques chez le rat. C R Seances Sot Biol Fil 164: 2106-13.

Lesica P, Young TP (2005) A demographic model explains life-history variation in Arabis fecunda. Funct Ecol 19: 471–7.

Lessells CM (1986) Brood size in Canada Geese: a manipulation experiment. J Anim Ecol 55: 669-89.

Lessells CM (1999) Sexual conflict in animals. In: Keller L, ed. Levels of Selection in Evolution. Princeton, NJ: Princeton University Press. pp 75–99.

Lessells CM (2006) The evolutionary outcome of sexual conflict. Philos Trans R Soc Lond B Biol Sci 361: 301-17.

Lester NP, Shuter BJ, Abrams PA (2004) Interpreting the von Bertalanffy model of somatic growth in fishes: the cost of reproduction. Proc Biol Sci 271: 1625-31.

Le Trionnaire G, Hardie J, Jaubert-Possamai S, Simon JC, Tagu D (2008) Shifting from clonal to sexual reproduction in aphids: physiological and developmental aspects. Biol Cell 100: 441–51.

Lettre G, Kritikou EA, Jaeggi M, Calixto A, Fraser AG, et al. (2004) Genome-wide RNAi identifies p53-dependent and -independent regulators of germ cell apoptosis in C. elegans. Cell Death Differ 11: 1198–203.

Leung AK, Sharp PA (2010) MicroRNA functions in stress responses. Mol Cell 40: 205-15.

Leung DC, Lorincz MC (2012) Silencing of endogenous retroviruses: when and why do histone marks predominate? Trends Biochem Sci 37: 127-33.

Leung TLF, King KC, Wolinska J (2012) Escape from the Red Queen: an overlooked scenario in coevolutionary studies. Oikos 121: 641–5.

Levadoux-Martin M, Gouble A, Jegou B, Vallet-Erdtmann V, Auriol J, et al. (2003) Impaired gametogenesis in mice that overexpress the RNA-binding protein HuR. EMBO Rep 4: 394-399.

Levasseur A, Orlando L, Bailly X, Milinkovitch MC, Danchin EG, Pontarotti P (2007) Conceptual bases for quantifying the role of the environment on gene evolution: the participation of positive selection and neutral evolution. Biol Rev Camb Philos Soc 82: 551–72.

Levin BR (1988) The evolution of sex in bacteria. In: Michod RE, Levin BR, eds. The Evolution of Sex: An Examination of Current Ideas. Sunderland, MA: Sinauer Associates. pp 194–211.

Levin DA (1975) Pest pressure and recombination systems in plants. Am Nat 109: 437–51.

Levin DA (1983) Polyploidy and novelty in flowering plants. Am Nat 122: 1–25.

Levin DA (1988) Local differentiation and the breeding structure of plant populations. In: Gottlieb LD, Jain SK, eds. Plant Evolutionary Biology. New York, NY: Chapman & Hall. pp 305–329.

Levin SA, Hastings A, Cohen D (1984) Dispersal strategies in patchy environments. Theor Popul Biol 26: 165-91.

Levine AJ, Tomasini R, McKeon FD, Mak TW, Melino G (2011) The p53 family: guardians of maternal reproduction. Nat Rev Mol Cell Biol 12: 259-65.

Levins R (1968) Evolution in changing environments. Princeton, NJ: Princeton University Press.

Levins R (1969) The effect of random variations of different types on population growth. Proc Natl Acad Sci USA 62: 1061-5.

Levinson G, Gutman GA (1987a) High frequencies of short frameshifts in poly-CA/TG tandem repeats borne by bacteriophage M13 in Escherichia coli K-12. Nucleic Acids Res 15: 5323–38.

Levinson G, Gutman GA (1987b) Slipped-strand mispairing - a major mechanism for DNA-sequence evolution. Mol Biol Evol 4: 203–21.

Levy AP, Levy NS, Wegner S, Goldberg MA (1995) Transcriptional regulation of the rat vascular endothelial growth factor gene by hypoxia. J Biol Chem 270: 13333–40.

Levy AP, Levy NS, Goldberg MA (1996) Posttranscriptional regulation of vascular endothelial growth factor by hypoxia. J Biol Chem 271: 2746–53.

Levy SF, Ziv N, Siegal ML (2012) Bet hedging in yeast by heterogeneous, age-correlated expression of a stress protectant. PLoS Biol 10: e1001325.

Lewens T (2009) Seven types of adaptationism. Biol Philos 24: 161-82.

Lewin R (1980) Evolutionary theory under fire. Science 210: 883–7.

Lewis BP, Burge CB, Bartel DP (2005) Conserved seed pairing, often flanked by adenosines, indicates that thousands of human genes are microRNA targets. Cell 120: 15–20.

Lewis WH (1980) Polyploidy: biological relevance. New York, NY: Plenum Press.

Lewis WM (1987) The cost of sex. EXS 55: 33-57.

Lewontin R (1961) Evolution and the theory of games. J Theor Biol 1: 382-403.

Lewontin RC (1966) Is nature probable or capricious? BioScience 16: 25–7.

Lewontin R (1970) The units of selection. Annu Rev Ecol Syst 1: 1–14.

Lewontin RC (1974) The genetic basis of evolutionary change. Cambridge, MA: Harvard University Press.

Lewontin RC (1983) Biological determinism. Tanner Lectures on Human Values 4: 149-83. Salt Lake City, UT: University of Utah Press.

Lewontin RC, Cohen D (1969) On population growth in a randomly varying environment. Proc Natl Acad Sci USA 62: 1056-60.

Leyens G, Knoops B, Donnay I (2004a) Expression of peroxiredoxins in bovine oocytes and embryos produced in vitro. Mol Reprod Dev 69: 243-51.

Leyens G, Verhaeghe B, Landtmeters M, Marchandise J, Knoops B, Donnay I (2004b) Peroxiredoxin 6 is upregulated in bovine oocytes and cumulus cells during in vitro maturation: role of intercellular communication. Biol Reprod 71: 1646-51.

Lezberg AL, Halpern CB, Antos JA (2001) Clonal development of Maianthemum dilatatum in forests of differing age and structure. Can J Bot 79: 1028-38.

Li C, Vagin VV, Lee S, Xu J, Ma S, et al. (2009) Collapse of germline piRNAs in the absence of Argonaute3 reveals somatic piRNAs in flies. Cell 137: 509-21.

Li E (2002) Chromatin modification and epigenetic reprogramming in mammalian development. Nat Rev Genet 3: 662–73.

Li E, Bestor TH, Jaenisch R (1992) Targeted mutation of the DNA methyltransferase gene results in embryonic lethality. Cell 69: 915–26.

Li H, Roossinck MJ (2004) Genetic bottlenecks reduce population variation in an experimental RNA virus population. J Virol 78: 10582-7.

Li H, Coghlan A, Ruan J, Coin LJ, Heriche JK, et al. (2006) TreeFam: a curated database of phylogenetic trees of animal gene families. Nucleic Acids Res 34: D572–D580.

Li H, Lu Y (2009) Phenotypic inheritance induced by hairpin RNA in Drosophila. Acta Biochim Biophys Sin (Shanghai) 41: 922-8.

Li J, Li Q, Xie C, Zhou H, Wang Y, et al. (2004) Beta-actin is required for mitochondria clustering and ROS generation in TNF-induced, caspase-independent cell death. J Cell Sci 117: 4673-80.

Li JY, Lees-Murdock DJ, Xu GL, Walsh CP (2004) Timing of establishment of paternal methylation imprints in the mouse. Genomics 84: 952–60.

Li L, Oppenheim RW, Milligan CE (2001) Characterization of the execution pathway of developing motoneurons deprived of trophic support. J Neurobiol 46: 249–64.

Li LH, Wine RN, Chapin RE (1996) 2-Methoxyacetic acid (MAA)-induced spermatocyte apoptosis in human and rat testes: an in vitro comparison. J Androl 17: 538–49.

Li LY, Seddon AP, Meister A, Risley MS (1989) Spermatogenic cell-somatic cell interactions are required for maintenance of spermatogenic cell glutathione. Biol Reprod 40: 317-31.

Li LY, Luo X, Wang X (2001) Endonuclease G is an apoptotic DNase when released from mitochondria. Nature 412: 95–9.

Li M, Wang IX, Li Y, Bruzel A, Richards AL, et al. (2011) Widespread RNA and DNA sequence differences in the human transcriptome. Science 333: 53-8.

Li M, Vascotto C, Xu S, Dai N, Qing Y, et al. (2012) Human AP endonuclease/redox factor APE1/ref-1 modulates mitochondrial function after oxidative stress by regulating the transcriptional activity of NRF1. Free Radic Biol Med 53: 237-48.

Li N, Karin M (1999) Is NF-kappaB the sensor of oxidative stress? Faseb J 13: 1137-43.

Li N, Lv J, Niu DK (2009) Low contents of carbon and nitrogen in highly abundant proteins: evidence of selection for the economy of atomic composition. J Mol Evol 68: 248-55.

Li S, Zhou W, Doglio L, Goldberg E (1998) Transgenic mice demonstrate a testis-specific promoter for lactate dehydrogenase, LDHC. J Biol Chem 273: 31191–4.

Li W, Baker NE (2007) Engulfment is required for cell competition. Cell 15: 1215–25.

Li WH, Wu CI, Luo CC (1984) Nonrandomness of point mutation as reflected in nucleotide substitutions in pseudogenes and its evolutionary implications. J Mol Evol 21: 58-71.

Li WH, Ellsworth DL, Krushkal J, Chang BH, Hewett-Emmett D (1996) Rates of nucleotide substitution in primates and rodents and the generation-time effect hypothesis. Mol Phylogenet Evol 5: 182-7.

Li WH, Yi SJ, Makova K (2002) Male-driven evolution. Curr Opin Genet Dev 12: 650–6.

Li X, Youngblood GL, Payne AH, Hales DB (1995) Tumor necrosis factor-α inhibition of 17α-hydroxylase/C17–20 lyase gene (Cyp17) expression. Endocrinology 136: 3519–26.

Li YC, Korol AB, Fahima T, Beiles A, Nevo E (2002) Microsatellites: genomic distribution, putative functions and mutational mechanisms: a review. Mol Ecol 11: 2453-65.

Li YC, Korol AB, Fahima T, Nevo E (2004) Microsatellites within genes: structure, function, and evolution. Mol Biol Evol 21: 991-1007.

Li Z, Yang J, Huang H (2006) Oxidative stress induces H2AX phosphorylation in human spermatozoa. FEBS Lett 580: 6161–8.

Liancourt P, Callaway RM, Michalet R (2005) Stress tolerance and competitive response ability determine the outcome of biotic interactions. Ecology 86: 1611–8.

Liang F, Han M, Romanienko PJ, Jasin M (1998) Homology-directed repair is a major double-strand break repair pathway in mammalian cells. Proc Natl Acad Sci USA 95: 5172–7.

Liang G, Chan MF, Tomigahara Y, Tsai YC, Gonzales FA, et al. (2002) Cooperativity between DNA methyltransferases in the maintenance methylation of repetitive elements. Mol Cell Biol 22: 480-91.

Liang Y, Liu J, Feng Z (2013a) The regulation of cellular metabolism by tumor suppressor p53. Cell Biosci 3: 9.

Liang Y, Zhang H, Feng QS, Cai MB, Deng W, et al. (2013b) The propensity for tumorigenesis in human induced pluripotent stem cells is correlated with genomic instability. Chin J Cancer 32: 205-12.

Liao BY, Zhang J (2006) Low rates of expression profile divergence in highly expressed genes and tissue-specific genes during mammalian evolution. Mol Biol Evol 23: 1119-28.

Liao BY, Scott NM, Zhang J (2006) Impacts of gene essentiality, expression pattern, and gene compactness on the evolutionary rate of mammalian proteins. Mol Biol Evol 23: 2072-80.

Liao GC, Rehm EJ, Rubin GM (2000) Insertion site preferences of the P transposable element in Drosophila melanogaster. Proc Natl Acad Sci USA 97: 3347-51.

Liao W, Cai M, Chen J, Huang J, Liu F, Jiang C, Gao Y (2010) Hypobaric hypoxia causes deleterious effects on spermatogenesis in rats. Reproduction 139: 1031-8.

Liberman U, Feldman MW (1986) Modifiers of mutation rate: a general reduction principle. Theor Popul Biol 30: 125–42.

Licht P (1965) The relation between preferred body temperatures and testicular heat sensitivity in lizards. Copeia 1965: 428-36.

Licht P, McCreery BR, Barnes R, Pang P (1983) Seasonal and stress related changes in plasma gonadotropins, sex steroids, and corticosterone in the bullfrog, Rana catesbeiana. Gen Comp Endocrinol 50: 124-45.

Lichten M, Haber JE (1989) Position effects in ectopic and allelic mitotic recombination in Saccharomyces cerevisiae. Genetics 123: 261-8.

Lidzbarsky GA, Shkolnik T, Nevo E (2009) Adaptive response to DNA-damaging agents in natural Saccharomyces cerevisiae populations from ‘‘Evolution Canyon’’, Mt. Carmel, Israel. PLoS ONE 4: e5914.

Lidstrom ME, Konopka MC (2010) The role of physiological heterogeneity in microbial population behavior. Nat Chem Biol 6: 705-12.

Lieber MR (2010) The mechanism of double-strand DNA break repair by the nonhomologous DNA end-joining pathway. Annu Rev Biochem 79: 181–211.

Lieberman BS, Brett CE, Eldredge N (1995) A study of stasis and change in two species lineages from the Middle Devonian of New York State. Paleobiology 21: 15–27.

Ligorio M, Izzotti A, Pulliero A, Arrigo P (2011) Mutagens interfere with microRNA maturation by inhibiting DICER. An in silico biology analysis. Mutat Res 717: 116-28.

Lilienbaum A, Sage J, Mémet S, Rassoulzadegan M, Cuzin F, Israël A (2000) NF-kappa B is developmentally regulated during spermatogenesis in mice. Dev Dyn 219: 333-40.

Lillycrop KA, Phillips ES, Jackson AA, Hanson MA, Burdge GC (2005) Dietary protein restriction of pregnant rats induces and folic acid supplementation prevents epigenetic modification of hepatic gene expression in the offspring. J Nutr 135: 1382-6.

Lim AK, Kai T (2007) Unique germ-line organelle, nuage, functions to repress selfish genetic elements in Drosophila melanogaster. Proc Natl Acad Sci USA 104: 6714–9.

Lim HN, van Oudenaarden A (2007) A multistep epigenetic switch enables the stable inheritance of DNA methylation states. Nat Genet 39: 269–75.

Lim J, Luderer U (2011) Oxidative damage increases and antioxidant gene expression decreases with aging in the mouse ovary. Biol Reprod 84: 775-82.

Lim JA, Kim SH (2009) Transcriptional activation of an anti-oxidant mouse Pon2 gene by dexamethasone. BMB Rep 42: 421-6.

Lim JP, Brunet A (2013) Bridging the transgenerational gap with epigenetic memory. Trends Genet 29: 176-86.

Lim SO, Gu JM, Kim MS, Kim HS, Park YN, et al. (2008) Epigenetic changes induced by reactive oxygen species in hepatocellular carcinoma: methylation of the E-cadherin promoter. Gastroenterology 135: 2128–40.

Limbourg T, Mateman AC, Andersson S, Lessells CM (2004) Female blue tits adjust parental effort to manipulated male UV attractiveness. Proc R Soc Lond B 271: 1903–8.

Limoli CL, Hartmann A, Shephard L, Yang CR, Boothman DA, et al. (1998) Apoptosis, reproductive failure, and oxidative stress in Chinese hamster ovary cells with compromised genomic integrity. Cancer Res 58: 3712-8.

Lin CY, Lovén J, Rahl PB, Paranal RM, Burge CB, et al. (2012) Transcriptional amplification in tumor cells withelevated c-Myc. Cell 151: 56-67.

Lin H, Decuypere E, Buyse J (2004) Oxidative stress induced by corticosterone administration in broiler chickens (Gallus gallus domesticus) 2. Short-term effect. Comp Biochem Physiol Part B Biochem Mol Biol 139: 745–51.

Lin H, Yin H (2008) A novel epigenetic mechanism in Drosophila somatic cells mediated by Piwi and piRNAs. Cold Spring Harb Symp Quant Biol 73: 273-81.

Lin H, Wang Y, Wang Y, Tian F, Pu P, et al. (2010) Coordinated regulation of active and repressive histone methylations by a dual-specificity histone demethylase ceKDM7A from Caenorhabditis elegans. Cell Res 20: 899-907.

Lin R, Ding L, Casola C, Ripoll DR, Feschotte C, Wang H (2007) Transposase-derived transcription factors regulate light signaling in Arabidopsis. Science 318: 1302–5.

Lin SR, Hsieh SC, Yueh YY, Lin TH, Chao DY, et al. (2004) Study of sequence variation of dengue type 3 virus in naturally infected mosquitoes and human hosts: Implications for transmission and evolution. J Virol 78: 12717–21.

Lin T, Wang D, Nagpal ML, Calkins JH, Chang WC, Chi R (1991) Interleukin-1 inhibits cholesterol side-chain cleavage cytochrome P450 expression in primary cultures of Leydig cells. Endocrinology 129: 1305-11.

Lin T, Wang D, Nagpal ML (1993) Human chorionic gonadotropin induces interleukin-1 gene expression in rat Leydig cells in vivo. Mol Cell Endocrinol 95: 139–45.

Lin X, Ramamurthi K, Mishima M, Kondo A, Howell SB (2000) p53 interacts with the DNA mismatch repair system to modulate the cytotoxicity and mutagenicity of hydrogen peroxide. Mol Pharmacol 58: 1222–9.

Lin X, Ruan X, Anderson MG, McDowell JA, Kroeger PE, et al. (2005) siRNA-mediated off-target gene silencing triggered by a 7 nt complementation. Nucleic Acids Res 33: 4527–35.

Lin Y, Dion V, Wilson JH (2006) Transcription promotes contraction of CAG repeat tracts in human cells. Nat Struct Mol Biol 13: 179-80.

Lin Y, Wilson JH (2007) Transcription-induced CAG repeat contraction in human cells is mediated in part by transcription-coupled nucleotide excision repair. Mol Cell Biol 27: 6209-17.

Lin Y, Hubert L Jr, Wilson JH (2009) Transcription destabilizes triplet repeats. Mol Carcinog 48: 350-61.

Lin Y, Dent SY, Wilson JH, Wells RD, Napierala M (2010) R loops stimulate genetic instability of CTG•CAG repeats. Proc Natl Acad Sci USA 107: 692-7.

Lin YL, Pasero P (2012) Interference between DNA replication and transcription as a cause of genomic instability. Curr Genomics 13: 65-73.

Lincoln GA (2002) Neuroendocrine regulation of seasonal gonadotrophin and prolactin rhythms: lessons from the Soay ram model. Reprod Suppl 59: 131–47.

Lindahl T (1993) Instability and decay of the primary structure of the DNA. Nature 362: 709–15.

Lindahl T, Nyberg B (1972) Rate of depurination of native deoxyribonucleic acid. Biochemistry 11: 3610–8.

Lindahl T, Nyberg B (1974) Heat-induced deamination of cytosine residues in deoxyribonucleic acid. Biochemistry 13: 3405–10.

Lindahl T, Barnes DE (2000) Repair of endogenous DNA damage. Cold Spring Harb Symp Quant Biol 65: 127-33.

Linder B, Plöttner O, Kroiss M, Hartmann E, Laggerbauer B, et al. (2008) Tdrd3 is a novel stress granule-associated protein interacting with the Fragile-X syndrome protein FMRP. Hum Mol Genet 17: 3236–46.

Linder P (2006) Dead-box proteins: a family affair–active and passive players in RNP-remodeling. Nucleic Acids Res 34: 4168–80.

Linder P, Lasko P (2006) Bent out of shape: RNA unwinding by the DEAD-box helicase Vasa. Cell 125: 219–21.

Lindgren D (1972a) The temperature influence on the spontaneous mutation rate. I. Literature review. Hereditas 70: 165-78.

Lindgren D (1972b) The temperature influence on the spontaneous mutation rate. II. Investigation by the aid of waxy mutants. Hereditas 70: 179-84.

Lindow M, Gorodkin J (2007) Principles and limitations of computational microRNA gene and target finding. DNA Cell Biol 26: 339-351.

Lindquist S (1986) The heat-shock response. Annu Rev Biochem 55: 1151-91.

Lindquist S, Craig EA (1988) The heat-shock proteins. Annu Rev Genet 22: 631−77.

Lindsley DL, Tokuyasu KT (1980) Spermatogenesis. In: Ashburner M, Wright TRF, eds. The genetics and biology of Drosophila, vol. 2a. New York, NY: Academic. pp 225–294.

Lindström J (1999) Early development and fitness in birds and mammals. Trends Ecol Evol 14: 343-8.

Linsen SE, de Wit E, de Bruijn E, Cuppen E (2010) Small RNA expression and strain specificity in the rat. BMC Genomics 11: 249.

Liou GY, Storz P (2010) Reactive oxygen species in cancer. Free Radic Res 44: 479–96.

Lipinski P, Drapier JC, Oliveira L, Retmanska H, Sochanowicz B, Kruszewski M (2000) Intracellular iron status as a hallmark of mammalian cell susceptibility to oxidative stress: a study of L5178Y mouse lymphoma cell lines differentially sensitive to H2O2. Blood 95: 2960-6.

Lipow SR, Wyatt R (2000) Single gene control of postzygotic self-incompatibility in poke milkweed, Asclepias exaltata L. Genetics 154:893–907.

Lippert MJ, Kim N, Cho JE, Larson RP, Schoenly NE, et al. (2011) Role for topoisomerase 1 in transcription-associated mutagenesis in yeast. Proc Natl Acad Sci USA 108: 698-703.

Lippman Z, May B, Yordan C, Singer T, Martienssen R (2003) Distinct mechanisms determine transposon inheritance and methylation via small interfering RNA and histone modification. PLoS Biol 1: E67.

Lippman Z, Gendrel AV, Black M, Vaughn MW, Dedhia N, et al. (2004) Role of transposable elements in heterochromatin and epigenetic control. Nature 430: 471-6.

Lips K (2001) Reproductive trade-offs and bet-hedging in Hyla calypsa, a Neotropical treefrog. Oecologia 128: 509–18.

Lipson H, Pollack JB, Suh NP (2002) On the origin of modular variation. Evolution 56: 1549–56.

Lisch D, Bennetzen JL (2011) Transposable element origins of epigenetic gene regulation. Curr Opin Plant Biol 14: 156-61.

Lissbrant E, Collin O, Bergh A (1997a) Localization and effects of calcitonin gene-related peptide in the testicular vasculature of the rat. J Androl 18: 385–92.

Lissbrant E, Löfmark U, Collin O, Bergh A (1997b) Is nitric oxide involved in the regulation of the rat testicular vasculature? Biol Reprod 56: 1221–7.

Lissbrant E, Collin O, Damber JE, Bergh A (2006) Effects of haemorrhagic hypotension on the subcapsular artery and microvasculature of the rat testis. Int J Androl 29: 434-40.

Lister A, Van Der Kraak G (2002) Modulation of goldfish testicular testosterone production in vitro by tumor necrosis factor alpha, interleukin-1beta, and macrophage conditioned media. J Exp Zool 292: 477-86.

Lister R, Pelizzola M, Dowen RH, Hawkins RD, Hon G, et al. (2009) Human DNA methylomes at base resolution show widespread epigenomic differences. Nature 462: 315–22.

Little T, Hebert PDN (1996) Ancient asexuals: scandal or artifact? Trend Ecol Evol 11: 296.

Little TJ, Hebert PDN (1997) Clonal diversity in high arctic ostracodes. J Evol Biol 10: 233–52.

Little TJ, DeMelo R, Tailor DJ, Hebert PDN (1997) Genetic characterization of an arctic zooplankter: insights into geographic polyploidy. Proc R Soc Lond Ser B Biol Sci 264: 1363–70.

Little TJ, Ebert D (1999) Associations between parasitism and host genotype in natural populations of Daphnia (Crustacea: Cladocera). J Anim Ecol 68: 134–49.

Little TJ, Ebert D (2001) Temporal patterns of genetic variation for resistance and infectivity in a Daphnia-microparasite system. Evolution 55: 1146–52.

Littlewood TD, Kreuzaler P, Evan GI (2012) All things to all people. Cell 151: 11-3.

Liu A, Mosher DF, Murphy-Ullrich JE, Goldblum SE (2009) The counteradhesive proteins, thrombospondin 1 and SPARC/osteonectin, open the tyrosine phosphorylation-responsive paracellular pathway in pulmonary vascular endothelia. Microvasc Res 77: 13-20.

Liu B, Kanazawa A, Matsumura H, Takahashi R, Harada K, Abe J (2008) Genetic redundancy in soybean photoresponses associated with duplication of the phytochrome A gene. Genetics 180: 995–1007.

Liu B, Chen Y, St Clair DK (2008) ROS and p53: a versatile partnership. Free Radic Biol Med 44: 1529-35.

Liu D, Li L, Fu H, Li S, Li J (2012) Inactivation of Dicer1 has a severe cumulative impact on the formation of mature germ cells in mouse testes. Biochem Biophys Res Commun 422: 114-20.

Liu F, Lou YL, Wu J, Ruan QF, Xie A, et al. (2012) Upregulation of microRNA-210 regulates renal angiogenesis mediated by activation of VEGF signaling pathway under ischemia/perfusion injury in vivo and in vitro. Kidney Blood Press Res 35: 182-91.

Liu G, Chen X (2002) The ferredoxin reductase gene is regulated by the p53 family and sensitizes cells to oxidative stress-induced apoptosis. Oncogene 21: 7195–204.

Liu H, Kim JM, Aoki F (2004) Regulation of histone H3 lysine 9 methylation in oocytes and early pre-implantation embryos. Development 131: 2269–80.

Liu K, Nagase H, Huang CG, Calamita G, Agre P (2006) Purification and functional characterization of aquaporin-8. Biol Cell 98: 153-61.

Liu L, Trimarchi JR, Keefe DL (2000a) Involvement of mitochondria in oxidative stress-induced cell death in mouse zygotes. Biol Reprod 62: 1745–53.

Liu L, Dybvig K, Panangala VS, van Santen VL, French CT (2000b) GAA trinucleotide repeat region regulates M9/pMGA gene expression in Mycoplasma gallisepticum. Infect Immun 68: 871-6.

Liu N, Wang X, McCoubrey WK, Maines MD (2000) Developmentally regulated expression of two transcripts for heme oxygenase-2 with a first exon unique to rat testis: control by corticosterone of the oxygenase protein expression. Gene 241: 175-83.

Liu SK, Tessman I (1990a) groE genes affect SOS repair in Escherichia coli. J Bacteriol 172: 6135-8.

Liu SK, Tessman I (1990b) Error-prone SOS repair can be error-free. J Mol Biol 216: 803-7.

Liu X, Yang Y (2004) Effect of VEGF on the angiogenesis in male reproduction system. Zhonghua Nan Ke Xue 10: 49–51.

Liu X, Hajnóczky G (2011) Altered fusion dynamics underlie unique morphological changes in mitochondria during hypoxia-reoxygenation stress. Cell Death Differ 18: 1561-72.

Liu XD, Thiele DJ (1996) Oxidative stress induced heat shock factor phosphorylation and Hsf-dependent activation of yeast metallothionein gene transcription. Genes Dev 10: 592–603.

Liu XM, Chapman GB,Wang H, Durante W (2000) Adenovirus-mediated heme oxygenase-1 gene expression stimulates apoptosis in vascular smooth muscle cells. Circulation 105: 79–84.

Liu Y, Cox SR, Morita T, Kourembanas S (1995) Hypoxia regulates vascular endothelial growth factor gene expression in endothelial cells. Circ Res 77: 638–43.

Liu Y, Elf SE, Miyata Y, Sashida G, Liu Y, et al. (2009) p53 regulates hematopoietic stem cell quiescence. Cell Stem Cell 4: 37-48.

Liu Y, Lu X (2012) Non-coding RNAs in DNA damage response. Am J Cancer Res 2: 658-75.

Liu Z, Lin H, Ye S, Liu QY, Meng Z, et al. (2006) Remarkably high activities of testicular cytochrome c in destroying reactive oxygen species and in triggering apoptosis. Proc Natl Acad Sci USA 103: 8965–70.

Lively CM (1986) Canalization versus developmental conversion in a spatially-variable environment. Am Nat 128: 561–72.

Lively CM (1987) Evidence from a New-Zealand snail for the maintenance of sex by parasitism. Nature 328: 519-21.

Lively CM (1989) Adaptation by a parasitic trematode to local populations of its snail host. Evolution 43: 1663-71.

Lively CM (1999) Migration, virulence, and the geographic mosaic of adaptation by parasites. Am Nat 153: S34-S47.

Lively CM (2010) A review of Red Queen models for the persistence of obligate sexual reproduction. J Hered 101(Suppl 1): S13–S20.

Lively CM, Craddock C, Vrijenhoek RC (1990) Red Queen hypothesis supported by parasitism in sexual and clonal fish. Nature 344: 864–6.

Lively CM, Howard RS (1994) Selection by parasites for clonal diversity and mixed mating. Philos Trans R Soc Lond B 346: 271–81.

Lively CM, Dybdahl MF (2000) Parasite adaptation to locally common host genotypes. Nature 405: 679-81.

Lively CM, Jokela J (2002) Temporal and spatial distributions of parasites and sex in a freshwater snail. Evol Ecol Res 4: 219–26.

Lively CM, Dybdahl MF, Jokela J, Osnas EE, Delph LF (2004) Host sex and local adaptation by parasites in a snail–trematode interaction. Am Nat 164: S6–S18.

Livnat A, Papadimitriou C, Dushoff J, Feldman MW (2008) A mixability theory for the role of sex in evolution. Proc Natl Acad Sci USA 105: 19803–8.

Livneh Z (2001) DNA damage control by novel DNA polymerases: Translesion replication and mutagenesis. J Biol Chem 276: 25639-42.

Livneh Z (2006) Keeping mammalian mutation load in check: regulation of the activity of error-prone DNA polymerases by p53 and p21. Cell Cycle 5: 1918-22.

Livneh Z, Cohen-Fix O, Skaliter R, Elizur T (1993) Replication of damaged DNA and the molecular mechanism of ultraviolet light mutagenesis. CRC Crit Rev Biochem Mol Biol 28: 465-513.

Lizama C, Alfaro I, Reyes JG, Moreno RD (2007) Up-regulation of CD95 (Apo-1/Fas) is associated with spermatocyte apoptosis during the first round of spermatogenesis in the rat. Apoptosis 12: 499-512.

Ljung R, Petrini P, Tengborn L, Sjörin E (2001) Haemophilia B mutations in Sweden: a population-based study of mutational heterogeneity. Br J Haematol 113: 81–6.

Llewellyn KS, Loxdale HD, Harrington R, Clark SJ, Sunnucks P (2004) Evidence for gene flow and local clonal selection in field populations of the grain aphid (Sitobion avenae) in Britain revealed using microsatellites. Heredity 93: 143–53.

Lloyd DG (1980) Benefits and handicaps of sexual reproduction. Evol Biol 13: 69-111.

Lloyd Morgan C (1896) On modification and variation. Science 4: 733-40.

Lo YY, Cruz TF (1995) Involvement of reactive oxygen species in cytokine and growth factor induction of c-fos expression in chondrocytes. J Biol Chem 270: 11727-30.

Lo YY, Conquer JA, Grinstein S, Cruz TF (1998) Interleukin-1 beta induction of c-fos and collagenase expression in articular chondrocytes: involvement of reactive oxygen species. J Cell Biochem 69: 19-29.

Lobry JR (1997) Influence of genomic G1C content on average amino-acid composition of proteins from 59 bacterial species. Gene 205: 309–16.

Lockless SW, Ranganathan R (1999) Evolutionarily conserved pathways of energetic connectivity in protein families. Science 286: 295–9.

Loeb LA (1998) Cancer cells exhibit a mutator phenotype. Adv Cancer Res 72: 25-56.

Loeb LA (2001) A mutator phenotype in cancer. Cancer Res 61: 3230–9.

Loeb LA, Bielas JH, Beckman RA (2008) Cancers exhibit a mutator phenotype: clinical implications. Cancer Res 68: 3551-7.

Loebel DA, Tsoi B,Wong N, O’Rourke MP, Tam PP (2004) Restricted expression of ETn-related sequences during post-implantation mouse development. Gene Expr Patterns 4: 467–71.

Loenarz C, Coleman ML, Boleininger A, Schierwater B, Holland PWH, et al. (2011) The hypoxia-inducible transcription factor pathway regulates oxygen sensing in the simplest animal, Trichoplax adhaerens. EMBO Rep 12: 63–70.

Loeschcke V, ed. (1987) Genetic constraints on adaptive evolution. New York, NY: Springer.

Loewe L, Textor V, Scherer S (2003) High deleterious mutation rate in stationary phase of Escherichia coli. Science 302: 1558-60.

Loewe L, Charlesworth B (2006) Inferring the distribution of mutational effects on fitness in Drosophila. Biol Lett 2: 426–30.

Loewe L, Lamatsch DK (2008) Quantifying the threat of extinction from Muller’s ratchet in the diploid Amazon molly (Poecilia formosa). BMC Evol Biol 8: 88–108.

Loft S, Astrup A, Buemann B, Poulsen HE (1994) Oxidative DNA damage correlates with oxygen consumption in humans. FASEB J 8:534-7.

Loft S, Poulsen HE (1996) Cancer risk and oxidative DNA damage in man. J Mol Med 74: 297-312.

Loftus B, Anderson I, Davies R, Alsmark UC, Samuelson J, et al. (2005) The genome of the protist parasite Entamoeba histolytica. Nature 433: 865–8.

Loh E, Choe J, Loeb LA (2007) Highly tolerated amino acid substitutions increase the fidelity of E. coli DNA polymerase I. J Biol Chem 282: 12201–9.

Loidl J (1994) Cytological aspects of meiotic recombination. Experientia 50: 285-94.

Lolo FN, Casas-Tintó S, Moreno E (2012) Cell competition time line: winners kill losers, which are extruded and engulfed by hemocytes. Cell Rep 2: 526-39.

Lombaert E, Malausa T, Devred R, Estoup A (2008) Phenotypic variation in invasive and biocontrol populations of the harlequin ladybird, Harmonia axyridis. Biocontrol 53: 89–102.

Long A, Michod RE (1995) Origin of sex for error repair. I. Sex, diploidy, and haploidy. Theor Popul Biol 47: 18-55.

Long TA, Agrawal AF, Rowe L (2012) The effect of sexual selection on offspring fitness depends on the nature of genetic variation. Curr Biol 22: 204-8.

Loomis WF, Kuspa A, Shaulsky G (1998) Two-component signal transduction systems in eukaryotic microorganisms. Curr Opin Microbiol 1: 643-8.

Lopes M, Foiani M, Sogo JM (2006) Multiple mechanisms control chromosome integrity after replication fork uncoupling and restart at irreparable UV lesions. Mol Cell 21: 15-27.

Lopes S, Jurisicova A, Sun JG, Casper RF (1998) Reactive oxygen species: potential cause for DNA fragmentation in human spermatozoa. Hum Reprod 13: 896–900.

López MA, López-Fanjul C (1993) Spontaneous mutation for a quantitative trait in Drosophila melanogaster. I. Response to artificial selection. Genet Res 61: 107-16.

López Castel A, Cleary JD, Pearson CE (2010) Repeat instability as the basis for human diseases and as a potential target for therapy. Nat Rev Mol Cell Biol 11: 165–70.

Lorch PD, Proulx S, Rowe L, Day T (2003) Condition-dependent sexual selection can accelerate adaptation. Evol Ecol Res 5: 867–81.

Loreau M (1998) Ecosystem development explained by competition within and between material cycles. Proc R Soc Lond B Biol Sci 265: 33–8.

Lorenz MG, Wackernagel W (1994) Bacterial gene transfer by natural genetic transformation in the environment. Microbiol Rev 58: 563-602.

Lorenzo-Carballa MO, Cordero-Rivera A (2009) Thelytokous parthenogenesis in the damselfly Ischnura hastata (Odonata, Coenagrionidae): genetic mechanisms and lack of bacterial infection. Heredity 103: 377–84.

Loreto ELS, Carareto CMA, Capy P (2008) Revisiting horizontal transfer of transposable elements in Drosophila. Heredity 100: 545–54.

Lorkovic ZJ, Naumann U, Matzke AJM, Matzke M (2012) Involvement of a GHKL ATPase in RNA-directed DNA methylation in Arabidopsis thaliana. Curr Biol 22: 933–8.

Loscalzo J (2010) The cellular response to hypoxia: tuning the system with microRNAs. J Clin Invest 120: 3815-7.

Losdat S, Helfenstein F, Saladin V, Richner H (2011) Higher in vitro resistance to oxidative stress in extra-pair offspring. J Evol Biol 24: 2525-30.

Losick R, Stragier P (1992) Crisscross regulation of cell-type-specific gene expression during development in B. subtilis. Nature 355: 601-4.

Losick R, Desplan C (2008) Stochasticity and cell fate. Science 320: 65-8.

Losos JB, Warheit KI, Schoener TW (1997) Adaptive differentiation following experimental island colonization in Anolis lizards. Nature 387: 70–3.

Loudon A, Racey PA, eds. (1987) The reproductive energetics of mammals. Oxford, UK: Oxford University Press.

Loukinov DI, Pugacheva E, Vatolin S, Pack SD, Moon H, et al. (2002) BORIS, a novel male germ-line-specific protein associated with epigenetic reprogramming events, shares the same 11-zinc-finger domain with CTCF, the insulator protein involved in reading imprinting marks in the soma. Proc Natl Acad Sci USA 99: 6806–11.

Lourdes de Pereira M, Garcia e Costa F (2003) Spermatogenesis recovery in the mouse after iron injury. Hum Exp Toxicol 22: 275-9.

Love OP, Chin EH, Wynne-Edwards KE, Williams TD (2005) Stress hormones: A link between maternal condition and sex-biased reproductive investment. Am Nat 166: 751–66.

Love OP, Williams TD (2008) The adaptive value of stress-induced phenotypes: effects of maternally-derived corticosterone on sex-biased investment, cost of reproduction, and maternal fitness. Am Nat 172: E135–E149.

Love OP, Gilchrist HG, Bêty J, Wynne-Edwards KE, Berzins L, et al. (2009) Using life-histories to predict and interpret variability in yolk hormones. Gen Comp Endocrinol 163: 169−74.

Love PE, Lyle MJ, Yasbin RE (1985) DNA-damage-inducible (din) loci are transcriptionally activated in competent Bacillus subtilis. Proc Natl Acad Sci USA 82: 6201-5.

Lovell SC, Robertson DL (2010) An integrated view of molecular coevolution in protein-protein interactions. Mol Biol Evol 27:2567-75.

Lovett CM Jr, Love PE, Yasbin RE (1989) Competence-specific induction of the Bacillus subtilis RecA protein analog: evidence for dual regulation of a recombination protein. J Bacteriol 171: 2318-22.

Lowe CB, Bejerano G, Haussler D (2007) Thousands of human mobile element fragments undergo strong purifying selection near developmental genes. Proc Natl Acad Sci USA 104: 8005–10.

Loxdale HD (2008) The nature and reality of the aphid clone: genetic variation, adaptation and evolution. Agricult Forest Ent 10:81-90.

Loxdale HD (2009) What’s in a clone: The rapid evolution of aphid asexual lineages in relation to geography, host plant adaptation and resistance to pesticides. In: Schön I, Martens K, van Dijk P, eds. Lost sex. New York, NY: Springer. pp 535–557.

Loxdale HD (2010) Rapid genetic changes in natural insect populations. Ecol Entomol 35: 155–64.

Loxdale HD, Lushai G (2003) Rapid changes in clonal lines: the death of a ‘sacred cow’. Biol J Linn Soc 79: 3–16.

Loxdale HD, Weisser WW (2011) Evidence for clonal selection and the need for the maintenance of sexual recombination: a reappraisal and overview. Acta Hort (ISHS) 904: 133-149.

Loxdale HD, Lushai G, Harvey JA (2011) The evolutionary improbability of ‘generalism’ in nature, with special reference to insects. Biol J Linn Soc 103: 1-18.

Lozovsky ER, Chookajorn T, Brown KM, Imwong M, Shaw PJ, et al. (2009) Stepwise acquisition of pyrimethamine resistance in the malaria parasite. Proc Natl Acad Sci USA 106: 12025–30.

Lu C, Brauer MJ, Botstein D (2009) Slow growth induces heat-shock resistance in normal and respiratory-deficient yeast. Mol Biol Cell 20: 891-903.

Lu J, Wu CI (2005) Weak selection revealed by the whole-genome comparison of the X chromosome and autosomes of human and chimpanzee. Proc Natl Acad Sci USA 102: 4063–7.

Lu JP, Monardo L, Bryskin I, Hou ZF, Trachtenberg J, et al. (2010) Androgens induce oxidative stress and radiation resistance in prostate cancer cells though NADPH oxidase. Prostate Cancer Prostatic Dis 13: 39-46.

Lu KP, Ramos KS (2003) Redox regulation of a novel L1Md-A2 retrotransposon in vascular smooth muscle cells. J Biol Chem 278: 28201-9.

Lu WJ, Chapo J, Roig I, Abrams JM (2010) Meiotic recombination provokes functional activation of the p53 regulatory network. Science 328: 1278-81.

Lu WJ, Lee NP, Kaul SC, Lan F, Poon RT, et al. (2011) Mortalin-p53 interaction in cancer cells is stress dependent and constitutes a selective target for cancer therapy. Cell Death Differ 18: 1046-56.

Lu X, Lane DP (1993) Differential induction of transcriptionally active p53 following UV or ionizing radiation: defects in chromosome instability syndromes? Cell 75: 765–78.

Luce K, Weil AC, Osiewacz HD (2010) Mitochondrial protein quality control systems in aging and disease. Adv Exp Med Biol 694: 108-25.

Lucesoli F, Caligiuri M, Roberti MF, Perazzo JC, Fraga CG (1999) Dose-dependent increase of oxidative damage in the testes of rats subjected to acute iron overload. Toxicology 132: 179-86.

Lucht JM, Mauch-Mani B, Steiner HY, Metraux JP, Ryals J, Hohn B (2002) Pathogen stress increases somatic recombination frequency in Arabidopsis. Nat Genet 30: 311–4.

Lue YH, Sinha Hikim AP, Swerdloff RS, Im P, Taing KS, et al. (1999) Single exposure to heat induces stage-specific germ cell apoptosis in rats: role of intratesticular testosterone on stage specificity. Endocrinology 140: 1709-17.

Lue YH, Lasley BL, Laughlin LS, Swerdloff RS, Hikim AP, et al. (2002) Mild testicular hyperthermia induces profound transitional spermatogenic suppression through increased germ cell apoptosis in adult cynomolgus monkeys (Macaca fascicularis). J Androl 23: 799-805.

Lue Y, Sinha Hikim AP, Wang C, Leung A, Swerdloff RS (2003) Functional role of inducible nitric oxide synthase in the induction of male germ cell apoptosis, regulation of sperm number, and determination of testes size: evidence from null mutant mice. Endocrinology 144: 3092–100.

Lue Y, Wang C, Liu YX, Hikim AP, Zhang, XS, et al. (2006) Transient testicular warming enhances the suppressive effect of testosterone on spermatogenesis in adult cynomolgus monkeys (Macaca fascicularis). J Clin Endocrinol Metab 91: 539–45.

Luetjens CM, Weinbauer GF, Wistuba J (2005) Primate spermatogenesis: new insights into comparative testicular organisation, spermatogenic efficiency and endocrine control. Biol Rev Camb Philos Soc 80: 475-88.

Lukas J, Lukas C, Bartek J (2011) More than just a focus: The chromatin response to DNA damage and its role in genome integrity maintenance. Nat Cell Biol 13: 1161-9.

Lukens LN, Pires JC, Leon E, Vogelzang R, Oslach L, Osborn T (2006) Patterns of sequence loss and cytosine methylation within a population of newly resynthesized Brassica napus allopolyploids. Plant Physiol 140: 336–48.

Lum JJ, Bui T, Gruber M, Gordan JD, DeBerardinis RJ, et al. (2007) The transcription factor HIF-1alpha plays a critical role in the growth factor-dependent regulation of both aerobic and anaerobic glycolysis. Genes Dev 21: 1037–49.

Lummaa V (2003) Early developmental condtions and reproductive success in humans: downstream effects of prenatal famine, birthweight, and timing of birth. Am J Hum Biol 15: 370-9.

Lummaa V, Clutton-Brock TH (2002) Early development, survival and reproduction in humans. Trends Ecol Evol 17: 141-7.

Lundmark M, Saura A (2006) Asexuality alone does not explain the success of clonal forms in insects with geographical parthenogenesis. Hereditas 143: 23–32.

Lunt DH (2008) Genetic tests of ancient asexuality in root knot nematodes reveal recent hybrid origins. BMC Evol Biol 8: 194.

Lunt DH, Hyman BC (1997) Animal mitochondrial DNA recombination. Nature 387: 247.

Lunt ID (1990) Seed longevity of six native forbs in a closed Themeda triandra grassland. Aust J Bot 43: 439–49.

Luo L, Chen H, Trush MA, Show MD, Anway MD, Zirkin BR (2006) Aging and the Brown Norway rat Leydig cell antioxidant defense system. J Androl 27: 240–7.

Luo L, Ye L, Liu G, Shao G, Zheng R, et al. (2010) Microarray-based approach identifies differentially expressed microRNAs in porcine sexually immature and mature testes. PLoS ONE 5: e11744.

Luo M, He H, Kelley MR, Georgiadis MM (2010) Redox regulation of DNA repair: implications for human health and cancer therapeutic development. Antioxid Redox Signal 12: 1247-69.

Luo S, Kleemann GA, Shaw WM, Murphy CT (2010) TGF-beta and insulin signaling regulate reproductive aging via oocyte and germline quality maintenance. Cell 143: 299–312.

Luo S, Murphy CT (2011) Caenorhabditis elegans reproductive aging: regulation and underlying mechanisms. Genesis 49: 53–65.

Luo ZC, Fraser WD, Julien P, Deal CL, Audibert F, et al. (2006) Tracing the origins of "fetal origins" of adult diseases: programming by oxidative stress? Med Hypotheses 66: 38-44.

Lüpold S, Linz GM, Rivers JW, Westneat DF, Birkhead TR (2009) Sperm competition selects beyond relative testes size in birds. Evolution 63: 391–402.

Lupu A, Pechkovskaya A, Rashkovetsky E, Nevo E, Korol A (2004) DNA repair efficiency and thermotolerance in Drosophila melanogaster from ‘Evolution Canyon’. Mutagenesis 19: 383–90.

Lupu A, Nevo E, Zamorzaeva I, Korol A (2006) Ecological-genetic feedback in DNA repair in wild barley, Hordeum spontaneum. Genetica 127: 121–32.

Luria SE, Delbrück M (1943) Mutations of bacteria from virus sensitivity to virus resistance. Genetics 28: 491–511.

Lürling M, Roozen F, Van Donk E, Goser B (2003) Response of Daphnia to substances released from crowded congeners and conspecifics. J Plankton Res 25: 967–78.

Lushai G, De Barro PJ, David O, Sherratt TN, Maclean N (1998) Genetic variation within a parthenogenetic lineage. Insect Mol Biol 7: 337–44.

Lushai G, Loxdale HD, Maclean N (2000) Genetic diversity in clonal lineages. J Reprod Dev 46: 21–2.

Lushai G, Loxdale HD (2002) The biological improbability of a clone (mini-review). Genet Res 79: 1–9.

Lushai G, Loxdale HD, Allen JA (2003) The dynamic clonal genome and its adaptive potential. Biol J Linn Soc 79: 193-208.

Lushchak VI (2011) Adaptive response to oxidative stress: Bacteria, fungi, plants and animals. Comp Biochem Physiol C Toxicol Pharmacol 153: 175-90.

Luttikhuizen PC, Stift M, Kuperus P, VAN Tienderen PH (2007) Genetic diversity in diploid vs. tetraploid Rorippa amphibia (Brassicaceae). Mol Ecol 16: 3544-53.

Lutz S, Weisser HJ, Heizmann J, Pollak S (2000) Mitochondrial heteroplasmy among maternally related individuals. Int J Legal Med 113: 155-61.

Lutz WK (1990) Endogenous genotoxic agents and processes as a basis of spontaneous carcinogenesis. Mutat Res 238: 287-95.

Lutzoni F, Pagel M (1997) Accelerated evolution as a consequence of transitions to mutualism. Proc Natl Acad Sci USA 94: 11422-7.

Lv DK, Bai X, Li Y, Ding XD, Ge Y, et al. (2010) Profiling of cold-stress-responsive miRNAs in rice by microarrays. Gene 459: 39–47.

Lynch AJJ, Barnes RW, Cambecedes J, Vaillancourt RE (1998) Genetic evidence that Lomatia tasmanica (Proteaceae) is an ancient clone. Aust J Bot 46: 25–33.

Lynch M (1984) Destabilizing hybridization, general-purpose genotypes and geographic parthenogenesis. Q Rev Biol 59: 257–90.

Lynch M (2006) The origins of eukaryotic gene structure. Mol Biol Evol 23: 450–68.

Lynch M (2007a) The origins of genome architecture. Sunderland, MA: Sinauer Associates.

Lynch M (2007b) The frailty of adaptive hypotheses for the origins of organismal complexity. Proc Natl Acad Sci USA 104 Suppl 1: 8597-604.

Lynch M (2010a) Evolution of the mutation rate. Trends Genet 26: 345–52.

Lynch M (2010b) Rate, molecular spectrum, and consequences of human mutation. Proc Natl Acad Sci USA 107: 961–8.

Lynch M (2011) The lower bound to the evolution of mutation rates. Genome Biol Evol 3: 1107–18.

Lynch M, Hill WG (1986) Phenotypic evolution by neutral mutation. Evolution 40: 915-35.

Lynch M, Gabriel W (1990) Mutation load and the survival of small populations. Evolution 44: 1725-37.

Lynch M, Bürger R, Butcher D, Gabriel W (1993) The mutational meltdown in asexual populations. J Hered 84: 339-44.

Lynch M, Spitze K (1994) Evolutionary genetics of Daphnia. In: Real LA, ed. Ecological Genetics. Princeton, NJ: Princeton University Press. pp 109–128.

Lynch M, Conery J, Bürger R (1995a) Mutational meltdowns in sexual populations. Evolution 49: 1067–80.

Lynch M, Conery J, Bürger R (1995b) Mutation accumulation and the extinction of small populations. Am Nat 146: 489–518.

Lynch M, Walsh B (1998) Genetics and Analysis of Quantitative Traits. Sunderland, MA: Sinauer Associates.

Lynch M, Latta L, Hicks J, Giorgianni M (1998) Mutation, selection, and the maintenance of life-history variation in a natural population. Evolution 52: 727–33.

Lynch M, Blanchard J, Houle D, Kibota T, Schultz S, Vassilieva L, Willis J (1999) Perspective: spontaneous deleterious mutation. Evolution 53: 645–63.

Lynch M, O’Hely M (2001) Captive breeding and the genetic fitness of natural populations. Conserv Genet 2: 363–78.

Lynch M, Conery JS (2003) The origins of genome complexity. Science 302: 1401–4.

Lynch M, Koskella B, Schaack S (2006) Mutation pressure and the evolution of organelle genomic architecture. Science 311: 1727–30.

Lynch M, Sung W, Morris K, Coffey N, Landry CR, et al. (2008) A genomewide view of the spectrum of spontaneous mutations in yeast. Proc Natl Acad Sci USA 105: 9272–7.

Lynch VJ, Leclerc RD, May G, Wagner GP (2011) Transposon-mediated rewiring of gene regulatory networks contributed to the evolution of pregnancy in mammals. Nat Genet 43: 1154-9.

Lysiak JJ, Nguyen QA, Turner TT (2000a) Fluctuations in rat testicular interstitial oxygen tensions are linked to testicular vasomotion: persistence after repair of torsion. Biol Reprod 63: 1383-9.

Lysiak JJ, Turner SD, Turner TT (2000b) Molecular pathway of germ cell apoptosis following ischemia/reperfusion of the rat testis. Biol Reprod 63: 1465–72.

Lysiak JJ, Zheng S, Woodson R, Turner TT (2007) Caspase-9-dependent pathway to murine germ cell apoptosis: mediation by oxidative stress, BAX, and caspase 2. Cell Tissue Res 328: 411-9.

Lysiak JJ, Kirby JL, Tremblay JJ, Woodson RI, Reardon MA, et al. (2009) Hypoxia-inducible factor-1alpha is constitutively expressed in murine Leydig cells and regulates 3beta-hydroxysteroid dehydrogenase type 1 promoter activity. J Androl 30:146-56.

Lythgoe KA (2000) The coevolution of parasites with host-acquired immunity and the evolution of sex. Evolution 54: 1142–56.

Ma A, Qi S, Chen H (2008) Antioxidant therapy for prevention of inflammation, ischemic reperfusion injuries and allograft rejection. Cardiovasc Hematol Agents Med Chem 6: 20-43.

Ma L, Buchold GM, Greenbaum MP, Roy A, Burns KH, et al. (2009) GASZ is essential for male meiosis and suppression of retrotransposon expression in the male germline. PLoS Genet 5: e1000635.

Ma T, Yang B, Verkman AS (1997) Cloning of a novel water and urea-permeable aquaporin from mouse expressed strongly in colon, placenta, liver, and heart. Biochem Biophys Res Commun 240: 324–8.

Ma W, Tessarollo L, Hong SB, Baba M, Southon E, et al. (2003) Hepatic vascular tumors, angiectasis in multiple organs, and impaired spermatogenesis in mice with conditional inactivation of the VHL gene. Cancer Res 63: 5320–8.

Ma X, Fan L, Meng Y, Hou Z, Mao YD, Wang W, Ding W, Liu JY (2007) Proteomic analysis of human ovaries from normal and polycystic ovarian syndrome. Mol Hum Reprod 13: 527–35.

Ma Y, Creanga A, Lum L, Beachy PA (2006) Prevalence of off-target effects in Drosophila RNA interference screens. Nature 443: 359-63.

Maamar H, Raj A, Dubnau D (2007) Noise in gene expression determines cell fate in Bacillus subtilis. Science 317: 526–9.

Maan ME, Seehausen O (2011) Ecology, sexual selection and speciation. Ecol Lett 14: 591-602.

Maatouk DM, Loveland KL, McManus MT, Moore K, Harfe BD (2008) Dicer1 is required for differentiation of the mouse male germline. Biol Reprod 79: 696-703.

Mable BK, Otto SP (2001) Masking and purging mutations following EMS treatment in haploid, diploid and tetraploid yeast (Saccharomyces cerevisiae). Genet Res 77: 9–26.

MacArthur RH (1962) Some generalized theorems of natural selection. Proc Natl Acad Sci USA 48: 1893-7.

MacArthur R (1970) Species packing and competitive equilibrium for many species. Theor Popul Biol 1: 1–11.

MacArthur RH (1972a) Geographical ecology. New York, NY: Harper and Row.

MacArthur RH (1972b) Geographical Ecology: Patterns in the Distribution of Species. Princeton, NJ: Princeton University Press

MacArthur R, Levins R (1964) Competition, habitat selection, and character displacement in a patchy environment. Proc Natl Acad Sci USA 51: 1207-10.

MacArthur RH, Pianka ER (1966) On optimal use of a patchy environment. Am Nat 100: 603-9.

MacArthur RH, Wilson EO (1967) The theory of island biogeography. Princeton, NJ: Princeton University Press.

MacCarthy T, Bergman A (2007) Coevolution of robustness, epistasis, and recombination favors asexual reproduction. Proc Natl Acad Sci USA 104: 12801–6.

Macía J, Solé RV, Elena SF (2012) The causes of epistasis in genetic networks. Evolution 66: 586-96.

Maciá MD, Blanquer D, Togores B, Sauleda J, Pérez JL, Oliver A (2005) Hypermutation is a key factor in development of multiple-antimicrobial resistance in Pseudomonas aeruginosa strains causing chronic lung infections. Antimicrob Agents Chemother 49: 3382–6.

Macip S, Igarashi M, Berggren P, Yu J, Lee SW, Aaronson SA (2003) Influence of induced reactive oxygen species in p53-mediated cell fate decisions. Mol Cell Biol 23: 8576–85.

MacKay PA (1989) Clonal variation in sexual morph production in Acyrthosiphon pisum (Homoptera: Aphididae). Environ Entomol 18: 558–62.

Mackay TF, Fry JD, Lyman RF, Nuzhdin SV (1994) Polygenic mutation in Drosophila melanogaster: estimates from response to selection of inbred strains. Genetics 136: 937-51.

Mackay TFC, Lyman R, Jackson MS (1992) Effects of P elements on quantitative traits in Drosophila melanogaster. Genetics 130: 315-32.

Mackey ZB, Ramos W, Levin DS, Walter CA, McCarrey JR, Tomkinson AE (1997) An alternative splicing event which occurs in mouse pachytene spermatocytes generates a form of DNA ligase III with distinct biochemical properties that may function in meiotic recombination. Mol Cell Biol 17: 989–98.

Mackwan RR, Carver GT, Drake JW, Grogan DW (2007) An unusual pattern of spontaneous mutations recovered in the halophilic archaeon Haloferax volcanii. Genetics 176: 697–702.

MacNaughton J, Banah M, McCloud P, Hee J, Burger H (1992) Age related changes in follicle stimulating hormone, luteinizing hormone, oestradiol and immunoreactive inhibin in women of reproductive age. Clin Endocrinol (Oxf) 36: 339–45.

MacLean R (2008) The tragedy of the commons in microbial populations: insights from theoretical, comparative and experimental studies. Heredity 100: 471–7.

MacLeod MC (1995) Interaction of bulky chemical carcinogens with DNA in chromatin. Carcinogenesis 16: 2009–14.

Macmillan-Crow LA, Cruthirds DL (2001) Invited review: manganese superoxide dismutase in disease. Free Radic Res 34: 325–36.

MacNeill A (2009) http://evolutionlist.blogspot.com/2009/03/are-mechanisms-that-produce-phenotypic.html

MacPhee DG (1985) Indications that mutagenesis in Salmonella may be subject to catabolite repression. Mutat Res 151: 35-41.

MacPhee DG (1994) Regulatory processes and the origins of spontaneous mutations. Mutat Res 307: 115-20.

MacPhee DG (1996) Mismatch repair as a source of mutations in non-dividing cells. Genetica 97: 183-95.

MacPhee DG (1999) Adaptive mutability in bacteria. J Genet 78: 29-33.

MacPhee DG, Ambrose M (2010) Catabolite repression of SOS-dependent and SOS-independent spontaneous mutagenesis in stationary-phase Escherichia coli. Mutat Res 686: 84–9.

Macpherson JM, Sella G, Davis JC, Petrov DA (2007) Genomewide spatial correspondence between nonsynonymous divergence and neutral polymorphism reveals extensive adaptation in Drosophila. Genetics 177: 2083–99.

MacRae AF, Anderson WW (1988) Evidence for non-neutrality of mitochondrial DNA haplotypes in Drosophila pseudoobscura. Genetics 120: 485-94.

Madan E, Gogna R, Bhatt M, Pati U, Kuppusamy P, Mahdi AA (2011) Regulation of glucose metabolism by p53: emerging new roles for the tumor suppressor. Oncotarget 2: 948-57.

Maddocks OD, Vousden KH (2011) Metabolic regulation by p53. J Mol Med (Berl) 89: 237-45.

Maderspacher F (2008) Sex and the drought. Curr Biol 18: R983–5.

Maderspacher F (2011) Asexuality: the insects that stick with it. Curr Biol 21: R495-7.

Madhani HD, Fink GR (1998) The control of filamentous differentiation and virulence in fungi. Trends Cell Biol 8: 348-53.

Madison-Villar MJ, Michalak P (2011) Misexpression of testicular microRNA in sterile Xenopus hybrids points to tetrapod-specific microRNAs associated with male fertility. J Mol Evol 73: 316-24.

Madlung A, Masuelli RW, Watson B, Reynolds SH, Davison J, Comai L (2002) Remodeling of DNA methylation and phenotypic and transcriptional changes in synthetic Arabidopsis allotetraploids. Plant Physiol 129: 733–46.

Madlung A, Tyagi AP, Watson B, Jiang H, Kagochi T, et al. (2005) Genomic changes in synthetic Arabidopsis polyploids. Plant J 41: 221–30.

Madrid E, Reyes JG, Hernández B, García JM, San Martín S, et al. (2012) Effect of normobaric hypoxia on the testis in a murine model. Andrologia. 2012 Sep 11. doi: 10.1111/and.12019. [Epub ahead of print]

Madsen BE, Villesen P, Wiuf C (2008) Short tandem repeats in human exons: a target for disease mutations. BMC Genomics 9: 410.

Madsen T, Shine R, Loman J, Hakansson T (1992) Why do female adders copulate so frequently? Nature 335: 440–1.

Maeda Y, Shiratsuchi A, Namiki M, Nakanishi Y (2002) Inhibition of sperm production in mice by annexin V microinjected into seminiferous tubules: possible etiology of phagocytic clearance of apoptotic spermatogenic cells and male infertility. Cell Death Differ 9: 742—9.

Magni GE (1963) The origin of spontaneous mutations during meiosis. Proc Natl AcadSci USA 50: 975-80.

Magni GE,von Borstel RC (1962) Different rates of spontaneous mutation during mitosis and meiosis in yeast. Genetics 47: 1097-108.

Magwire MM, Bayer F, Webster CL, Cao C, Jiggins FM (2011) Successive increases in the resistance of Drosophila to viral infection through a transposon insertion followed by a duplication. PLoS Genet 7: e1002337.

Mahadev K, Zilbering A, Zhu L, Goldstein BJ (2001) Insulin-stimulated hydrogen peroxide reversibly inhibits protein-tyrosine phosphatase 1b in vivo and enhances the early insulin action cascade. J Biol Chem 276: 21938-42.

Mahadevaiah SK, Turner JM, Baudat F, Rogakou EP, de Boer P, et al. (2001) Recombinational DNA double-strand breaks in mice precede synapsis. Nat Genet 27: 271-6.

Mahadevan M, Tsilfidis C, Sabourin L, Shutler G, Amemiya C, et al. (1992) Myotonic dystrophy mutation: an unstable CTG repeat in the 3' untranslated region of the gene. Science 255: 1253-5.

Mahakali Zama A, Hudson III FP, Bedell MA (2005) Analysis of hypomorphic KitlSl mutants suggests different requirements for KITL in proliferation and migration of mouse primordial germ cells. Biol Reprod 73: 639–47.

Maharjan RP, Liu B, Li Y, Reeves PR, Wang L, Ferenci T (2012) Mutation accumulation and fitness in mutator subpopulations of Escherichia coli. Biol Lett 9: 20120961.

Maheshwari A, Misro MM, Aggarwal A, Sharma RK, Nandan D (2009) Pathways involved in testicular germ cell apoptosis induced by H2O2 in vitro. FEBS J 276: 870-81.

Mahowald AP (1962) Fine structure of pole cells and polar granules in Drosophila melanogaster. J Exp Zool 151: 201–15.

Mahowald AP (1968) Polar granules of Drosophila. II. Ultrastructural changes during early embryogenesis. J Exp Zool 167: 237–61.

Mahowald AP (1971) Polar granules of Drosophila. 3. The continuity of polar granules during the life cycle of Drosophila. J Exp Zool 176: 329–43.

Mahpatra S, Firpo MT, Bacanamwo M (2010) Inhibition of DNA methyltransferases and histone deacetylases induces bone marrow-derived multipotent adult progenitor cells to differentiate into endothelial cells. Ethn Dis 20(1 Suppl 1): S1-60-4.

Mai B, Breeden L (2000) CLN1 and its repression by Xbp1 are important for efficient sporulation in budding yeast. Mol Cell Biol 20: 478–87.

Mai WJ, Yan JL, Wang L, Zheng Y, Xin Y, Wang WN (2010) Acute acidic exposure induces p53-mediated oxidative stress and DNA damage in tilapia (Oreochromis niloticus) blood cells. Aquat Toxicol 100: 271-81.

Maia LP (2009) Analytical results on Muller’s ratchet effect in growing populations. Phys Rev E 79: 032903.

Mailloux RJ, Seifert EL, Bouillaud F, Aguer C, Collins S, Harper ME (2011) Glutathionylation acts as a control switch for uncoupling proteins UCP2 and UCP3. J Biol Chem 286: 21865-75.

Mailloux RJ, Harper ME (2011) Uncoupling proteins and the control of mitochondrial reactive oxygen species production. Free Radic Biol Med 51: 1106-15.

Mailloux RJ, Harper ME (2012) Mitochondrial proticity and ROS signaling: lessons from the uncoupling proteins. Trends Endocrinol Metab 23: 451-8.

Maines JZ, Stevens LM, Tong X, Stein D (2004) Drosophila dMyc is required for ovary cell growth and endoreplication. Development 131: 775-86.

Maines MD (1997) The heme oxygenase system: a regulator of second messenger gases. Annu Rev Pharmacol Toxicol 37: 517-54.

Maines MD, Gibbs PE (2005) 30 some years of heme oxygenase: from a "molecular wrecking ball" to a "mesmerizing" trigger of cellular events. Biochem Biophys Res Commun 338: 568-77.

Maisnier-Patin S, Roth JR, Fredriksson A, Nyström T, Berg OG, Andersson DI (2005) Genomic buffering mitigates the effects of deleterious mutations in bacteria. Nat Genet 37: 1376-9.

Majewski J (2001) Sexual isolation in bacteria. FEMS Microbiol Lett 199:161-9.

Majewski J, Ott J (2000) GT repeats are associated with recombination on human chromosome 22. Genome Res 10: 1108-14.

Majmundar AJ, Wong WJ, Simon MC (2010) Hypoxia-inducible factors and the response to hypoxic stress. Mol Cell 40: 294-309.

Majumdar A, Chatterjee AG, Ripmaster TL, Levin HL (2011) Determinants that specify the integration pattern of retrotransposon Tf1 in the fbp1 promoter of Schizosaccharomyces pombe. J Virol 85: 519-29.

Makalowski W, Zhang J, Boguski MS (1996) Comparative analysis of 1196 orthologous mouse and human full-length mRNA and protein sequences. Genome Res 6: 846-57.

Maki H (2002) Origins of spontaneous mutations: specificity and directionality of base-substitution, frameshift, and sequence-substitution mutagenesis. Annu Rev Genet 36: 279–303.

Makino N, Sasaki K, Hashida K, Sakakura Y (2004) A metabolic model describing the H2O2elimination by mammalian cells including H2O2 permeation through cytoplasmic and peroxisomal membranes: comparison with experimental data. Biochim Biophys Acta 1673: 149–59.

Makino Y, Okamoto K, Yoshikawa N, Aoshima M, Hirota K, et al. (1996) Thioredoxin: a redox-regulating cellular cofactor for glucocorticoid hormone action. Cross talk between endocrine control of stress response and cellular antioxidant defense system. J Clin Invest 98: 2469–77.

Makova KD, Li WH (2002) Strong male-driven evolution of DNA sequences in humans and apes. Nature 416: 624–6.

Makova KD, Yang S, Chiaromonte F (2004) Insertions and deletions are male biased too: a whole-genome analysis in rodents. Genome Res 14: 567–73.

Maksakova IA, Romanish MT, Gagnier L, Dunn CA, van de Lagemaat LN, et al. (2006) Retroviral elements and their hosts: Insertional mutagenesis in the mouse germ line. PLoS Genet 2: e2.

Maksakova IA, Mager DL, Reiss D (2008) Keeping active endogenous retroviral-like elements in check: the epigenetic perspective. Cell Mol Life Sci 65: 3329–47.

Malagnac F, Lalucque H, Lepère G, Silar P (2004) Two NADPH oxidase isoforms are required for sexual reproduction and ascospore germination in the filamentous fungus Podospora anserina. Fungal Genet Biol 41: 982-97.

Malena A, Loro E, Di Re M, Holt IJ, Vergani L (2009) Inhibition of mitochondrial fission favours mutant over wild-type mitochondrial DNA. Hum Mol Genet 18: 3407-16.

Maley CC, Forrest S (2000) Exploring the relationship between neutral and selective mutations in cancer. Artif Life 6: 325–45.

Malik S-B, Pightling AW, Stefaniak LM, Schurko AM, Logsdon JM Jr (2008) An expanded inventory of conserved meiotic genes provides evidence for sex in Trichomonas. PLoS ONE 3: e2879.

Malinouski M, Zhou Y, Belousov VV, Hatfield DL, Gladyshev VN (2011) Hydrogen peroxide probes directed to different cellular compartments. PLoS One 6: e14564.

Mallet J (2010) The 'struggle for existence:' why the mismatch of basic theory in ecology and evolution?http://abacus.gene.ucl.ac.uk/jim/pap/mallet_the_struggle_2010.pdf

Mallet MA, Chippindale AK (2011) Inbreeding reveals stronger net selection on Drosophila melanogaster males: implications for mutation load and the fitness of sexual females. Heredity 106: 994-1002.

Mallette FA, Calabrese V, Ilangumaran S, Ferbeyre G (2010) SOCS1, a novel interaction partner of p53 controlling oncogene-induced senescence. Aging (Albany NY) 2: 445-52.

Malo AF, Garde JJ, Soler AJ, García AJ, Gomendio M, et al. (2005) Male fertility in natural populations of red deer is determined by sperm velocity and the proportion of normal spermatozoa. Biol Reprod 72: 822–9.

Malo AF, Martinez-Pastor F, Alaks G, Dubach J, Lacy RC (2010) Effects of genetic captive-breeding protocols on sperm quality and fertility in the white-footed mouse. Biol Reprod 83: 540-8.

Malone CD, Hannon GJ (2009) Small RNAs as guardians of the genome. Cell 136: 656–68.

Malone CD, Brennecke J, Dus M, Stark A, McCombie WR, et al. (2009) Specialized piRNA pathways act in germline and somatic tissues of the Drosophila ovary. Cell 137: 522–35.

Malone JH, Michalak P (2008) Gene expression analysis of the ovary of hybrid females of Xenopus laevis and X. muelleri. BMC Evol Biol 8: 82.

Malone RE, Esposito RE (1980) The RAD52 gene is required for homothallic interconversion of mating types and spontaneous mitotic recombination in yeast. Proc Natl Acad Sci USA 77: 503-7.

Maloyan A, Palmon A, Horowitz M (1999) Heat acclimation increases the basal Hsp72 level and alters its production dynamics during heat stress. Am J Physiol Regul Integr Comp Physiol 276: R1506–R1515.

Malpaux B, Migaud M, Tricoire H, Chemineau P (2001) Biology of mammalian photoperiodism and the critical role of the pineal gland and melatonin. J Biol Rhythms 16: 336-47.

Maltepe E, Schmidt JV, Baunoch D, Bradfield CA, Simon MC (1997) Abnormal angiogenesis and responses to glucose and oxygen deprivation in mice lacking the protein ARNT. Nature 386: 403–7.

Maltepe E, Simon MC (1998) Oxygen, genes, and development: an analysis of the role of hypoxic gene regulation during murine vascular development. J Mol Med 76: 391–401.

Maltepe E, Krampitz GW, Okazaki KM, Red-Horse K, Mak W, et al. (2005) Hypoxia-inducible factor-dependent histone deacetylase activity determines stem cell fate in the placenta. Development 132: 3393-403.

Mammoto A, Masumoto N, Tahara M, Ikebuchi Y, Ohmichi M, et al. (1996) Reactive oxygen species block sperm-egg fusion via oxidation of sperm sulfhydryl proteins in mice. Biol Reprod 55: 1063-8.

Managadze D, Rogozin IB, Chernikova D, Shabalina SA, Koonin EV (2011) Negative correlation between expression level and evolutionary rate of long intergenic noncoding RNAs. Genome Biol Evol 3: 1390-404.

Manalo DJ, Lin Z, Liu AY (2002) Redox-dependent regulation of the conformation and function of human heat shock factor 1. Biochemistry 41: 2580–8.

Manandhar G, Miranda-Vizuete A, Pedrajas JR, Krause WJ, Zimmerman S, et al. (2009) Peroxiredoxin 2 and peroxidase enzymatic activity of mammalian spermatozoa. Biol Reprod 80: 1168-77.

Mancera E, Bourgon R, Brozzi A, Huber W, Steinmetz LM (2008) High-resolution mapping of meiotic crossovers and non-crossovers in yeast. Nature 454: 479–85.

Mancini D, Singh S, Ainsworth P, Rodenhiser D (1997) Constitutively methylated CpG dinucleotides as mutation hot spots in the retinoblastoma gene (RB1). Am J Hum Genet 61: 80–7.

Mandegar MA, Otto SP (2007) Mitotic recombination counteracts the benefits of genetic segregation. Proc Biol Sci 274: 1301-7.

Mangiarini L, Sathasivam K, Mahal A, Mott R, Seller M, Bates GP (1997) Instability of highly expanded CAG repeats in mice transgenic for the Huntington’s disease mutation. Nat Genet 15: 197-200.

Mani SR, Juliano CE (2013) Untangling the web: The diverse functions of the PIWI/piRNA pathway. Mol Reprod Dev 2013 May 24. doi: 10.1002/mrd.22195. [Epub ahead of print]

Manier MK, Belote JM, Berben KS, Novikov D, Stuart WT, et al. (2010) Resolving mechanisms of competitive fertilization success in Drosophila melanogaster. Science 328: 354–7.

Mann DR, Jackson GG, Blank MS (1982) Influence of adrenocorticotropin and adrenalectomy on gonadotropin secretion in immature rats. Neuroendocrinology 34: 20-6.

Mann T, Lutwak-Mann C (1981) Enzymes as biochemical markers in differentiating germ cells. In: Mann T, ed.Male Reproductive Function and Semen: Themes and Trends in Physiology, Biochemistry, and Investigative Andrology. Berlin, Germany: Springer-Verlag. pp 101–106.

Manna I, Jana K, Samanta PK (2003) Effect of intensive exercise-induced testicular gametogenic and steroidogenic disorders in mature male Wistar strain rats: A correlative approach to oxidative stress. Acta Physiol Scand 178: 33-40.

Manning JA, Kumar S (2010) A potential role for NEDD1 and the centrosome in senescence of mouse embryonic fibroblasts. Cell Death Dis 1: e35.

Manning JT (1976) Is sex maintained to fascilitate or minimise mutational advance? Heredity (Edinburgh) 36: 351-7.

Manning JT, Dickson DP (1986) Environmental change, mutational load and the advantage of sexual reproduction. Acta Biotheor 35: 149-62.

Manning JT, Chamberlain AT (1994) Sib competition and sperm competitiveness: an answer to 'why so many sperms?' and the recombination/sperm number correlation. Proc Biol Sci 256: 177-82.

Manoel D, Carvalho S, Phillips PC, Teotónio H (2007) Selection against males in Caenorhabditis elegans under two mutational treatments. Proc R Soc B Biol Sci 274: 417–24.

Manoli I, Le H, Alesci S, McFann KK, Su YA, et al. (2005) Monoamine oxidase-A is a major target gene for glucocorticoids in human skeletal muscle cells. FASEB J 19: 1359-61.

Manoli I, Alesci S, Blackman MR, Su YA, Rennert OM, Chrousos GP (2007) Mitochondria as key components of the stress response. Trends Endocrinol Metab 18: 190-8.

Manolio TA, Brooks LD, Collins FS (2008) A HapMap harvest of insights into the genetics of common disease. J Clin Invest 118: 1590–1605.

Manova K, Huang EJ, Angeles M, De Leon V, Sanchez S, et al. (1993) The expression pattern of the c-kit ligand in gonads of mice supports a role for the c-kit receptor in oocyte growth and in proliferation of spermatogonia. Dev Biol 157: 85-99.

Manrubia SC, Escarmís C, Domingo E, Lázaro E (2005) High mutation rates, bottlenecks, and robustness of RNA viral quasispecies. Gene 347: 273-82.

Mansfield KD, Guzy RD, Pan Y, Young RM, Cash TP, et al. (2005) Mitochondrial dysfunction resulting from loss of cytochrome c impairs cellular oxygen sensing and hypoxic HIF-alpha activation. Cell Metab 1: 393-9.

Mao EF, Lane L, Lee J, Miller JH (1997) Proliferation of mutators in a cell population. J Bacteriol 179: 417–22.

Mao Z, Bozzella M, Seluanov A, Gorbunova V (2008) DNA repair by nonhomologous end joining and homologous recombination during cell cycle in human cells. Cell Cycle 7: 2902–6.

Mappes J, Mappes T, Lappalainen T (1997) Unequal maternal investment in offspring quality in relation to predation risk. Evol Ecol 11: 237-43.

Mappes T, Koskela E (2004) Genetic basis of the trade-off between offspring number and quality in the bank vole. Evolution 58: 645–50.

Mappes T, Grapputo A, Hakkarainen H, Huhta E, Koskela E, et al. (2008) Island selection on mammalian life-histories: genetic differentiation in offspring size. BMC Evol Biol 8: 296.

Marais G (2003) Biased gene conversion: implications for genome and sex evolution. Trends Genet 19: 330-8.

Marais G, Duret L (2001) Synonymous codon usage, accuracy of translation, and gene length in Caenorhabditis elegans. J Mol Evol 52: 275–80.

Marais G, Charlesworth B (2003) Genome evolution: recombination speeds up adaptive evolution. Curr Biol 13: R68-70.

Marchenko ND, Hanel W, Li D, Becker K, Reich N, Moll UM (2010) Stress-mediated nuclear stabilization of p53 is regulated by ubiquitination and importin-alpha3 binding. Cell Death Differ 17: 255-67.

Marchetti F, Essers J, Kanaar R, Wyrobek AJ (2007) Disruption of maternal DNA repair increases sperm-derived chromosomal aberrations. Proc Natl Acad Sci USA 104: 17725-9.

Marchetti F, Wyrobek AJ (2008) DNA repair decline during mouse spermiogenesis results in the accumulation of heritable DNA damage. DNA Repair (Amst) 7: 572-81.

Marchington DR, Hartshorne GM, Barlow D, Poulton J (1997) Homopolymeric tract heteroplasmy in mtDNA from tissues and single oocytes: support for a genetic bottleneck. Am J Hum Genet 60: 408–16.

Marchington DR, Macaulay V, Hartshorne GM, Barlow D, Poulton J (1998) Evidence from human oocytes for a genetic bottleneck in an mtDNA disease. Am J Hum Genet 63: 769–75.

Marchissio MJ, Francés DE, Carnovale CE, Marinelli RA (2012) Mitochondrial aquaporin-8 knockdown in human hepatoma HepG2 cells causes ROS-induced mitochondrial depolarization and loss of viability. Toxicol Appl Pharmacol 264: 246-54.

Marcon E, Moens PB (2005) The evolution of meiosis: recruitment and modification of somatic DNA-repair proteins. BioEssays 27: 795–808.

Marcon E, Babak T, Chua G, Hughes T, Moens PB (2008) miRNA and piRNA localization in the male mammalian meiotic nucleus. Chromosome Res 16: 243-60.

Marcon L, Boissonneault G (2004) Transient DNA strand breaks during mouse and human spermiogenesis new insights in stage specificity and link to chromatin remodeling. Biol Reprod 70: 910–8.

Marcus NH, Lutz R, Burnett W, Cable P (1994) Age, viability, and vertical distribution of zooplankton resting eggs from an anoxic basin: evidence of an egg bank. Limnol Oceanogr 39: 154–8.

Mares D, Romagnoli C, Rubini M, Fasulo MP (1993) Cytological characterization of a giant strain of Euglena gracilis obtained from dark-starved cultures. Botanica Acta 106: 473-9.

Maresca B, Schwartz JH (2006) Sudden origins: a general mechanism of evolution based on stress protein concentration and rapid environmental change. Anat Rec B New Anat 289: 38–46.

Margulis L, Sagan D (1986) Origins of Sex: Three Billion Years of Genetic Recombination. New Haven, CT: Yale University Press.

Marin S, Chiang K, Bassilian S, Lee WN, Boros LG, et al. (2003) Metabolic strategy of boar spermatozoa revealed by a metabolomics characterization. FEBS Lett 554: 342–6.

Mariner PD, Walters RD, Espinoza CA, Drullinger LF, Wagner SD, et al. (2008) Human Alu RNA is a modular transacting repressor of mRNA transcription during heat shock. Mol Cell 29: 499–509.

Marion RM, Strati K, Li H, Murga M, Blanco R, et al. (2009) A p53-mediated DNA damage response limits reprogramming to ensure iPS cell genomic integrity. Nature 460: 1149–53.

Mark Welch DB, Meselson MS (2000) Evidence for the evolution of bdelloid rotifers without sexual recombination or genetic exchange. Science 288: 1211–5.

Welch DBM, Meselson MS (2001) Rates of nucleotide substitution in sexual and anciently asexual rotifers. Proc Natl Acad Sci USA 98: 6720-4.

Mark Welch DB, Meselson M (2003) Oocyte nuclear DNA and GC proportion in rotifers of an anciently asexual Class Bdelloidea. Biol J Linn Soc 79:85-91.

Mark Welch DB, Cummings MP, Hillis DM, Meselson M (2004a) Divergent gene copies in the asexual class Bdelloidea (Rotifera) separated before the bdelloid radiation of within bdelloid families. Proc Natl Acad Sci USA 101: 1622-5.

Mark Welch JL, Mark Welch DB, Meselson M (2004b) Cytogenetic evidence for asexual evolution of bdelloid rotifers. Proc Natl Acad Sci USA 101: 1618-21.

Mark Welch DB, Mark Welch JL,Meselson M (2008) Evidence for degenerate tetraploidy in bdelloid rotifers. Proc NatlAcad Sci USA 105: 5145–9.

Mark Welch DB, Ricci C, Meselson M (2009) Bdelloid rotifers: progress in understanding the success of an evolutionary scandal. In: Schön I, Martens K, van Dijk P, eds. Lost sex. New York, NY: Springer. pp 259–279.

Marletta MA (1988) Mammalian synthesis of nitrite, nitrate, and N-nitrosating agents. Chem Res Toxicol 1: 249-57.

Marnett LJ (2000) Oxyradicals and DNA damage. Carcinogenesis 21: 361-70.

Marnett LJ, Plastaras JP (2001) Endogenous DNA damage and mutation. Trends Genet 17: 214-21.

Margueron R, Reinberg D (2011) The Polycomb complex PRC2 and its mark in life. Nature 469: 343-9.

Marrow P, Dieckmann U, Law R (1996) Evolutionary dynamics of predator-prey systems: an ecological perspective. J Math Biol 34:556-78.

Marshall AJ, Corbet PS (1959) The breeding biology of equatorial vertebrates: reproduction of the bat Chaerephon hindei Thomas at latitude 0°26'N. Proc Zool Soc Lond 132: 607-16.

Marshall CR (2006) Explaining the Cambrian “Explosion” of animals. Annu Rev Earth Planet Sci 34: 355–84.

Marshall DJ, Bolton TF, Keough MJ (2003) Offspring size affects the post-metamorphic performance of a colonial marine invertebrate. Ecology 84: 3131-7.

Marshall DJ, Cook CN, Emlet RB (2006) Offspring size effects mediate competitive interactions in a colonial marine invertebrate. Ecology 87: 214-25.

Marshall DJ, Keough MJ (2007) The evolutionary ecology of offspring size in marine invertebrates. Adv Mar Biol 53: 1-60.

Marshall DJ, Bonduriansky R, Bussière LF (2008) Offspring size variation within broods as a bet-hedging strategy in unpredictable environments. Ecology 89: 2506-17.

Marshall DL, Ellstrand NC (1986) Sexual selection in Raphanus sativus: experimental data on nonrandom fertilization, maternal choice, and consequences of multiple paternity. Am Nat 127:446–61.

Marshall DL, Ellstrand NC (1988) Effective mate choice in wild radish: evidence for selective seed abortion and its mechanism. Am Nat 131: 739–56.

Marshall E (2002) Plant genetics. A hidden Arabidopsis emerges under stress. Science 296: 1218.

Marshall FHA (1911) The male generative cycle in the hedgehog; with experiments on the functional correlation between the essential & the accessory sexual organs. J Physiol Lond 43: 247-60.

Marteinsson VT, Reysenbach A-L, Birrien J-L, Prieur D (1999) A stress protein is induced in the deep-sea barophilic hyperthermophile Thermococcus barophilus when grown under atmospheric pressure. Extremophiles 3: 277-82.

Martens K, ed. (1998) Sex and parthenogenesis. Evolutionary ecology of reproductive modes in non-marine ostracods. Leiden, The Netherlands: Backhuys Publishers.

Martens K, Rossetti G, Horne DJ (2003) How ancient are ancient asexuals? Proc R Soc Lond B 270: 723–9.

Martens K, Schön I (2008) Opinion: ancient asexuals: darwinulids not exposed. Nature 453: 587.

Marti A, Jaggi R, Vallan C, Ritter PM, Baltzer A, et al. (1999) Physiological apoptosis in hormone-dependent tissues: involvement of caspases. Cell Death Differ 6: 1190-200.

Marti HH, Risau W (1998) Systemic hypoxia changes the organ-specific distribution of vascular endothelial growth factors and its receptors. Proc Natl Acad Sci USA 95: 15809–14.

Marti HH, Katschinski DM, Wagner KF, Schaffer L, Stier B, Wenger RH (2002) Isoform-specific expression of hypoxia-inducible factor-1alpha during the late stages of mouse spermiogenesis. Mol Endocrinol 16: 234–43.

Martienssen R (2008) Great leap forward? Transposable elements, small interfering RNA and adaptive Lamarckian evolution. New Phytol 179: 570–2.

Martienssen RA, Colot V (2001) DNA methylation and epigenetic inheritance in plants and filamentous fungi. Science 293: 1070-4.

Martin AP (1999) Substitution rates of organelle and nuclear genes in sharks: implicating metabolic rate (again). Mol Biol Evol 16: 996–1002.

Martin AP, Naylor G, Palumbi SR (1992) Rates of mitochondrial DNA evolution in sharks are slow compared with mammals. Nature 357: 153–5.

Martin AP, Palumbi SR (1993) Body size, metabolic rate, generation time, and the molecular clock. Proc Natl Acad Sci USA 90: 4087–91.

Martin E (1991) The egg and the sperm: how science has constructed a romance based on stereotypical male-female roles. Signs 16: 485-501.

Martin G, Elena SF, Lenormand T (2007) Distributions of epistasis in microbes fit predictions from a fitness landscape model. Nat Genet 39: 555–60.

Martin JL, McMillan FM (2002) SAM (dependent) I AM: the S-adenosylmethionine-dependent methyltransferase fold. Curr Opin Struct Biol 12: 783–93.

Martin KR, Barrett JC (2002) Reactive oxygen species as double-edged swords in cellular processes: low-dose cell signaling versus high-dose toxicity. Hum Exp Toxicol 21: 71-5.

Martin OY, Hosken DJ (2003) The evolution of reproductive isolation through sexual conflict. Nature 423: 979–82.

Martin PA, Reimers TJ, Lodge JR, Dziuk PJ (1974) The effect of ratios and numbers of spermatozoa mixed from two males on proportions of offspring. J Reprod Fert 39: 251–8.

Martin R, Santamaria L, Fraile B, Paniagua R, Polak JM (1995) Ultrastructural localization of PGP 9.5 and ubiquitin immunoreactivities in rat ductus epididymidis epithelium. Histochem J 27: 431–9.

Martin SJ, Reutelingsperger CP, McGahon AJ, Rader JA, van Schie RC, et al. (1995) Early redistribution of plasma membrane phosphatidylserine is a general feature of apoptosis regardless of the initiating stimulus: inhibition by overexpression of Bcl-2 and Abl. J Exp Med 182: 1545–56.

Martindale JL, Holbrook NJ (2002) Cellular response to oxidative stress: signaling for suicide and survival. J Cell Physiol 192: 1-15.

Martindale MQ (2005) The evolution of metazoan axial properties. Nat Rev Genet 6: 917-27.

Martindale MQ, Pang K, Finnerty JR (2004) Investigating the origins of triploblasty: 'mesodermal' gene expression in a diploblastic animal, the sea anemone Nematostella vectensis (phylum, Cnidaria; class, Anthozoa). Development 131: 2463-74.

Martínez G, Slotkin RK (2012) Developmental relaxation of transposable element silencing in plants: functional or byproduct? Curr Opin Plant Biol 15: 496-502.

Martínez MA, Dopazo J, Hernández J, Mateu MG, Sobrino F, et al. (1992) Evolution of the capsid protein genes of foot-and-mouth disease virus: antigenic variation without accumulation of amino acid substitutions over six decades. J Virol 66: 3557-65.

Martínez-Heredia J, Estanyol JM, Ballescà JL, Oliva R (2006) Proteomic identification of human sperm proteins. Proteomics 6: 4356-69.

Martins JSS (2000) Simulated coevolution in a mutating ecology. Phys Rev 61: R2212-R2215.

Martins MJF, Vandekerkhove J, Namiotko T (2008) Environmental stability and the distribution of the sexes: insights from life history experiments with the geographic parthenogen Eucypris virens (Crustacea: Ostracoda). Oikos 117: 829–36.

Marumo T, Schini-Kerth VB, Brandes RP, Busse R (1998) Glucocorticoids inhibit superoxide anion production and p22phox mRNA expression in human aortic smooth muscle cells. Hypertension 32: 1083–8.

Marusyk A, Porter CC, Zaberezhnyy V, DeGregori J (2010) Irradiation selects for p53-deficient hematopoietic progenitors. PLoS Biol 8: e1000324.

Maruyama K, Hartl DL (1991) Evidence for interspecific transfer of the transposable element mariner between Drosophila and Zaprionus. J Mol Evol 33: 514–24.

Maruyama T, Crow JF (1975) Heterozygous effects of x-ray induced mutations on viability of Drosophila melanogaster. Mutat Res 27:241-8.

Marx CJ (2013) Can you sequence ecology? Metagenomics of adaptive diversification. PLoS Biol 11: e1001487.

Marx JL (1988) Cytokines are two-edged swords in disease. Science 239: 257-8.

Maryanovich M, Gross A (2013) A ROS rheostat for cell fate regulation. Trends Cell Biol 23: 129-34.

Más A, López-Galíndez C, Cacho I, Gómez J, Martínez MA (2010) Unfinished stories on viral quasispecies and Darwinian views of evolution. J Mol Biol 397: 865-77.

Masel J (2004) Genetic assimilation can occur in the absence of selection for the assimilating phenotype, suggesting a role for the canalization heuristic. J Evol Biol 17: 1106–10.

Masel J (2006) Cryptic genetic variation is enriched for potential adaptations. Genetics 172: 1985–91.

Masel J, Bergman A (2003) The evolution of the evolvability properties of the yeast prion [PSI+]. Evolution 57: 1498–1512.

Masel J, King OD, Maughan H (2007) The loss of adaptive plasticity during long periods of environmental stasis. Am Nat 169: 38–46.

Masel J, Siegal ML (2009) Robustness: mechanisms and consequences. Trends Genet 25: 395–403.

Masel J, Trotter MV (2010) Robustness and evolvability. Trends Genet 26: 406–14.

Maside X, Bartolome C, Charlesworth B (2002) S-element insertions are associated with the evolution of the HSP70 genes in Drosophila melanogaster. Curr Biol 12: 1686-91.

Mason G, Noris E, Lanteri S, Acquadro A, Accotto GP, Portis E (2008) Potentiality of methylation-sensitive amplification polymorphism (MSAP) in identifying genes involved in tomato response to tomato yellow leaf curl Sardinia virus. Plant Mol Biol Rep 26: 156–73.

Mason MG, Nicholls P, Wilson MT, Cooper CE (2006) Nitric oxide inhibition of respiration involves both competitive (heme) and noncompetitive (copper) binding to cytochrome c oxidase. Proc Natl Acad Sci USA 103: 708–13.

Masoro EJ (2005) Overview of caloric restriction and ageing. Mech Ageing Dev 126:913–22.

Massicotte L, Coenen K, Mourot M, Sirard MA (2006) Maternal housekeeping proteins translated during bovine oocyte maturation and early embryo development. Proteomics 6: 3811–20.

Massicotte R, Whitelaw E, Angers B (2011) DNA methylation: A source of random variation in natural populations. Epigenetics 6: 421-7.

Massicotte R, Angers B (2012) General-purpose genotype or how epigenetics extend the flexibility of a genotype. Genet Res Int 2012: 317175.

Massie ED, Gomes WR, VanDemark NL (1969) Oxygen tension in testes of normal and cryptorchid rats. J Reprod Fertil 19: 559-61.

Masterson J (1994) Stomatal size in fossil plants: evidence for polyploidy in the majority of angiosperms. Science 264: 421-4.

Mastorakos G, Pavlatou MG, Mizamtsidi M (2006) The hypothalamic-pituitary-adrenal and the hypothalamic- pituitary-gonadal axes interplay. Pediatr Endocrinol Rev 3 Suppl 1: 172-81.

Mateo Leach I, Pannebakker BA, Schneider MV, Driessen G, van de Zande L, Beukeboom LW (2009) Thelytoky in hymenoptera with Venturia canescens and Leptopilina clavipes as case studies. In: Schön I, Martens K, van Dijk P, eds. Lost sex. Amsterdam, The Netherlands: Springer. pp 347-375.

Matés JM, Segura JA, Alonso FJ, Márquez J (2008) Intracellular redox status and oxidative stress: implications for cell proliferation, apoptosis, and carcinogenesis. Arch Toxicol 82: 273-99.

Mathai JC, Sitaramam V (1994) Stretch sensitivity of transmembrane mobility of hydrogen peroxide through voids in the bilayer. Role of cardiolipin. J Biol Chem 269: 17784-93.

Mather KA, Jorm AF, Parslow RA, Christensen H (2011) Is telomere length a biomarker of aging? A review. J Gerontol A Biol Sci Med Sci 66: 202–13.

Mathew A, Mathur SK, Jolly C, Fox SG, Kim S, Morimoto RI (2001) Stress-specific activation and repression of heat shock factors 1 and 2. Mol Cell Biol 21: 7163-71.

Mathews MB, Sonenberg N, Hershey JWB (2000) Origins and principles of translational control. In: Sonenberg N, Hershey JWB, Mathews MB, eds. Translational Control of Gene Expression. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press. pp 1–32.

Mathieu O, Bender J (2004) RNA-directed DNA methylation. J Cell Sci 117: 4881–8.

Mathur PP, Francispillai M, Vaithinathan S, Agarwal A (2011) NF-κB in male reproduction: a boon or a bane?Open Reprod Sci J 3: 85-91.

Matic I, Radman M, Taddei F, Picard B, Doit C, et al. (1997) Highly variable mutation rates in commensal and pathogenic Escherichia coli. Science 277: 1833–4.

Matoba S, Kang JG, Patino WD, Wragg A, Boehm M, et al. (2006) p53 regulates mitochondrial respiration. Science 312: 1650–3.

Matoba S, Hiramatsu R, Kanai-Azuma M, Tsunekawa N, Harikae K, et al. (2008) Establishment of testis-specific SOX9 activation requires high glucose metabolism in mouse sex differentiation. Dev Biol 324: 76–87.

Matova N, Cooley L (2001) Comparative aspects of animal oogenesis. Dev Biol 16: 1–30.

Matsui Y (1998) Regulation of germ cell death in mammalian gonads. APMIS 106: 147-8.

Matsumoto M, Fujimoto H (1990) Cloning of a hsp70-related gene expressed in mouse spermatids. Biochem Biophys Res Commun 166: 43–9.

Matsunaga E, Minoda K, Sasaki MS (1990) Parental age and seasonal variation in the births of children with sporadic retinoblastoma: a mutation-epidemiologic study. Hum Genet 84: 155-8.

Matsuura ET, Chigusa SI, Niki Y (1989) Induction of mitochondrial DNA heteroplasmy by intra- and interspecific transplantation of germ plasm in Drosophila. Genetics 122: 663-7.

Matsuura ET, Niki Y, Chigusa SI (1991) Selective transmission of mitochondrial DNA in heteroplasmic lines for intra- and interspecific combinations in Drosophila melanogaster. Jpn J Genet 66: 197-207.

Matta SL, Vilela DA, Godinho HP, França LR (2002) The goitrogen 6-n-propyl-2-thiouracil (PTU) given during testis development increases Sertoli and germ cell numbers per cyst in fish: the tilapia (Oreochromis niloticus) model. Endocrinology 143: 970-8.

Mattews G, Goodwin TJD, Butler MI, Berryman TA, Poulter RTM (1997) pCal, a highly unusual Ty1/copia retrotransposon from the pathogenic yeast Candida albicans. J Bact 179: 7118–28.

Matthews SB, Waud JP, Roberts AG, Campbell AK (2005) Systemic lactose intolerance: a new perspective on an old problem. Postgrad Med J 81: 167–73.

Mattiasson G, Sullivan PG (2006) The emerging functions of UCP2 in health, disease, and therapeutics. Antioxid Redox Signal 8: 1–38.

Mattick JS (2012) Rocking the foundations of molecular genetics. Proc Natl Acad Sci USA 109: 16400-1.

Mattick JS, Amaral PP, Dinger ME, Mercer TR, Mehler MF (2009) RNA regulation of epigenetic processes. Bioessays 31: 51–9.

Mattila TM, Bokma F (2008) Extant mammal body masses suggest punctuated equilibrium. Proc R Soc B 275: 2195–9.

Matzke MA, Mette MF, Matzke AJ (2000) Transgene silencing by the host genome defense: implications for the evolution of epigenetic control mechanisms in plants and vertebrates. Plant Mol Biol 43: 401–15.

Matzke M, Aufsatz W, Kanno T, Daxinger L, Papp I, et al. (2004) Genetic analysis of RNA-mediated transcriptional gene silencing. Biochim Biophys Acta 1677: 129–41.

Matzke MA, Birchler JA (2005) RNAi-mediated pathways in the nucleus. Nat Rev Genet 6: 24–35.

Matzuk MM, Dionne L, Guo Q, Kumar TR, Lebovitz RM (1998) Ovarian function in superoxide dismutase 1 and 2 knockout mice. Endocrinology 139: 4008–11.

Matzuk MM, Burns KH, Viveiros MM, Eppig JJ (2002) Intercellular communication in the mammalian ovary: oocytes carry the conversation. Science 296: 2178–80.

Matsuzawa A, Ichijo H (2008) Redox control of cell fate by MAP kinase: physiological roles of ASK1-MAP kinase pathway in stress signaling. Biochim Biophys Acta 1780: 1325–36.

Mauck RA, Huntingdon CE, Grubb TC Jr (2004) Age-specific reproductive success: evidence for the selection hypothesis. Evolution 58: 880–5.

Mauduit C, Chauvin MA, Hartman DJ, Revol A, Morera AM, Benahmed M (1992) Interleukin-1 alpha as potent inhibitor of gonadotropin action in porcine Leydig cells: site(s) of action. Biol Reprod 46: 1119-26.

Mauduit C, Besset V, Caussanel V, Benahmed M (1996) Tumor necrosis factor α receptor p55 is under hormonal (follicle-stimulating hormone) control in testicular Sertoli cells. Biochem Biophys Res Commun 224: 631–7.

Mauduit C, Gasnier F, Rey C, Chauvin MA, Stocco DM, et al. (1998) Tumor necrosis factor-α inhibits Leydig cell steroidogenesis through a decrease in steroidogenic acute regulatory protein expression. Endocrinology 139: 2863–8.

Mauduit C, Hamamah S, Benahmed M (1999) Stem cell factor/c-kit system in spermatogenesis. Hum Reprod Update 5: 535-45.

Maumus F, Allen AE, Mhiri C, Hu H, Jabbari K, et al. (2009) Potential impact of stress activated retrotransposons on genome evolution in a marine diatom. BMC Genomics 10: 624.

Max B (1992) This and that: hair pigments, the hypoxic basis of life and the Virgilian journey of the spermatozoon. Trends Pharmacol Sci 13: 272–6.

Maxwell PH, Wiesener MS, Chang GW, Clifford SC, Vaux EC, et al. (1999) The tumour suppressor protein VHL targets hypoxia-inducible factors for oxygen-dependent proteolysis. Nature 399: 271–5.

May RM (1974) Ecosystem patterns in randomly fluctuating environments. In: Rosen R, Snell F, eds. Progress in Theoretical Biology. New York, NY: Academic Press. pp 1-50.

May RM, Anderson RM (1983) Epidemiology and genetics in the coevolution of parasites and hosts. Proc R Soc Lond B Biol Sci 219: 281–313.

Mayer I, Bornestaf C, Borg B (1997) Melatonin in non-mammalian vertebrates: physiological role in reproduction? Comp Biochem Physiol 118A: 515-31.

Mayer W, Niveleau A, Walter J, Fundele R, Haaf T (2000) Demethylation of the zygotic paternal genome. Nature 403: 501–2.

Mayerhofer A (1996) Leydig cell regulation by catecholamines and neuroendocrine messengers. In: Payne A, Hardy M, Russel L, eds. The Leydig cell. Vienna, IL: Cache River Press. pp 407-418.

Mayerhofer A, Amador AG, Steger RW, Bartke A (1990) Testicular function after local injection of 6-hydroxydopamine or norepinephrine in the golden hamster (Mesocricetus auratus). J Androl 11: 301-11.

Mayley G (1997) Guiding or hiding: explorations into the effects of learning on the rate of evolution. In: Husbands P, Harvey I, eds. Proceedings of the Fourth European Conference on Artificial Life. Cambridge, MA: Bradford/MIT Press. pp 135–144.

Maynard Smith J (1964) Group selection and kin selection: a rejoinder. Nature 201: 1145-7.

Maynard Smith J (1968) Evolution in sexual and asexual populations. Am Nat 102: 469–73.

Maynard Smith J (1971a) The origin and maintenance of sex. In: Williams GC, ed. Group selection. Chicago, IL: Aldine-Atherton. pp 163–175.

Maynard Smith J (1971b) What use is sex? J Theor Biol 30: 319–35.

Maynard Smith J (1976a) A short-term advantage for sex and recombination through sib-competition. J Theor Biol 63: 245-58.

Maynard-Smith J (1976b) Group selection. Q Rev Biol 5l: 277-83.

Maynard Smith J (1978a) The evolution of sex. London, UK: Cambridge University Press.

Maynard Smith J (1978b) Optimization theory in evolution. Annu Rev Ecol Syst 9: 31-56.

Maynard Smith J (1982) Evolution and the Theory of Games. Cambridge, UK: Cambridge University Press.

Maynard Smith J (1986) Contemplating life without sex. Nature 324: 300–1.

Maynard Smith J (1988) Evolutionary progress and levels of selection. In: Nitecki MH, ed. Evolutionary Progress. Chicago, IL: University of Chicago Press. pp 219–230.

Maynard Smith J (1989) Did Darwin get it right? New York, NY: Chapman & Hall.

Maynard Smith J, Price GR (1973) The logic of animal conflict. Nature 246: 15-8.

Maynard Smith J, Haigh J (1974) The hitch-hiking effect of a favourable gene. Genet Res Camb 23: 23–35.

Maynard Smith J, Brookfield JFY (1983) Models of evolution [and discussion]. Proc R Soc Lond Ser B Biol Sci 219: 315-25.

Maynard Smith J, Szathmáry E (1995) The major transitions in evolution. Oxford, UK: W.H. Freeman/Spektrum.

Maynard Smith J, Smith NH (1996) Synonymous nucleotide divergence: What is saturation? Genetics 142: 1033–6.

Mayr E (1961) Cause and effect in biology. Science 134: 1501-6.

Mayr E (1980) Some thoughts on the history of the evolutionary synthesis. In: Mayr E, Provine W, eds. The evolutionary synthesis. Cambridge, MA: Harvard University Press. pp 1–48.

Mayr E (1982) The growth of biological thought. Cambridge, MA: Belknap Press of Harvard University Press.

Mayr EW (1992) The idea of teleology. J Hist Ideas 53: 117–35.

Mayr E (2001) What Evolution Is. New York, NY: Basic Books.

Mayrose I, Doron-Faigenboim A, Bacharach E, Pupko T (2007) Towards realistic codon models: among site variability and dependency of synonymous and non-synonymous rates. Bioinformatics 23: i319-27.

Mays HL, Hill GE (2004) Choosing mates: good genes versus genes that are a good fit. Trends Ecol Evol 19: 554–9.

Mayshar Y, Ben-David U, Lavon N, Biancotti JC, Yakir B, et al. (2010) Identification and classification of chromosomal aberrations in human induced pluripotent stem cells. Cell Stem Cell 7: 521–31.

Mazin AL (2009) Suicidal function of DNA methylation in age-related genome disintegration. Ageing Res Rev 8:314-27.

McBride RC, Ogbunugafor CB, Turner PE (2008) Robustness promotes evolvability of thermotolerance in an RNA virus. BMC Evol Biol 8: 231.

McBride TJ, Preston BD, Loeb LA (1991) Mutagenic spectrum resulting from DNA damage by oxygen radicals. Biochemistry 30: 207-13.

McCain CM (2009) Global analysis of bird elevational diversity. Global Ecol Biogeogr 18: 346–60.

McCall K (2004) Eggs over easy: cell death in the Drosophila ovary. Dev Biol 274: 3-14.

McCall K, Steller H (1998) Requirement for DCP-1 caspase during Drosophila oogenesis. Science 279: 230–4.

McCarrey JR, Kumari M, Aivaliotis MJ, Wang Z, Zhang P, et al. (1996) Analysis of the cDNA and encoded protein of the human testis-specific PGK-2 gene. Dev Genet 19: 321–32.

McCarter J, Bartlett B, Dang T, Schedl T (1999) On the control of oocyte meiotic maturation and ovulation in Caenorhabditis elegans. Dev Biol 205: 111–28.

McCauley E, Murdoch WW (1987) Cyclic and stable populations: plankton as paradigm. Am Nat 129: 97–121.

McClellan AJ, Tam S, Kaganovich D, Frydman J (2005) Protein quality control: chaperones culling corrupt conformations. Nat Cell Biol 7: 736-41.

McClellan AJ, Xia Y, Deutschbauer AM, Davis RW, Gerstein M, Frydman J (2007) Diverse cellular functions of the Hsp90 molecular chaperone uncovered using systems approaches. Cell 131: 121-35.

McClellan KA, Gosden R, Taketo T (2003) Continuous loss of oocytes throughout meiotic prophase in the normal mouse ovary. Dev Biol 258: 334-48.

McClintock B (1951) Chromosome organization and genic expression. Cold Spring Harb Symp Quant Biol 16: 13–47.

McClintock B (1984) The significance of responses of the genome to challenge. Science 26: 792-801.

McClusky LM (2005) Stage and season effects on cell cycle and apoptotic activities of germ cells and Sertoli cells during spermatogenesis in the spiny dogfish (Squalus acanthias). Reproduction 129: 89-102.

McClusky LM (2011) Testicular degeneration during spermatogenesis in the blue shark, Prionace glauca: nonconformity with expression as seen in the diametric testes of other carcharhinids. J Morphol 272: 938-48.

McClusky LM, Patrick S, Barnhoorn IE, van Dyk JC, de Jager C, Bornman MS (2009) Immunohistochemical study of nuclear changes associated with male germ cell death and spermiogenesis. J Mol Histol 40: 287-99.

McConnell NA, Yunus RS, Gross SA, Bost KL, Clemens MG, Hughes FM Jr (2002) Water permeability of an ovarian antral follicle is predominantly transcellular and mediated by aquaporins. Endocrinology 143: 2905-12.

McCord AM, Jamal M, Shankavaram UT, Lang FF, Camphausen K, Tofilon PJ (2009) Physiologic oxygen concentration enhances the stem-like properties of CD133+ human glioblastoma cells in vitro. Mol Cancer Res 7: 489–97.

McCormick MI, Hoey AS (2004) Larval growth history determines juvenile growth and survival in a tropical marine fish. Oikos 106: 225-42.

McCormick MI, Meekan MG (2007) Social facilitation of selective mortality. Ecology 88: 1562–70.

McCormick MI, Gagliano M (2009) Carry-over effects-the importance of a good start. In: Proceedings of the 11th International Coral Reef Symposium. Florida, USA: Nova South Eastern University. pp 305-310.

McCoubrey WK Jr, Maines MD (1994) The structure, organization and differential expression of the gene encoding rat heme oxygenase-2. Gene 139: 155–161.

McCoubrey WK Jr, Eke B, Maines MD (1995) Multiple transcripts encoding heme oxygenase-2 in rat testis: developmental and cell-specific regulation of transcripts and protein. Biol Reprod 53: 1330–8.

McCracken RD (1971) Lactase deficiency: an example of dietary evolution. Curr Anthropol 12: 479-515.

McCue AD, Nuthikattu S, Reeder SH, Slotkin RK (2012) Gene expression and stress response mediated by the epigenetic regulation of a transposable element small RNA. PLoS Genet 8: e1002474.

McCue AD, Slotkin RK (2012) Transposable element small RNAs as regulators of gene expression. Trends Genet 28: 616–23.

McCue KA, Holtsford TP (1998) Seed bank influences on genetic diversity in the rare annual Clarkia springvillensis (Onagraceae). Am J Bot 85: 30-6.

McCusker C, Gardiner DM (2010) The axolotl model for regeneration and aging research: a mini-review. Gerontology 57: 565–71.

McDaniel CD, Bramwell RK, Wilson JL, Howrath BJr (1995) Fertility of male and female broiler breeders following exposure to elevated ambient temperatures. Poult Sci 74: 1029–38.

McDonald JF (1987) The potential evolutionary significance of retroviral-like transposable elements in peripheral populations. In: Fontdevila A, ed. Evolutionary biology of transient unstable populations. Berlin, Germany: Springer. pp 190-205.

McDonald JF (1998) Transposable elements, gene silencing and macroevolution. Trends Ecol Evol 13: 94–5.

McDonald JF, Matzke MA, Matzke AJ (2005) Host defenses to transposable elements and the evolution of genomic imprinting. Cytogenet Genome Res 110: 242–9.

McDonald JH, Kreitman M (1991) Adaptive protein evolution at the Adh locus in Drosophila. Nature 351: 652-4.

McDonald JP, Rapic-Otrin V, Epstein JA, Broughton BC, Wang X, et al. (1999) Novel human and mouse homologs of Saccharomyces cerevisiae DNA polymerase eta. Genomics 60: 20-30.

McDuffee AT, Senisterra G, Huntley S, Lepock JR, Sekhar KR, et al. (1997) Proteins containing non-native disulfide bonds generated by oxidative stress can act as signals for the induction of the heat shock response. J Cell Physiol 171: 143-51.

McEwan DL, Weisman AS, Hunter CP (2012) Uptake of extracellular double-stranded RNA by SID-2. Mol Cell 47: 746-54.

McEwen BS (1980) Binding and metabolism of sex steroids by the hypothalamic-pituitary unit: physiological implications. Annu Rev Physiol 42: 97-110.

McEwen BS, Wingfield JC (2003) The concept of allostasis in biology and biomedicine. Horm Behav 43: 2–15.

McEwen E, Kedersha N, Song B, Scheuner D, Gilks N, et al. (2005) Heme-regulated inhibitor kinase-mediated phosphorylation of eukaryotic translation initiation factor 2 inhibits translation, induces stress granule formation, and mediates survival upon arsenite exposure. J Biol Chem 280: 16925-33.

McGee EA, Hsueh AJ (2000) Initial and cyclic recruitment of ovarian follicles. Endocr Rev 21: 200–14.

McGee MD, Day N, Graham J, Melov S (2012) cep-1/p53-dependent dysplastic pathology of the aging C. elegans gonad. Aging (Albany NY) 4: 256-69.

McGill CB, Shafer BK, Derr LK, Strathern JN (1993) Recombination initiated by double-strand breaks. Curr Genet 23: 305–14.

McGill CB, Holbeck SL, Strathern JN (1998) The chromosome bias of misincorporations during double-strand break repair is not altered in mismatch repair-defective strains of Saccharomyces cerevisiae. Genetics 148: 1525–33.

Mcgowan PO, Meaney MJ, Szyf M (2008) Diet and the epigenetic (re)programming of phenotypic differences in behavior. Brain Res 1237: 12–24.

McGrady AV (1984) Effects of psychological stress on male reproduction: a review. Arch Androl 13: 1–7.

McGraw JE, Brookfield JF (2006) The interaction between mobile DNAs and their hosts in a fluctuating environment. J Theor Biol 243: 13-23.

McGraw LA, Fiumera AC, Ramakrishnan M, Madhavarapu S, Clark AG, et al. (2007) Larval rearing environment affects several post-copulatory traits in Drosophila melanogaster. Biol Lett 3: 607–10.

McGregor DL (1969) The reproductive potential, life history and parasitism of the freshwater ostracods Darwinula stevensoni (Brady and Robertson). In: Neale JW, ed. The Taxonomy, Morphology and Ecology of Recent Ostracoda. Edinburgh, UK: Oliver and Boyd. pp 194–221.

McGuigan K, Sgrò CM (2009) Evolutionary consequences of cryptic genetic variation. Trends Ecol Evol 24: 305-11.

McGuire NL, Kangas K, Bentley GE (2011) Effects of melatonin on peripheral reproductive function: regulation of testicular GnIH and testosterone.Endocrinology 152: 3461-70.

McInnis SM, Desikan R, Hancock JT, Hiscock SJ (2006) Production of reactive oxygen species and reactive nitrogen species by angiosperm stigmas and pollen: potential signalling crosstalk? New Phytol 172: 221-8.

McIver SC, Stanger SJ, Santarelli DM, Roman SD, Nixon B, et al. (2012) A unique combination of male germ cell miRNAs coordinates gonocyte differentiation. PLoS ONE 7: e35553.

McIvor EI, Polak U, Napierala M (2010) New insights into repeat instability: role of RNA•DNA hybrids. RNA Biol 7: 551-8.

Mckay FE (1971) Behavioral aspects of population dynamics in unisexual-bisexual Poeciliopsis (Pisces: Poeciliidae). Ecology 52: 778-90.

McKee BD (2004) Homologous pairing and chromosome dynamics in meiosis and mitosis. Biochim Biophys Acta 1677:165–80.

McKee BD (2009) Homolog pairing and segregation in Drosophila meiosis. Genome Dyn 5:56-68.

McKee CM, Ye Y, Richburg JH (2006) Testicular germ cell sensitivity to TRAIL-induced apoptosis is dependent upon p53 expression and is synergistically enhanced by DR5 agonistic antibody treatment. Apoptosis 11: 2237–50.

McKeehan WL (1982) Glycolysis, glutaminolysis and cell proliferation. Cell Biol Int Rep 6: 635–50.

McKim KS, Green-Marroquin BL, Sekelsky JJ, Chin G, Steinberg C, Khodosh R, Hawley RS (1998) Meiotic synapsis in the absence of recombination. Science 279: 876-8

McLachlan RI, O'Donnell L, Meachem SJ, Stanton PG, de Kretser DM, et al. (2002) Identification of specific sites of hormonal regulation in spermatogenesis in rats, monkeys, and man. Recent Prog Horm Res 57: 149-79.

McLain DK, Moulton MP, Sanderson JG (1999) Sexual selection and extinction: the fate of plumage-dimorphic and plumage-monomorphic birds introduced onto islands. Evol Ecol Res 1: 549–65.

McLaren A (1999) Signaling for germ cells. Genes Dev 13: 373-6.

McLaren A (2000) Germ and somatic cell lineages in the developing gonad. Mol Cell Endocrinol 163: 3–9.

McLean AR, Rosado MM, Agenes F, Vasconcellos R, Freitas AA (1997) Resource competition as a mechanism for B cell homeostasis. Proc Natl Acad Sci USA 94: 5792-7.

McMillan WO, Jiggins CD, Mallet J (1997) What initiates speciation in passion-vine butterflies? Proc Natl Acad Sci USA 94: 8628–33.

McNally JS, Davis ME, Giddens DP, Saha A, Hwang J, et al. (2003) Role of xanthine oxidoreductase and NAD(P)H oxidase in endothelial superoxide production in response to oscillatory shear stress. Am J Physiol Heart Circ Physiol 285: H2290–H2297.

McNamara JM (1995) Implicit frequency dependence and kin selection in fluctuating environments. Evol Ecol 9: 185-203.

McNamara JM (1998) Phenotypic plasticity in fluctuating environments: consequences of the lack of individual optimization. Behav Ecol 9: 642–8.

McNamara JM, Webb JN, Collins EJ (1995) Dynamic optimization in fluctuating environments. Proc R Soc Lond B 261: 279–84.

McNeill DR, Wilson DM 3rd (2007) A dominant-negative form of the major human abasic endonuclease enhances cellular sensitivity to laboratory and clinical DNA-damaging agents. Mol Cancer Res 5: 61–70.

McPeekMA, Gavrilets S (2006) The evolution of female mating preferences: Differentiation from species with promiscuous males can promote speciation. Evolution 60: 1967–80.

McPhee CP, Robertson A (1970) Effect of suppressing crossing-over on response to selection in Drosophila melanogaster. Genet Res 16: 1–16.

McPherson SM, Longo FJ (1993) Nicking of rat spermatid and spermatozoa DNA: possible involvement of DNA topoisomerase II. Dev Biol 158: 122–30.

McTavish KJ, Jimenez M, Walters KA, Spaliviero J, Groome NP, et al. (2007) Rising follicle-stimulating hormone levels with age accelerate female reproductive failure. Endocrinology 148: 4432-9.

McVean GAT, Charlesworth B (2000) The effects of Hill–Robertson interference between weakly selected mutations on patterns of molecular evolution and variation. Genetics 155: 929–44.

McVean GAT, Myers SR, Hunt S, Deloukas P, Bentley DR, Donnelly P (2004) The fine-scale structure of recombination rate variation in the human genome. Science 304: 581-4.

McVicker G, Gordon D, Davis C, Green P (2009) Widespread genomic signatures of natural selection in hominid evolution. PLoS Genet 5: e1000471.

Meachem SJ, McLachlan RI, de Kretser DM, Robertson DM, Wreford NG (1996) Neonatal exposure of rats to recombinant follicle stimulating hormone increases adult Sertoli and spermatogenic cell numbers. Biol Reprod 54: 36–44.

Meagher S, Penn DJ, Potts WK (2000) Male-male competition magnifies inbreeding depression in wild house mice. Proc Natl Acad Sci USA 97: 3324–9.

Méchali M, Gusse M, Vriz S, Taylor M, Andéol Y, et al. (1988)Proto-oncogenes and embryonic development. Biochimie 70: 895-8.

Medvedev ZA (1981) On the immortality of the germ line: genetic and biochemical mechanisms. A review. Mech Ageing Develop 17: 331–59.

Mee JA, Rowe L (2006) A comparison of parasite loads on asexual and sexual Phoxinus (Pisces: Cyprinidae). Can J Zool-Rev Can Zool 84: 808–16.

Mee JA, Otto SP (2010) Variation in the strength of male mate choice allows long-term coexistence of sperm-dependent asexuals and their sexual hosts. Evolution 64: 2808-19.

Meert JG, Lieberman BS (2008) The Neoproterozoic assembly of Gondwana and its relationship to the Ediacaran-Cambrian radiation. Gondwana Res 14: 5-21.

Mehr IJ, Seifert HS (1998) Differential roles of homologous recombination pathways in Neisseria gonorrhoeae pilin antigenic variation, DNA transformation and DNA repair. Mol Microbiol 30: 697–710.

Meier B, Radeke HH, Selle S, Younes M, Sies H, et al. (1989) Human fibroblasts release reactive oxygen species in response to interleukin-1 or tumour necrosis factor-alpha. Biochem J 263: 539–45.

Meier P, Finch A, Evan G (2000) Apoptosis in development. Nature 407: 796-801.

Meij JJ, Van Bodegom D, Ziem JB, Amankwa J, Polderman AM, et al. (2009) Quality-quantity trade-off of human offspring under adverse environmental conditions. J Evol Biol 22: 1014–23.

Meikar O, Da Ros M, Liljenback H, Toppari J, Kotaja N (2010) Accumulation of piRNAs in the chromatoid bodies purified by a novel isolation protocol. Exp Cell Res 316: 1567–75.

Meikar O, Da Ros M, Korhonen H, Kotaja N (2011) Chromatoid body and small RNAs in male germ cells. Reproduction 142: 195–209.

Meikar O, Da Ros M, Kotaja N (2013) Epigenetic regulation of male germ cell differentiation. In: Kundu TK, ed. Epigenetics: Development and Disease. Subcellular Biochemistry 61. Dordrecht, The Netherlands: Springer Science+Business Media. pp 119-138.

Meiklejohn CD, Hartl DL (2002) A single mode of canalization. Trends Ecol Evol 17: 468–73.

Meiller A, Alvarez S, Drané P, Lallemand C, Blanchard B, et al. (2007) p53-dependent stimulation of redox-related genes in the lymphoid organs of gamma-irradiated–mice – identification of haeme-oxygenase 1 as a direct p53 target gene. Nucleic Acids Res 35: 6924–34.

Meinhardt A, Parvinen M, Bacher M, Aumuller G, Hakovirta H, et al. (1995) Expression of mitochondrial heat shock protein 60 in distinct cell types and defined stages of rat seminiferous epithelium. Biol Reprod 52: 798–807.

Meinhardt A, Wilhelm B, Seitz J (1999) New aspects of spermatogenesis. Expression of mitochondrial marker proteins during spermatogenesis. Hum Reprod Update 5: 108–19.

Meirelles FV, Smith LC (1997) Mitochondrial genotype segregation in a mouse heteroplasmic lineage produced by embryonic karyoplast transplantation. Genetics 145: 445–51.

Meister A, Anderson ME (1983) Glutathione. Annu Rev Biochem 52: 711-60.

Meistertzheim AL, Lejart M, Le Goïc N, Thébault MT (2009) Sex-, gametogenesis, and tidal height-related differences in levels of HSP70 and metallothioneins in the Pacific oyster Crassostrea gigas. Comp Biochem Physiol A Mol Integr Physiol 152: 234-9.

Meistrich ML (1989) Histone and basic nuclear protein transitions in mammalian spermatogenesis. In: Hnilica LS, Stein GS, Stein JL, eds. Histones and other basic nuclear proteins. Boca Raton, FL: CRC Press. pp 165-182.

Meistrich ML (1993) Effects of chemotherapy and radiotherapy on spermatogenesis. Eur Urol 23: 136–41.

Meites J (1990) Aging: hypothalamic catecholamines, neuroendocrine-immune interactions, and dietary restriction. Proc Soc Exp Biol Med 195: 304-11.

Meley D, Spiller DG, White MR, McDowell H, Pizer B, Sée V (2010) p53-mediated delayed NF-κB activity enhances etoposide-induced cell death in medulloblastoma. Cell Death Dis 1: e41.

Melián CJ, Alonso D, Allesina S, Condit RS, Etienne RS (2012) Does sex speed up evolutionary rate and increase biodiversity? PLoS Comput Biol 8: e1002414.

Mello-Filho AC, Meneghini R (1984) In vivo formation of single-strand breaks in DNA by hydrogen peroxide is mediated by the Haber-Weiss reaction. Biochim Biophys Acta 781: 56-63.

Meloni R, Albanese V, Ravassard P, Treilhou F, Mallet J (1998) A tetranucleotide polymorphic microsatellite, located in the first intron of the tyrosine hydroxylase gene, acts as a transcription regulatory element in vitro. Hum Mol Genet 7: 423-8.

Melser C, Klinkhamer PGL (2001) Selective seed abortion increases offspring survival in Cynoglossum officinale (Boraginaceae). Am J Bot 88: 1033–40.

Melzer AL, Koeslag JH (1991) Mutations do not accumulate in asexual isolates capable of growth and extinction—Muller's ratchet reexamined. Evolution 45: 649-55.

Mendell JT (2008) miRiad roles for the miR-17-92 cluster in development and disease. Cell 133: 217-22.

Menegazzi M, Grassi-Zucconi G, Carcerero De Prati A, Ogura T, Poltronieri P, et al. (1991) Differential expression of poly(ADP-ribose) polymerase and DNA polymerase beta in rat tissues. Exp Cell Res 197: 66–74.

Meneghini R (1988) Genotoxicity of active oxygen species in mammalian cells. Mutat Res 195: 215-30.

Meneghini R (1997) Iron homeostasis, oxidative stress, and DNA damage. Free Radic Biol Med 23: 783-92.

Menella MR, Jones R (1980) Properties of spermatozoal superoxide dismutase and lack of involvement of superoxides in metal ion catalyzed lipid peroxidation reactions in semen. Biochem J 191: 289-97.

Menendez S, Camus S, Izpisua Belmonte JC (2010) p53: guardian of reprogramming. Cell Cycle 9: 3887–91.

Meng AX, Jalali F, Cuddihy A, Chan N, Bindra RS, et al. (2005) Hypoxia down-regulates DNA double strand break repair gene expression in prostate cancer cells. Radiother Oncol 76: 168–76.

Meng D, Lv DD, Fang J (2008) Insulin-like growth factor-I induces reactive oxygen species production and cell migration through Nox4 and Rac1 in vascular smooth muscle cells. Cardiovasc Res 80: 299-308.

Meng TC, Fukada T, Tonks NK (2002) Reversible oxidation and inactivation of protein tyrosine phosphatases in vivo. Mol Cell 9: 387-99.

Meng TC, Buckley DA, Galic S, Tiganis T, Tonks NK (2004) Regulation of insulin signaling through reversible oxidation of the protein-tyrosine phosphatases TC45 and PTP1B. J Biol Chem 279: 37716-25.

Menon SG, Sarsour EH, Spitz DR, Higashikubo R, Sturm M, et al. (2003) Redox regulation of the G1 to S phase transition in the mouse embryo fibroblast cell cycle. Cancer Res 63: 2109-17.

Menon SG, Goswami PC (2007) A redox cycle within the cell cycle: ring in the old with the new. Oncogene 26: 1101-9.

Menu F, Roebuck JP, Viala M (2000) Bet-hedging diapause strategies in stochastic environments. Am Nat 155: 724–34.

Mercurio F, Manning AM (1999) NF-kappaB as a primary regulator of the stress response. Oncogene 18: 6163-71.

Merilä J, Crnokrak P (2001) Comparison of genetic differentiation at marker loci and quantitative traits. J Evol Biol 14: 892–903.

Merilä J, Sheldon BC, Kruuk LEB (2001) Explaining stasis: microevolutionary studies in natural populations. Genetica 112-113: 199-222.

Merilä J, Laurila A, Lindgren B (2004) Variation in the degree and costs of adaptive phenotypic plasticity among Rana temporaria populations. J Evol Biol 17: 1132–40.

Merlin C, Mahillon J, Nešvera J, Toussaint A (2000) Gene recruiters and transporters: the modular structure of bacterial mobile elements. In: Thomas CM, ed. The Horizontal Gene Pool. Amsterdam, The Netherlands: Harwood Academic Publishers. pp 301–361.

Merlin F (2010) Evolutionary chance mutation: a defense of the Modern Synthesis consensus view. Philos Theor Biol 2: e103.

Merlo LM, Pepper JW, Reid BJ, Maley CC (2006) Cancer as an evolutionary and ecological process. Nat Rev Cancer 6: 924–35.

Merlo LM, Kosoff RE, Gardiner KL, Maley CC (2011) An in vitro co-culture model of esophageal cells identifies ascorbic acid as a modulator of cell competition. BMC Cancer 11: 461.

Meroni SB, Suburo AM, Cigorraga SB (2000) Interleukin-1beta regulates nitric oxide production and gamma-glutamyl transpeptidase activity in Sertoli cells. J Androl 21: 855–61.

Merrifield M (2011) Radiation-induced deregulation of PiRNA pathway proteins: a possible molecular mechanism underlying transgenerational epigenomic instability. Master of Science Thesis. Lethbridge, Canada: University of Lethbridge.

Mery F, Kawecki TJ (2002) Experimental evolution of learning ability in fruit flies. Proc Natl Acad Sci USA 99: 14274–9.

Mery F, Kawecki TJ (2003) A fitness cost of learning ability in Drosophila melanogaster. Proc R Soc Lond B 270: 2465–9.

Mery F, Kawecki TJ (2005) A cost of long-term memory in Drosophila. Science 308: 1148.

Messier W, Stewart CB (1997) Episodic adaptive evolution of primate lysozymes. Nature 385: 151-4.

Metcalfe NB, Alonso-Alvarez C (2010) Oxidative stress as a life-history constraint: the role of reactive oxygen species in shaping phenotypes from conception to death. Funct Ecol 24: 947–9.

Metchat A, Akerfelt M, Bierkamp C, Delsinne V, Sistonen L, et al. (2009) Mammalian heat shock factor 1 is essential for oocyte meiosis and directly regulates Hsp90alpha expression. J Biol Chem 284: 9521–8.

Mette MF, Aufsatz W, van der Winden J, Matzke MA, Matzke AJ (2000) Transcriptional silencing and promoter methylation triggered by double-stranded RNA. EMBO J 19: 5194–201.

Metz EC, Palumbi SR (1996) Positive selection and sequence rearrangements generate extensive polymorphism in the gamete recognition protein bindin. Mol Biol Evol 13: 397–406.

Metz EC, Robles-Sikisaka R, Vacquier VD (1998) Nonsynonymous substitution in abalone sperm fertilization genes exceeds substitution in introns and mitochondrial DNA. Proc Natl Acad Sci USA 95: 10676–81.

Metzendorf C, Lind MI (2010) Drosophila mitoferrin is essential for male fertility: evidence for a role of mitochondrial iron metabolism during spermatogenesis. BMC Dev Biol 10: 68.

Metzgar D, Wills C (2000) Evidence for adaptive evolution of mutation rates. Cell 101: 581–4.

Metzgar D, Bytof J, Wills C (2000) Selection against frameshift mutations limits microsatellite expansion in coding DNA. Genome Res 10: 72–80.

Meunier L, Siddeek B, Vega A, Lakhdari N, Inoubli L, et al. (2012) Perinatal programming of adult rat germ cell death after exposure to xenoestrogens: role of microRNA miR-29 family in the down-regulation of DNA methyltransferases and Mcl-1. Endocrinology 153: 1936-47.

Mével-Ninio M, Pelisson A, Kinder J, Campos AR, Bucheton A (2007) The flamenco locus controls the gypsy and ZAM retroviruses and is required for Drosophila oogenesis. Genetics 175: 1615–24.

Meyer N, Penn LZ (2008) Reflecting on 25 years with MYC. Nat Rev Cancer 8: 976–90.

Meyer SC (2008) A scientific history – and philosophical defense – of the theory of intelligent design. Religion-Staat-Gesellschaft 7: 203-47.

Meyer-Ficca ML, Scherthan H, Bürkle A, Meyer RG (2005) Poly(ADPribosyl)ation during chromatin remodeling steps in rat spermiogenesis. Chromosoma 114: 67–74.

Meyer-Ficca ML, Lonchar JD, Ihara M, Meistrich ML, Austin CA, Meyer RG (2011) Poly(ADP-ribose) polymerases PARP1 and PARP2 modulate topoisomerase II beta (TOP2B) function during chromatin condensation in mouse spermiogenesis. Biol Reprod 84: 900-9.

Meyers LA, Bull JJ (2002) Fighting change with change: adaptive variation in an uncertain world. Trends Ecol Evol 17: 551–7.

Meylan S, Clobert J (2005) Is corticosterone-mediated phenotype development adaptive? Maternal corticosterone treatment enhances survival in male lizards. Horm Behav 48: 44–52.

Mezger V, Renard JP, Christians E, Morange M (1994) Detection of heat shock element binding activities by gel shift assay during mouse preimplantation development. Dev Biol 165: 627–38.

Mezquita B, Mezquita C and Mezquita J (1998) Marked differences between avian and mammalian testicular cells in the heat shock induction and polyadenylation of Hsp and ubiquitin transcripts. FEBS Lett 436: 382-6.

M’Gonigle LK, Shen JJ, Otto SP (2009) Mutating away from your enemies: The evolution of mutation rate in a host–parasite system. Theor Popul Biol 75: 301–11.

Michael AE, Cooke BA (1994) A working hypothesis for the regulation of steroidogenesis and germ cell development in the gonads by glucocorticoids and 11 beta-hydroxysteroid dehydrogenase (11 beta HSD). Mol Cell Endocrinol 100: 55-63.

Michaelson J (1987) Cell selection in development. Biol Rev Camb Philos Soc 62: 115-39.

Michaelson J (1993) Cellular selection in the genesis of multicellular organization. Lab Invest 69: 136-51.

Michalik KM, Böttcher R, Förstemann K (2012) A small RNA response at DNA ends in Drosophila. Nucleic Acids Res 40: 9596-603.

Michalik P, Uhl G (2005) The male genital system of the cellar spider Pholcus phalangioides (Fuesslin, 1775) (Pholcidae, Araneae): development of spermatozoa and seminal secretion. Front Zool 2: 12.

Michalik P, Huber BA (2006) Spermiogenesis in Psilochorus simony (Berland, 1911) (Pholcidae, Araneae): evidence for considerable within-family variation in sperm structure and development. Zoology 109: 14–25.

Michaud S, Marin R, Westwood JT, Tanguay RM (1997) Cell-specific expression and heat-shock induction of Hsps during spermatogenesis in Drosophila melanogaster. J Cell Sci 110: 1989-97.

Michels J, Johnson PW, Packham G (2005) Mcl-1. Int J Biochem Cell Biol 37: 267–71.

Michod RE (1993) Genetic error, sex, and diploidy. J Heredity 84: 360–71.

Michod RE (1995) Eros and evolution: a natural philosophy of sex. Reading, MA: Addison-Wesley Publishing.

Michod RE (1996) Cooperation and conflict in the evolution of individuality. II. Conflict mediation. Proc R Soc Lond B 263: 813–822.

Michod RE (1999a) Individuality, immortality, and sex. In: Keller L, ed. Levels of selection in evolution. Princeton, NJ: Princeton University Press. pp 53–74.

Michod RE (1999b) Darwinian dynamics: evolutionary transitions in fitness and individuality. Princeton, NJ: Princeton University Press.

Michod RE (2005) On the transfer of fitness from the cell to the multicellular organism. Biol Philos 20: 967–87.

Michod RE (2006) The group covariance effect and fitness trade-offs during evolutionary transitions in individuality. Proc Natl Acad Sci USA 103: 9113–7.

Michod RE (2007) Evolution of individuality during the transition from unicellular to multicellular life. Proc Natl Acad Sci USA 104: 8613–8.

Michod R, Wojciechowski M, Hoelzer M (1988) DNA repair and the evolution of transformation in the bacterium Bacillus subtilis. Genetics 118: 31-9.

Michod RE, Bernstein H, Nedelcu AM (2008) Adaptive value of sex in microbial pathogens. Infect Genet Evol 8: 267–85.

Michor F, Iwasa Y, Komarova NL, Nowak MA (2003) Local regulation of homeostasis favors chromosomal instability. Curr Biol 13: 581–4.

Micke A, Donini B, Maluzynski M (1987) Induced mutations for crop improvement- a review. Trop Agric 64: 259-78.

Miething A (1992) Germ-cell death during prespermatogenesis in the testis of the golden hamster. Cell Tissue Res 267: 583–90.

Migicovsky Z, Kovalchuk I (2012) Epigenetic modifications during angiosperm gametogenesis. Front Plant Sci 3: 20.

Mihaljev T, Drossel B (2009) Evolution of a population of random Boolean networks. Eur Phys J B 67: 259-67.

Mihara M, Erster S, Zaika A, Petrenko O, Chittenden T, et al. (2003) p53 has a direct apoptogenic role at the mitochondria. Mol Cell 11: 577–90.

Mihaylova VT, Bindra RS, Yuan J, Campisi D, Narayanan L, et al. (2003) Decreased expression of the DNA mismatch repair gene Mlh1 under hypoxic stress in mammalian cells. Mol Cell Biol 23:3265-73.

Mihola O, Trachtulec Z, Vlcek C, Schimenti JC, Forejt J (2009) A mouse speciation gene encodes a meiotic histone H3 methyltransferase. Science 323: 373–5.

Mikkelsen R, Wardman P (2003) Biological chemistry of reactive oxygen and nitrogen and radiation-induced signal transduction mechanisms. Oncogene 22: 5734–54.

Mikkelsen TS, Ku M, Jaffe DB, Issac B, Lieberman E, et al. (2007) Genome-wide maps of chromatin state in pluripotent and lineage-committed cells. Nature 448: 553-60.

Milán M, Pérez L, Cohen SM (2002) Short-range cell interactions and cell survival in the Drosophila wing. Dev Cell 2: 797-805.

Milani L, Scali V, Passamonti M (2009) The Clonopsis gallica puzzle: Mendelian species, polyploid parthenogens with karyotype re-diploidization and clonal androgens in Moroccan stick insects. J Zool Syst Evol Res 47: 132-40.

Miles GM, Sancar A (1989) DNA repair. Chem Res Toxicol 2: 197-226.

Miles JR, McDaneld TG, Wiedmann RT, Cushman RA, Echternkamp SE, et al. (2012) MicroRNA expression profile in bovine cumulus-oocyte complexes: possible role of let-7 and miR-106a in the development of bovine oocytes. Anim Reprod Sci 130: 16-26.

Milinski M (2006) The major histocompatibility complex, sexual selection and mate choice. Annu Rev Ecol Syst 37: 159-86.

Milinski M, Bakker TCM (1990) Female sticklebacks use male coloration in mate choice and hence avoid parasitized males. Nature 344: 330-3.

Milinski M, Parker GA (1991) Competition for resources. In: Krebs JR, Davies NB, eds. Behavioural ecology: an evolutionary approach. Oxford, UK: Blackwell. pp 137–168.

Milla S, Wang N, Mandiki SN, Kestemont P (2009) Corticosteroids: Friends or foes of teleost fish reproduction? Comp Biochem Physiol A Mol Integr Physiol 153: 242–51.

Millane RC, Kanska J, Duffy DJ, Seoighe C, Cunningham S, et al. (2011) Induced stem cell neoplasia in a cnidarian by ectopic expression of a POU domain transcription factor. Development 138: 2429–39.

Millar CD, Dodd A, Anderson J, Gibb GC, Ritchie PA, et al. (2008) Mutation and evolutionary rates in Adélie penguins from the Antarctic. PLoS Genet 4: e1000209.

Miller AA, De Silva TM, Jackman KA, Sobey CG (2007) Effect of gender and sex hormones on vascular oxidative stress. Clin Exp Pharmacol Physiol 34: 1037-43.

Miller AC, Blakely WF (1992) Inhibition of glutathione reductase activity by a carbamoylating nitrosourea: effect on cellular radiosensitivity. Free Radic Biol Med 12: 53-62.

Miller AD, Roxburgh SH, Shea K (2011) How frequency and intensity shape diversity–disturbance relationships. Proc Natl Acad Sci USA 108: 5643–8.

Miller EW, Dickinson BC, Chang CJ (2010) Aquaporin-3 mediates hydrogen peroxide uptake to regulate downstream intracellular signaling. Proc Natl Acad Sci USA 107: 15681-6.

Miller G, Mittler R (2006) Could heat shock transcription factors function as hydrogen peroxide sensors in plants? Ann Bot 98:279-88.

Miller G, Shulaev V, Mittler R (2008) Reactive oxygen signaling and abiotic stress. Physiol Plant 133: 481–9.

Miller JA (2007) Repeated evolution of male sacrifice behavior in spiders correlated with genital mutilation. Evolution 61: 1301–15.

Miller JH (2005) Perspective on mutagenesis and repair: the standard model and alternate modes of mutagenesis. Crit Rev Biochem Mol Biol 40: 155-79.

Miller JH, Suthar A, Tai J, Yeung A, Truong C, Stewart JL (1999) Direct selection for mutators in Escherichia coli. J Bacteriol 181: 1576–84.

Miller MA, Technau U, Smith KM, Steele RE (2000) Oocyte development in Hydra involves selection from competent precursor cells. Dev Biol 224: 326–38.

Miller PS (1994) Is inbreeding depression more severe in a stressful environment? Zoo Biol 13: 195–208.

Miller PS, Hedrick PW (1993) Inbreeding and fitness in captive populations: lessons from Drosophila. Zoo Biol 12: 333–51.

Miller PS, Glasner J, Hedrick PW (1993) Inbreeding depression and male-mating behavior in Drosophila melanogaster. Genetica 88: 29–36.

Miller RA (1938) Spermatogenesis in a sex-reversed female and in normal males of the domestic fowl, Gallus domesticus. Anat Rec 70: 155-83.

Miller WJ, Capy P (2004) Mobile genetic elements as natural tools for genome evolution. Methods Mol Biol 260: 1–20.

Miller WL (2007) Steroidogenic acute regulatory protein (StAR), a novel mitochondrial cholesterol transporter. Biochim Biophys Acta 1771: 663-76.

Mills DR, Peterson RL, Spiegelman S (1967) An extracellular Darwinian experiment with a selfduplicating nucleic acid molecule. Proc Natl Acad Sci USA 58: 217–24.

Mills KD, Ferguson DO, Alt FW (2003) The role of DNA breaks in genomic instability and tumorigenesis. Immunol Rev 194: 77–95.

Mills R, Watson RA (2006) On crossing fitness valleys with the Baldwin effect. In: Rocha LM, Yaeger LS, Bedau MA, Floreano D, Goldstone RL, Vespignani A, eds. Proceedings of the Tenth International Conference on the Simulation and Synthesis of Living Systems (ALifeX). Cambridge, MA: MIT Press. pp 493–499.

Millstein R (2002) Are random drift and natural selection conceptually distinct? Biol Philos 17: 33–53.

Milton CC, Huynh B, Batterham P, Rutherford SL, Hoffmann AA (2003) Quantitative trait symmetry independent of Hsp90 buffering: distinct modes of genetic canalization and developmental stability. Proc Natl Acad Sci USA 100: 13396–401.

Milton CC, Ulane CM, Rutherford S (2006) Control of canalization and evolvability by Hsp90. PLoS ONE 1: e75.

Mimura I, Tanaka T, Wada Y, Kodama T, Nangaku M (2011) Pathophysiological response to hypoxia - from the molecular mechanisms of malady to drug discovery: epigenetic regulation of the hypoxic response via hypoxia-inducible factor and histone modifying enzymes. J Pharmacol Sci 115: 453–8.

Min JY, Lim SO, Jung G (2010) Downregulation of catalase by reactive oxygen species via hypermethylation of CpG island II on the catalase promoter. FEBS Lett 584: 2427-32.

Minchenko O, Opentanova I, Caro J (2003) Hypoxic regulation of the 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase gene family (PFKFB-1–4) expression in vivo. FEBS Lett 554: 264–70.

Minchenko O, Opentanova I, Minchenko D, Ogura T, Esumi H (2004) Hypoxia induces transcription of 6-phosphofructo-2-kinase/fructose-2,6-biphosphatase-4 gene via hypoxia-inducible factor-1 activation. FEBS Lett 576: 14–20.

Minder AM, Hosken DJ, Ward PI (2005) Co-evolution of male and female reproductive characters across the Scathophagidae (Diptera). J Evol Biol 18: 60–9.

Minocherhomji S, Tollefsbol TO, Singh KK (2012) Mitochondrial regulation of epigenetics and its role in human diseases. Epigenetics 7: 326-34.

Mints B (1957) Embryological development of primordial germ cells in the mouse: influence of a new mutation, Wj. J Embryol Exp Morphol 5: 396-403.

Mintz B, Russell ES (1957) Gene-induced embryological modifications of primordial germ cells in the mouse. J Exp Zool 134: 207–30.

Miralles R, Gerrish PJ, Moya A, Elena SF (1999) Clonal interference and the evolution of RNA viruses. Science 285: 1745-7.

Miranda-Vizuete A, Sadek CM, Jimenez A, Krause WJ, Sutovsky P, Oko R (2004) The mammalian testis-specific thioredoxin system. Antioxid Redox Signal 6: 25–40.

Mirault ME, Tremblay A, Beaudoin N, Tremblay M (1991) Overexpression of seleno-glutathione peroxidase by gene transfer enhances the resistance of T47D human breast cells to clastogenic oxidants. J Biol Chem 266: 20752-60.

Miret JJ, Pessoa-Brandao L, Lahue RS (1998) Orientation-dependent and sequence-specific expansions of CTG/CAG trinucleotide repeats in Saccharomyces cerevisiae. Proc Natl Acad Sci USA 95: 12438-44.

Miriti MN, Wright SJ, Howe HF (2001) The effects of neighbors on the demography of a dominant desert shrub (Ambrosia dumosa). Ecol Monogr 71: 491–509.

Misawa K, Kamatani N, Kikuno RF (2008) The universal trend of amino acid gain-loss is caused by CpG hypermutability. J Mol Evol 67: 334–42.

Mishmar D, Ruiz-Pesini E, Golik P, Macaulay V, Clark AG, et al. (2003) Natural selection shaped regional mtDNA variation in humans. Proc Natl Acad Sci USA 100: 171–6.

Miska EA (2005) How microRNAs control cell division, differentiation and death? Curr Opin Genet Dev 15: 563–8.

Mitani K, Fujita H, Sassa S, Kappas A (1990) Activation of heme oxygenase and heat shock protein 70 genes by stress in human hepatoma cells. Biochem Biophys Res Commun 166: 1429-34.

Mitchell BM, Cook LG, Danchuk S, Puschett JB (2007) Uncoupled endothelial nitric oxide synthase and oxidative stress in a rat model of pregnancy-induced hypertension. Am J Hypertens 20: 1297-304.

Mitchell JA, Simmons MJ (1977) Fitness effects of EMS-induced mutations on the X chromosome of Drosophila melanogaster. II. Hemizygous fitness effects. Genetics 87: 775-83.

Mitchell LA, De Iuliis GN, Aitken RJ (2011) The TUNEL assay consistently underestimates DNA damage in human spermatozoa and is influenced by DNA compaction and cell vitality: development of an improved methodology. Int J Androl 34: 2-13.

Mitchell SE, Read AF (2005) Poor maternal environment enhances offspring disease resistance in an invertebrate. Proc Biol Sci 272: 2601-7.

Mitchell SE, Rogers ES, Little TJ, Read AF (2005) Host–parasite and genotype-by-environment interactions: temperature modifies potential for selection by a sterilizing pathogen. Evolution 59: 70–80.

Mitchell-Olds T, Knight CA (2002) Evolution. Chaperones as buffering agents? Science 296: 2348-9.

Mitomo K, Nakayama K, Fujimoto K, Sun X, Seki S, Yamamoto K (1994) Two different cellular redox systems regulate the DNA-binding activity of the p50 subunit of NF-kappa B in vitro. Gene 145: 197–203.

Mitra S, Boldogh I, Izumi T, Hazra TK (2001) Complexities of the DNA base excision repair pathway for repair of oxidative DNA damage. Environ Mol Mutagen 38: 180–90.

Mitra S, Izumi T, Boldogh I, Bhakat KK, Chattopadhyay R, Szczesny B (2007) Intracellular trafficking and regulation of mammalian AP-endonuclease 1 (APE1), an essential DNA repair protein. DNA Repair 6: 461–9.

Mittelbach GG, Schemske DW, Cornell HV, Allen AP, Brown JM, et al. (2007) Evolution and the latitudinal diversity gradient: speciation, extinction and biogeography. Ecol Lett 10: 315-31.

Mittelbrunn M, Sánchez-Madrid F (2012) Intercellular communication: diverse structures for exchange of genetic information. Nat Rev Mol Cell Biol 13: 328-35.

Mitteldorf J (2006) Chaotic population dynamics and the evolution of ageing. Evol Ecol Res 8: 561–74.

Mitteldorf J (2010) Evolutionary origins of aging. In: Fahy GM, West MD, Coles LS, Harris SB, eds. The future of aging. Pathways to human life extension. Dordrecht, NL: Springer. pp 87-126.

Mitteldorf J, Croll DH, Ravela SC (2002) Multilevel selection and the evolution of predatory restraint. Artif Life 8: 146–52.

Mittelman D, ed. (2013) Stress-Induced Mutagenesis. New York, NY: Springer.

Mittelman D, Sykoudis K, Hersh M, Lin Y, Wilson JH (2010) Hsp90 modulates CAG repeat instability in human cells. Cell Stress Chaperones 15: 753-9.

Mittelman D, Wilson JH (2010) Stress, genomes, and evolution. Cell Stress Chaperones 15: 463-6.

Mittler R (2002) Oxidative stress, antioxidants and stress tolerance. Trends Plant Sci 7: 405–10.

Mitton JB, Grant MC (1996) Genetic variation and the natural history of quaking aspen. BioScience 46: 25–31.

Mittwoch U, Mahadevaiah SK (1992) Unpaired chromosomes at meiosis: cause or effect of gametogenic insufficiency? Cytogenet Cell Genet 59: 274–9.

Miura A, Yonebayashi S, Watanabe K, Toyama T, Shimada H, Kakutani T (2001) Mobilization of transposons by a mutation abolishing full DNA methylation in Arabidopsis. Nature 411: 212–4.

Miwa Y, Sasaguri T, Kosaka C, Taba Y, Ishida A, et al. (2000) Differentiation-inducing factor-1, a morphogen of Dictyostelium, induces G1 arrest and differentiation of vascular smooth muscle cells. Circ Res 86: 68–75.

Miyamoto K, Sato EF, Kasahara E, Jikumaru M, Hiramoto K, et al. (2010) Effect of oxidative stress during repeated ovulation on the structure and functions of the ovary, oocytes, and their mitochondria. Free Radic Biol Med 49: 674-81.

Miyamoto MM, Fitch WM (1996) Constraints on protein evolution and the age of the eubacteria/eukaryote split. Syst Biol 45: 568–75.

Miyata T, Hayashida H, Kuma K, Mitsuyasu K, Yasunaga T (1987) Male-driven molecular evolution: a model and nucleotide sequence analysis. Cold Spring Harb Symp Quant Biol 52: 863–7.

Miyazaki T, Sueoka K, Dharmarajan AM, Atlas SJ, Bulkley GB, Wallach EE (1991) Effect of inhibition of oxygen free radical on ovulation and progesterone production by the in-vitro perfused rabbit ovary. J Reprod Fertil 91: 207–12.

Mizuno K, Shiratsuchi A, Masamune Y, Nakanishi Y (1996) The role of Sertoli cells in the differentiation and exclusion of rat testicular germ cells in primary culture. Cell Death Differ 3: 119–23.

Moberg G (1985) Influence of stress on reproduction: Measure of well-being. In: Moberg GP, ed. Animal stress. Bethesda, MD: American Physiological Society. pp 245-267.

Mochizuki K, Nishimiya-Fujisawa C, Fujisawa T (2001) Universal occurrence of the vasa-related genes among metazoans and their germline expression in Hydra. Dev Genes Evol 211: 299–308.

Mochizuki N, Yamamoto M (1992) Reduction of the intracellular cAMP level triggers initiation of sexual development in fission yeast. Mol Gen Genet 233: 17-24.

Mochmann LH, Wells RD (2004) Transcription influences the types of deletion and expansion products in an orientation-dependent manner from GAC*GTC repeats. Nucleic Acids Res 32: 4469-79.

Mock DW, Parker GA (1997) The evolution of sibling rivalry. New York, NY: Oxford University Press.

Mock DW, Parker GA (1998) Siblicide, family conflict and the evolutionary limits of selfishness. Anim Behav 56: 1–10.

Mock N, McConnell NA, Hughes FM Jr, Pandalai SG (2004) Potential role of aquaporins in the formation and expansion of the ovarian follicular antrum and their regulation by gonadotropins. Recent Res Dev Endocrinol 4: 169-79.

Moczek AP, Sultan S, Foster S, Ledón-Rettig C, Dworkin I, et al. (2011) The role of developmental plasticity in evolutionary innovation. Proc Biol Sci 278: 2705-13.

Modi RI, Castilla LH, Puskas-Rozsa S, Helling RB,Adams J (1992) Genetic changes accompanying increased fitness in evolving populations of Escherichia coli. Genetics 130: 241-9.

Moeller BJ, Cao Y, Li CY, Dewhirst MW (2004) Radiation activates HIF-1 to regulate vascular radiosensitivity in tumors: role of reoxygenation, free radicals, and stress granules. Cancer Cell 5: 429–41.

Mogilner JG, Elenberg Y, Lurie M, Shiloni E, Coran AG, Sukhotnik I (2006) Effect of dexamethasone on germ cell apoptosis in the contralateral testis after testicular ischemia-reperfusion injury in the rat. Fertil Steril 85 Suppl 1: 1111-7.

Mogulkoc R, Baltaci AK, Oztekin E, Aydin L, Tuncer I (2005) Hyperthyroidism causes lipid peroxidation in kidney and testis tissues of rats: protective role of melatonin. Neuro Endocrinol Lett 26: 806-10.

Mogulkoc R, Baltaci AK, Aydin L, Oztekin E, Tuncer I (2006) Pinealectomy increases oxidant damage in kidney and testis caused by hyperthyroidism in rats. Cell Biochem Funct 24: 449-53.

Mohanty P, Hamouda W, Garg R, Aljada A, Ghanim H, Dandona P (2000) Glucose challenge stimulates reactive oxygen species (ROS) generation by leucocytes. J Clin Endocrinol Metab 85: 2970-3.

Mohanty S, Firtel RA (1999) Control of spatial patterning and cell-type proportioning in Dictyostelium. Semin Cell Dev Biol 10: 597-607.

Mohn F, Schübeler D (2009)Genetics and epigenetics: stability and plasticity during cellular differentiation. Trends Genet 25: 129-36.

Mokany K, Raison RJ, Prokushkin AS (2006) Critical analysis of root: shoot ratios in terrestrial biomes. Glob Change Biol 12: 84–96.

Mokkapati SK, Bhagwat AS (2002) Lack of dependance of transcription-induced cytosine deaminations on protein synthesis. Mutat Res 508: 131-6.

Molinier J, Ries G, Zipfel C, Hohn B (2006) Transgeneration memory of stress in plants. Nature 442: 1046-9.

Møller AP (1988a) Testes size, ejaculate quality and sperm competition in birds. Biol J Linn Soc 33: 273–83.

Møller AP (1988b) Ejaculate quality, testes size and sperm competition in primates. J Human Evol 17: 479–88.

Møller AP (1989) Ejaculate quality, testes size and sperm production in mammals. Funct Ecol 3: 91–6.

Møller AP (1991) Sperm competition, sperm depletion, paternal care, and relative testis size in birds. Am Nat 137: 882–906.

Møller AP (1994) Sexual selection and the barn swallow. New York, NY: Oxford University Press.

Møller AP, Saino N (1994) Parasites, immunology of hosts, and host sexual selection. J Parasitol 80: 850-8

Møller AP, Briskie JV (1995) Extra-pair paternity, sperm competition and the evolution of testis size in birds. Behav Ecol Sociobiol 36: 357-65.

Møller AP, Swaddle JP (1996) Developmental stability and evolution. Oxford, UK: Oxford University Press.

Møller AP, Thornhill R (1998) Bilateral symmetry and sexual selection: a meta-analysis. Am Nat 151: 174–92.

Møller AP, Pagel M (1998) Developmental stability and signalling among cells. J Theor Biol 193: 497-506.

Møller AP, Dufva R, Erritzøe J (1998) Host immune function and sexual selection in birds. J Evol Biol 11: 703-19.

Møller AP, Alatalo RV (1999) Good-genes effects in sexual selection. Proc R Soc Lond B 266: 85-91.

Møller AP, Christe P, Lux E (1999) Parasitism, host immune function, and sexual selection. Q Rev Biol 74: 3-20.

Møller P, Loft S (2004) Interventions with antioxidants and nutrients in relation to oxidative DNA damage and repair. Mutat Res 551: 79-89.

Molnar A, Melnyk CW, Bassett A, Hardcastle TJ, Dunn R, Baulcombe DC (2010) Small silencing RNAs in plants are mobile and direct epigenetic modification in recipient cells. Science 328: 872–5.

Moloney DM, Slaney SF, Oldridge M, Wall SA, Sahlin P, et al. (1996) Exclusive paternal origin of new mutations in Apert syndrome. Nat Genet 13: 48-53.

Monaghan P, Metcalfe NB, Torres R (2009) Oxidative stress as a mediator of life history trade-offs: mechanisms, measurements and interpretation. Ecol Lett 12: 75-92.

Moncada S, Erusalimsky JD (2002) Does nitric oxide modulate mitochondrial energy generation and apoptosis? Nat Rev Mol Cell Biol 3: 214–20.

Monder C, Hardy MP, Blanchard RJ, Blanchard DC (1994a) Comparative aspects of 11β-hydroxysteroid dehydrogenase. Testicular 11β-hydroxysteroid dehydrogenase: development of a model for the mediation of Leydig cell function by corticosteroids. Steroids 59: 69–73.

Monder C, Miroff Y, Marandici A, Hardy M (1994b) 11β-Hydroxysteroid dehydrogenase alleviates glucocorticoid-mediated inhibition of steroidogenesis in rat Leydig cells. Endocrinology 134: 1199–204.

Monder C, Sakai RR, Miroff Y, Blanchard DC, Blanchard RJ (1994c) Reciprocal changes in plasma corticosterone and testosterone in stressed male rats maintained in a visible burrow system: evidence for a mediating role of testicular 11β-hydroxysteroid dehydrogenase. Endocrinology 134: 1193–8.

Mongold JA (1992) DNA repair and the evolution of transformation in Haemophilus influenzae. Genetics 132: 893-8.

Monk M, Boubelik M, Lehnert S (1987) Temporal and regional changes in DNA methylation in the embryonic, extraembryonic and germ cell lineages during mouse embryo development. Development 99: 371–82.

Monnat RJ, Loeb LA (1985) Nucleotide sequence preservation of human mitochondrial DNA. Proc Natl Acad Sci USA 82: 2895–9.

Monod J (1971) Chance and Necessity. New York: NY: Vintage Books.

Monro K, Poore AGB (2004) Selection in modular organisms: is intraclonal variation in macroalgae evolutionarily important? Am Nat 163: 564-78.

Monro K, Poore AGB (2009) The potential for evolutionary responses to cell-lineage selection on growth form and its plasticity in a red seaweed. Am Nat 173: 151–63.

Monro K, Sinclair-Taylor T, Marshall DJ (2010) Selection on offspring size among environments: the roles of environmental quality and variability. Funct Ecol 24: 676–84.

Montandon AJ, Green PM, Bentley DR, Ljung R, Kling S, et al. (1992) Direct estimate of the haemophilia B (factor IX deficiency) mutation rate and of the ratio of the sex-specific mutation rates in Sweden. Hum Genet 89: 319-322.

Monte M, Simonatto M, Peche LY, Bublik DR, Gobessi S, et al. (2006) MAGE-A tumor antigens target p53 transactivation function through histone deacetylase recruitment and confer resistance to chemotherapeutic agents. Proc Natl Acad Sci USA 103: 11160-5.

Monteleone P, Giovanni Artini P, Simi G, Casarosa E, Cela V, Genazzani AR (2008) Follicular fluid VEGF levels directly correlate with perifollicular blood flow in normoresponder patients undergoing IVF. J Assist Reprod Genet 25: 183-6.

Montero L, Müller N, Gallant P (2008) Induction of apoptosis by Drosophila Myc. Genesis 46: 104-11.

Montgomery TA, Howell MD, Cuperus JT, Li D, Hansen JE, Alexander AL, et al. (2008) Specificity of ARGONAUTE7-mir390 interaction and dual functionality in TAS3 trans-acting siRNA formation. Cell 133: 128–41.

Monti V, Mandrioli M, Rivi M, Manicardi GC (2012) The vanishing clone: karyotypic evidence for extensive intraclonal genetic variation in the peach potato aphid, Myzus persicae (Hemiptera: Aphididae). Biol J Linn Soc 105: 350–8.

Montrose VT, Harris WE, Moore AJ, Moore PJ (2008) Sperm competition within a dominance hierarchy: investment in social status vs. investment in ejaculates. J Evol Biol 21: 1290-6.

Montville R, Froissart R, Remold SK, Tenaillon O, Turner PE (2005) Evolution of mutational robustness in an RNA virus. PLoS Biol 3: e381.

Mooney HA, Cleland EE (2001) The evolutionary impact of invasive species. Proc Natl Acad Sci USA 98: 5446–51.

Moor RM, Seamark RF (1986) Cell signaling, permeability, and microvasculatory changes during antral follicle development in mammals. J Dairy Sci 69: 927-43.

Moore CR (1926) The biology of the mammalian testis and scrotum. Q Rev Biol 1: 4–50.

Moore CR, Simmons GF, Wells LJ, Zalesky M, Nelson WO (1934) On the control of reproductive activity in an annual-breeding mammal (Citellus tridecemlineatus). Anat Rec 60: 279-89.

Moore DH, Michael H, Tritt R, Parsons SH, Kelley MR (2000) Alterations in the expression of the DNA repair/redox enzyme APE/ref-1 in epithelial ovarian cancers. Clin Cancer Res 6: 602–9.

Moore FBG, Rozen DE, Lenski RE (2000) Pervasive compensatory adaptation in Escherichia coli. Proc R Soc Lond B 267: 515–22.

Moore FL, Zoeller RT (1985) Stress-induced inhibition of sexual reproduction: Evidence of suppressed secretion of LH-RH in an amphibian. Gen Comp Endocrinol 60: 252-8.

Moore FL, Reijo-Pera RA (2000) Male sperm motility dictated by mother’s mtDNA. Am J Hum Genet 67: 543–8.

Moore H, Dvoráková K, Jenkins N, Breed W (2002) Exceptional sperm cooperation in the wood mouse. Nature 418: 174-7.

Moore HDM, Akhondi MA (1996) Fertilizing capacity of rat spermatozoa is correlated with decline in straightline velocity measured by continuous computer-aided sperm analysis: epididymal rat spermatozoa form the proximal cauda have a greater fertilizing capacity in vitro than those from the distal cauda or vas deferens. J Androl 17: 50–60.

Moore IT, Jessop TS (2003) Stress, reproduction, and adrenocortical modulation in amphibians and reptiles. Horm Behav 43: 39-47.

Moore MC, Thompson CW, Marler CA (1991) Reciprocal changes in corticosterone and testosterone levels following acute and chronic handling stress in the tree lizard, Urosaurus ornatus. Gen Comp Endocrinol 81: 217-26.

Moore MJ (2005) From birth to death: the complex lives of eukaryotic mRNAs. Science 309: 1514-8.

Moore TDE, Edman JC (1993) The alpha-mating type locus of Cryptococcus neoformans contains a peptide pheromone gene. Mol Cell Biol 13: 1962-70.

Moore WS (1976) Components of fitness in the unisexual fish Poeciliopsis monacha-occidentalis. Evolution 30: 564-78.

Moore WS (1984) Evolutionary ecology of unisexual fishes. In: Turner BJ, ed. Evolutionary genetics of fishes. New York, NY: Plenum Press. pp 329-398.

Moore WS, Mckay FE (1971) Coexistence in unisexual-bisexual species complexes of Poeciliopsis (Pisces: Poeciliidae). Ecology 52: 791-9.

Moorhead PS, Kaplan MM, eds. (1967) Mathematical challenges to the neo-Darwinian interpretation of evolution. Philadelphia, PA: Wistar Institute Press.

Mootha VK, Bunkenborg J, Olsen JV, Hjerrild M, Wisniewski JR, et al. (2003) Integrated analysis of protein composition, tissue diversity, and gene regulation in mouse mitochondria. Cell 115: 629–40.

Moran NA (1992) The evolutionary maintenance of alternative phenotypes. Am Nat 139: 971–89.

Moran NA (1996) Accelerated evolution and Muller’s ratchet in endosymbiotic bacteria. Proc Natl Acad Sci USA 93: 2873–8.

Morange M (2008) What history tells us XIII. Fifty years of the Central Dogma. J Biosci 33: 171-5.

Morata G, Ripoll P (1975) Minutes: mutants of Drosophila autonomously affecting cell division rate. Dev Biol 42: 211–21.

Morata G, Martin FA (2007) Cell competition: the embrace of death. Dev Cell 13: 1–2.

Moreira PI, Santos MS, Sena C, Seiça R, Oliveira CR (2005) Insulin protects against amyloid β-peptide toxicity in brain mitochondria of diabetic rats. Neurobiol Dis 18: 628–37.

Moreira PI, Rolo AP, Sena C, Seiça R, Oliveira CR, Santos MS (2006) Insulin attenuates diabetes-related mitochondrial alterations: a comparative study. Med Chem 2: 299–308.

Moreira da Silva F, Marques A, Chaveiro A (2010) Reactive oxygen species: a double-edged sword in reproduction. Open Vet Sci J 4: 127-33.

Moreno E (2008) Is cell competition relevant to cancer? Nat Rev Cancer 8: 141-7.

Moreno E, Basler K, Morata G (2002) Cells compete for decapentaplegic survival factor to prevent apoptosis in Drosophila wing development. Nature 416: 755–9.

Moreno E, Basler K (2004) dMyc transforms cells into super-competitors. Cell 117: 117–29.

Moreno RD, Lizama C, Urzúa N, Vergara SP, Reyes JG (2006) Caspase activation throughout the first wave of spermatogenesis in the rat. Cell Tissue Res 325: 533-40.

Moreno-Loshuertos R, Acin-Perez R, Fernandez-Silva P, Movilla N, Perez-Martos A, et al. (2006) Differences in reactive oxygen species production explain the phenotypes associated with common mouse mitochondrial DNA variants. Nat Genet 38: 1261–8.

Morgan EA (2005) Environmental stress and its effects on mutation rates in Drosophila melanogaster. Thesis. Bowling Green State University, Ohio.

Morgan MJ, Liu ZG (2011) Crosstalk of reactive oxygen species and NF-κB signaling. Cell Res 21: 103-15.

Morgan HD, Sutherland HG, Martin DI, Whitelaw E (1999) Epigenetic inheritance at the agouti locus in the mouse. Nat Genet 23: 314–8.

Morgan HD, Santos F, Green K, Dean W, Reik W (2005) Epigenetic reprogramming in mammals. Hum Mol Genet 14: R47–R58.

Morgan JW (1995) Ecological studies of the endangered Rutidosis leptorrhynchoides. I. Seed production, soil seed bank dynamics, population density and their effects on recruitment. Aust J Bot 43: 1–11.

Morgan MJ, Liu ZG (2010) Reactive oxygen species in TNFalpha-induced signaling and cell death. Mol Cells 30: 1-12.

Morgan TH (1903) Evolution and adaptation. New York, NY: MacMillan.

Morgante M, Brunner S, Pea G, Fengler K, Zuccolo A, Rafalski A (2005) Gene duplication and exon shuffling by helitron-like transposons generate intraspecies diversity in maize. Nat Genet 37: 997–1002.

Morillon A, Springer M, Lesage P (2000) Activation of the Kss1 invasive-filamentous growth pathway induces Ty1 transcription and retrotransposition in Saccharomyces cerevisiae. Mol Cell Biol 20: 5766–76.

Morimoto RI (1993) Cells in stress: transcriptional activation of heat shock genes. Science 259: 1409-10.

Morimoto RI (1998) Regulation of the heat shock transcriptional response: cross talk between a family of heat shock factors, molecular chaperones, and negative regulators. Genes Dev 12: 3788–96.

Morimoto RI (2008) Proteotoxic stress and inducible chaperone networks in neurodegenerative disease and aging. Genes Dev 22: 1427-38.

Morita M, Best JB (1984) Effects of photoperiods and melatonin on planarian asexual reproduction. J Exp Zool 231: 273–82.

Morita M, Hall F, Best JB (1987) Photoperiodic modulation of cephalic melatonin in planarians. J Exp Zool 241: 383–8.

Morita RY (1993) Bioavailability of energy and the starvation state. In: Kjelleberg S, ed. Starvation in Bacteria. New York, NY: Plenum. pp 1–23.

Morita Y, Tilly JL (1999) Oocyte apoptosis: like sand through an hourglass. Dev Biol 213: 1–17.

Morita Y, Manganaro TF, Tao XJ, Martimbeau S, Donahoe PK, Tilly JL (1999) Requirement for phosphatidylinositol-3’- kinase in cytokine-mediated germ cell survival during fetal oogenesis in the mouse. Endocrinology 140: 941–9.

Morita Y, Perez GI, Paris F, Miranda SR, Ehleiter D, et al. (2000) Oocyte apoptosis is suppressed by disruption of the acid sphingomyelinase gene or by sphingosine-1-phosphate therapy. Nat Med 6: 1109–14.

Morita Y, Maravei DV, Bergeron L, Wang S, Perez GI, et al. (2001) Caspase-2 deficiency rescues female germ cells from death due to cytokine insufficiency but not meiotic defects caused by ataxia telangiectasia-mutated (Atm) gene inactivation. Cell Death Differ 8: 614–20.

Moritz C, Donnelan S, Adams M, Baverstock PR (1989) The origin and evolution of parthenogenesis in Heteronotia binoei (Gekkonidae): extensive genotypic diversity among parthenogens. Evolution 43: 994–1003.

Moritz C, McCallum H, Donnellan S, Roberts JD (1991) Parasite loads in parthenogenetic and sexual lizards (Heteronotia binoei ): support for the Red Queen hypothesis. Proc R Soc Lond Ser B 244: 145–9.

Morran LT, Parmenter MD, Phillips PC (2009a) Mutation load and rapid adaptation favor outcrossing over self-fertilization. Nature 462: 350–2.

Morran LT, Cappy BJ, Anderson JL, Phillips PC (2009b) Sexual partners for the stressed: facultative outcrossing in the self-fertilizing nematode Caenorhabditis elegans. Evolution 63: 1473–82.

Morran LT, Ohdera AH, Phillips PC (2010) Purging deleterious mutations under self fertilization: paradoxical recovery in fitness with increasing mutation rate in Caenorhabditis elegans. PLoS ONE 5: e14473.

Morran LT, Schmidt OG, Gelarden IA, Parrish RC 2nd, Lively CM (2011) Running with the Red Queen: host-parasite coevolution selects for biparental sex. Science 333: 216-8.

Morris ID, Ilott S, Dixon L, Brison DR (2002) The spectrum of DNA damage in human sperm assessed by single cell gel electrophoresis (Comet assay) and its relationship to fertilization and embryo development. Hum Reprod 17: 990-8.

Morris JK, Mutton DE, Alberman E (2002) Revised estimates of the maternal age specific live birth prevalence of Down’s syndrome. J Med Screen 9: 2–6.

Morris JK, de Vigan C, Mutton DE, Alberman E (2005) Risk of a Down syndrome live birth in women 45 years of age and older. Prenat Diagn 25: 275–8.

Morris KV (2008) RNA-mediated transcriptional gene silencing in human cells. Curr Top Microbiol Immunol 320: 211-24.

Morris KV, Chan SWL, Jacobsen SE, Looney DJ (2004) Small interfering RNA-induced transcriptional gene silencing in human cells. Science 305: 1289–92.

Morrison DA (1997) Streptococcal competence for genetic transformation: regulation by peptide pheromones. Microb Drug Resist 3: 27-37.

Morrison DA, Baker MF (1979) Competence for genetic transformation in Pneumococcus depends on synthesis of a small set of proteins. Nature 282: 215-7.

Morrongiello JR, Bond NR, Crook DA, Wong BB (2012) Spatial variation in egg size and egg number reflects trade-offs and bet-hedging in a freshwater fish. J Anim Ecol 81: 806-17.

Morrow EH, Gage MJG (2000) The evolution of sperm length in moths. Proc R Soc Lond B 267: 307–13.

Morselli E, Maiuri MC, Markaki M, Megalou E, Pasparaki A, et al. (2010) Caloric restriction and resveratrol promote longevity through the Sirtuin-1-dependent induction of autophagy. Cell Death Dis 1: e10.

Mortusewicz O, Schermelleh L, Walter J, Cardoso MC, Leonhardt H (2005) Recruitment of DNA methyltransferase I to DNA repair sites. Proc Natl Acad Sci USA 102: 8905–9.

Moser CM, Page CC, Dutton PL (2006) Darwin at the molecular scale: selection and variance in electron tunnelling proteins including cytochrome c oxidase. Phil Trans R Soc B 361: 1295-305.

Moskwa P, Buffa FM, Pan Y, Panchakshari R, Gottipati P, et al. (2011) miR-182-mediated downregulation of BRCA1 impacts DNA repair and sensitivity to PARP inhibitors. Mol Cell 41: 210-20.

Mott JL, Kobayashi S, Bronk SF, Gores GJ (2007) mir-29 regulates Mcl-1 protein expression and apoptosis. Oncogene 26: 6133–40.

Mouse Genome Sequencing Consortium (2002) Initial sequencing and comparative analysis of the mouse genome. Nature 420: 520-62.

Mousseau TA, Roff DA (1987) Natural selection and the heritability of fitness components. Heredity 59: 181–98.

Moxon ER, Rainey PB, Nowak MA, Lenski RE (1994) Adaptive evolution of highly mutable loci in pathogenic bacteria. Curr Biol 4: 24-33.

Moya A, Elena SF, Bracho A, Miralles R, Barrio E (2000) The evolution of RNA viruses: A population genetics view. Proc Natl Acad Sci USA 97: 6967-73.

Moynahan ME, Jasin M (2010) Mitotic homologous recombination maintains genomic stability and suppresses tumorigenesis. Nat Rev Mol Cell Biol 11: 196–207.

Mruk DD, Cheng CY (2004) Sertoli-Sertoli and Sertoli-germ cell interactions and their significance in germ cell movement in the seminiferous epithelium during spermatogenesis. Endocr Rev 25: 747-806.

Msadek T (1999) When the going gets tough: survival strategies and environmental signaling networks in Bacillus subtilis. Trends Microbiol 7: 201-7.

Mu J, Awadalla P, Duan J, McGee KM, Keebler J, et al. (2006) Genome-wide variation and identification of vaccine targets in the Plasmodium falciparum genome. Nat Genet 39: 126–30.

Mugal CF, von Grünberg H-H, Peifer M (2009) Transcription-induced mutational strand bias and its effect on substitution rates in human genes. Mol Biol Evol 26: 131–42.

Mujahid A, Yoshiki Y, Akiba Y, Toyomizu M (2005) Superoxide radical production in chicken skeletal muscle induced by acute heat stress. Poult Sci 84: 307-14.

Mujahid A, Akiba Y, Warden CH, Toyomizu M (2007) Sequential changes in superoxide production, anion carriers and substrate oxidation in skeletal muscle mitochondria of heat-stressed chickens. FEBS Lett 581: 3461-7.

Mukai T (1961) A radiation-genetical consideration concerning the structure of natural populations. Jpn J Genet 36 (Suppl.): 155-66.

Mukai T (1969) The genetic structure of natural populations of Drosophila melanogaster. VII. Synergistic interaction of spontaneous mutant polygenes controlling viability. Genetics 61: 749–61.

Mukai T (1970) Viability mutations induced by ethyl methanesulfonate in Drosophila melanogaster. Genetics 65: 335-48.

Mukai T, Chigusa SI, Mettler LE, Crow JF (1972) Mutation rate and dominance of genes affecting viability in Drosophila melanogaster. Genetics 72: 335–55.

Mularoni L, Zhou Y, Bowen T, Gangadharan S, Wheelan SJ, Boeke JD (2012) Retrotransposon Ty1 integration targets specifically positioned asymmetric nucleosomal DNA segments in tRNA hotspots. Genome Res 22: 693-703.

Mulcahy DL (1974) Correlation between speed of pollen tube growth and seedling height in Zea mays L. Nature 249: 491-3.

Mulcahy DL (1979) The rise of the angiosperms: a genecological factor. Science 206: 20–3.

Mulcahy DL, Bergamini-Mulcahy G (1987) The effects of pollen competition. Am Sci 75: 44-50.

Mulcahy DL, Sari-Gorla M, Bergamini-Mulcahy G (1996) Pollen selection: past, present and future. Sex Plant Reprod 9: 353–6.

Müller F, Tobler H (2000) Chromatin diminution in the parasitic nematodes Ascaris suum and Parascaris univalens. Int J Parasitol 30: 391–9.

Müller GB (1990) Developmental mechanisms at the origin of morphological novelty: a side-effect hypothesis. In: Nitecki MH, ed. Evolutionary innovations. Chicago, IL: University of Chicago Press. pp 99-132.

Müller GB (2007) Evo-devo: extending the evolutionary synthesis. Nat Rev Genet 8: 943–9.

Muller HJ (1928) The measurement of gene mutation in Drosophila, its high variability and its dependence upon temperature. Genetics 13: 279–357.

Muller HJ (1932) Some genetic aspects of sex. Am Nat 66: 118-38.

Muller HJ (1964) The relation of recombination to mutational advance. Mutat Res 1: 2-9.

Müller LU, Milsom MD, Harris CE, Vyas R, Brumme KM, et al. (2012) Overcoming reprogramming resistance of Fanconi anemia cells. Blood 119: 5449–57.

Müller M, Strand S, Hug H, Heinemann EM, Walczak H, et al. (1997) Drug-induced apoptosis in hepatoma cells is mediated by the CD95 (APO-1/Fas) receptor/ligand system and involves activation of wild-type p53. J Clin Invest 99: 403-13.

Müller WE (2006) The stem cell concept in sponges (Porifera): metazoan traits. Semin Cell Dev Biol 17: 481–91.

Multigner L, Spira A (1997) The epidemiology of male reproduction. In: Barrati C, de Jong C, Mortimer D, J Parinaud J, eds. Genetics of Human Male Fertility. Paris, France: EDK Press. pp 43-65.

Mumm S, Whyte MP, Thakker RV, Buetow KH, Schlessinger D (1997) mtDNA analysis shows common ancestry in two kindreds with X-linked recessive hypoparathyroidism and reveals a heteroplasmic silent mutation. Am J Hum Genet 60: 153–9.

Munday P, Caley J, Jones G (1998) Bi-directional sex change in a coral-dwelling goby. Behav Ecol Sociobiol 43: 371–7.

Munier FL, Thonney F, Balmer A, Héon E, Pescia G, Schorderet DF (1996) Sex mutation ratio in retinoblastoma and retinoma: relevance to genetic counseling. Klin Monbl Augenheilkd 208: 400-3.

Muñoz-Fernández MA, Fresno M (1998) The role of tumour necrosis factor, interleukin 6, interferon-gamma and inducible nitric oxide synthase in the development and pathology of the nervous system. Prog Neurobiol 56: 307-40.

Muotri AR, Chu VT, Marchetto MC, Deng W, Moran JV, Gage FH (2005) Somatic mosaicism in neuronal precursor cells mediated by L1 retrotransposition. Nature 435: 903-10.

Muotri AR, Gage FH (2006) Generation of neuronal variability and complexity. Nature 441: 1087-93.

Muotri AR, Marchetto MC, Coufal NG, Gage FH (2007) The necessary junk: new functions for transposable elements. Hum Mol Genet 16: R159-67.

Muotri AR, Zhao C, Marchetto MC, Gage FH (2009) Environmental influence on L1 retrotransposons in the adult hippocampus. Hippocampus 19: 1002-7.

Muotri AR, Marchetto MC, Coufal NG, Oefner R, Yeo G, et al. (2010) L1 retrotransposition in neurons is modulated by MeCP2. Nature 468: 443-6.

Muramoto T, Suzuki K, Shimizu H, Kohara Y, Koriki E, et al. (2003) Construction of a gamete-enriched gene pool and RNAi-mediated functional analysis in Dictyostelium discoideum. Mech Dev 120: 967-75.

Muratori M, Marchiani S, Maggi M, Forti G, Baldi E (2006) Origin and biological significance of DNA fragmentation in human spermatozoa. Front Biosci 11: 1491–9.

Murchison EP, Stein P, Xuan Z, Pan H, Zhang MQ, Schultz RM, Hannon GJ (2007) Critical roles for Dicer in the female germline. Genes Dev 21: 682–93.

Murdoch WW (1994) Population regulation in theory and practice. Ecology 75: 271–87.

Murie JO, Dobson FS (1987) The costs of reproduction in female Columbian ground squirrels (Spermophilus columbianus). Oecologia 73: 1-6.

Murphey P, McLean DJ, McMahan CA, Walter CA, McCarrey JR (2013) Enhanced genetic integrity in mouse germ cells. Biol Reprod 88: 6.

Murphy BD (2000) Models of luteinization. Biol Reprod 63: 2–11.

Murphy EC, Haukioja E (1986) Clutch size in nidicolous birds. In: Johnston RF, ed. Current Ornithology, Vol. 4. New York, NY: Plenum. pp 141-180.

Murphy GI (1968) Pattern in life history and the environment. Am Nat 102: 390-404.

Murphy MP (2009) How mitochondria produce reactive oxygen species. Biochem J 417: 1–13.

Murphy MP (2012) Modulating mitochondrial intracellular location as a redox signal. Sci Signal 5: pe39.

Murphy-Ullrich JE, Lane TF, Pallero MA, Sage EH (1995) SPARC mediates focal adhesion disassembly in endothelial cells through a follistatin-like region and the Ca2+-binding EF-hand. J Cell Biochem 57: 341–50.

Murray Jr BG (1979) Population Dynamics. New York, NY: Academic Press.

Murray DL, Keith LB, Cary JR (1998) Do parasitism and nutritional status interact to affect production in showshoe hares? Ecology 79: 1209–22.

Murray-Zmijewski F, Lane DP, Bourdon JC (2006) p53/p63/p73 isoforms: an orchestra of isoforms to harmonise cell differentiation and response to stress. Cell Death Differ 13: 962-72.

Murray-Zmijewski F, Slee EA, Lu X (2008) A complex barcode underlies the heterogeneous response of p53 to stress. Nat Rev Mol Cell Biol 9: 702-12.

Murrow L, Debnath J (2013) Autophagy as a stress-response and quality-control mechanism: implications for cell injury and human disease. Annu Rev Pathol Mech Dis 8: 105–37.

Murton RK, Lofts B, Westwood NJ (1970a) The circadian basis of photoperiodically controlled spermatogenesis in the greenfinch Chloris chloris. J Zool 161: 125-36.

Murton RK, Lofts B, Orr AH (1970b) The significance of circadian based photosensitivity in the house sparrow Passer domesticus. Ibis 112: 448-56.

Murugesan P, Kanagaraj P, Yuvaraj S, Balasubramanian K, Aruldhas MM, Arunakaran J (2005) The inhibitory effects of polychlorinated biphenyl Aroclor 1254 on Leydig cell LH receptors, steroidogenic enzymes and antioxidant enzymes in adult rats. Reprod Toxicol 20: 117–26.

Musselman CA, Lalonde ME, Côté J, Kutateladze TG (2012) Perceiving the epigenetic landscape through histone readers. Nat Struct Mol Biol 19: 1218-27.

Mustonen V, Lässig M (2009) From fitness landscapes to seascapes: Non-equilibrium dynamics of selection and adaptation. Trends Genet 25: 111–9.

Mustonen V, Lässig M (2010) Fitness flux and ubiquity of adaptive evolution. Proc Natl Acad Sci USA 107: 4248-53.

Mutharasan RK, Nagpal V, Ichikawa Y, Ardehali H (2011) microRNA-210 is upregulated in hypoxic cardiomyocytes through Akt- and p53-dependent pathways and exerts cytoprotective effects. Am J Physiol Heart Circ Physiol 301: H1519–1530.

Mydlarski MB, Liberman A, Schipper HM (1995) Estrogen induction of glial heat shock proteins: implications for hypothalamic aging. Neurobiol Aging 16: 977-81.

Myers RA, Barrowman NJ, Hutchings JA, Rosenberg AA (1995) Population dynamics of exploited fish stocks at low population levels. Science 269: 1106–8.

Myers RB, Abney TO (1988) The effects of reduced O2 and antioxidants on steroidogenic capacity of cultured rat Leydig cells. J Steroid Biochem 31: 305-9.

Myers S, Bottolo L, Freeman C, McVean G, Donnelly P (2005) A fine-scale map of recombination rates and hotspots across the human genome. Science 310: 321–4.

Myers S, Freeman C, Auton A, Donnelly P, McVean G (2008) A common sequence motif associated with recombination hot spots and genome instability in humans. Nat Genet 40: 1124-9.

Myers S, Bowden R, Tumian A, Bontrop RE, Freeman C, et al. (2010) Drive against hotspot motifs in primates implicates the PRDM9 gene in meiotic recombination. Science 327: 876-9.

Myohara M, Yoshida-Noro C, Kobari F, Tochinai S (1999) Fragmenting oligochaete Enchytraeus japonensis: a new material for regeneration study. Dev Growth Differ 41: 549–55.

Myung K, Datta A, Kolodner RD (2001) Suppression of spontaneous chromosomal rearrangements by S phase checkpoint functions in Saccharomyces cerevisiae. Cell 104: 397–408.

Nabholz B, Glémin S, Galtier N (2008a) Strong variations of mitochondrial mutation rate across mammals: the longevity hypothesis. Mol Biol Evol 25: 120–30.

Nabholz B, Mauffrey JF, Bazin E, Galtier N, Glémin S (2008b) Determination of mitochondrial genetic diversity in mammals. Genetics 178: 351-61.

Nabholz B, Glémin S, Galtier N (2009) The erratic mitochondrial clock: variations of mutation rate, not population size, affect mtDNA diversity across mammals and birds. BMC Evol Biol 9: 54.

Nachman MW (1997) Patterns of DNA variability at X-linked loci in Mus domesticus. Genetics 147: 1303–16.

Nachman MW (1998) Deleterious mutations in animal mitochondrial DNA. Genetica 102/103: 61-9.

Nachman MW, Boyer SN, Aquadro CF (1994) Non-neutral evolution at the mitochondrial ND3 gene in mice. Proc Natl Acad Sci USA 91: 6364-8.

Nachman MW, Brown WM, Stoneking M, Aquadro CF (1996) Nonneutral mitochondrial DNA variation in humans and chimpanzees. Genetics 142: 953-63.

Nachman MW, Bauer VL, Crowell SL, Aquadro CF (1998) DNA variability and recombination rates at X-linked loci in humans. Genetics 150: 1133–41.

Nachman MW, Crowell SL (2000) Estimate of the mutation rate per nucleotide in humans. Genetics 156: 297–304.

Nadeau JH (2009) Transgenerational genetic effects on phenotypic variation and disease risk. Hum Mol Genet 18: R202–10.

Nagae G, Isagawa T, Shiraki N, Fujita T, Yamamoto S, et al. (2011) Tissue-specific demethylation in CpG-poor promoters during cellular differentiation. Hum Mol Genet 20: 2710–21.

Nagai S, Mabuchi T, Hirata S, Shoda T, Kasai T, Yokota S, Shitara H, Yonekawa H, Hoshi K (2004) Oocyte mitochondria: strategies to improve embryogenesis. Hum Cell 17: 195–201.

Nagano T, Mitchell JA, Sanz LA, Pauler FM, Ferguson-Smith AC, et al. (2008) The Air noncoding RNA epigenetically silences transcription by targeting G9a to chromatin. Science 322: 1717-20.

Nagao M, Ebert BL, Ratcliffe PJ, Pugh CW(1996) Drosophila melanogaster SL2 cells contain a hypoxically inducible DNA binding complex which recognises mammalian HIF-1 binding sites. FEBS Lett 387: 161–6.

Nagaraja AK, Andreu-Vieyra C, Franco HL, Ma L, Chen RH, et al. (2008) Deletion of Dicer in somatic cells of the female reproductive tract causes sterility. Mol Endocrinol 22: 2336–52.

Nagata S (1997) Apoptosis by death factor. Cell 88: 355–65.

Nagata S, Golstein P (1995) The Fas death factor. Science 267: 1449–56.

Naito K, Zhang F, Tsukiyama T, Saito H, Hancock CN, et al. (2009) Unexpected consequences of a sudden and massive transposon amplification on rice gene expression. Nature 461: 1130–4.

Nakae D, Akai H, Kishida H, Kusuoka O, Tsutsumi M, Konishi Y (2000) Age and organ dependent spontaneous generation of nuclear 8-hydroxydeoxyguanosine in male Fischer 344 rats. Lab Invest 80: 249-61.

Nakai A, Suzuki M, Tanabe M (2000) Arrest of spermatogenesis in mice expressing an active heat shock transcription factor 1. EMBO J 19: 1545–54.

Nakai Y, Kubota S, Kohno S (1991) Chromatin diminution and chromosome elimination in four Japanese hagfish species. Cytogenet Cell Genet 56: 196-8.

Nakai Y, Kubota S, Goto Y, Ishibashi T, Davison W, Kohno S (1995) Chromosome elimination in three Baltic, south Pacific and north-east Pacific hagfish species. Chromosome Res 3: 321-30.

Nakajima Y, DelliPizzi AM, Mallouh C, Ferreri NR (1996) TNF-mediated cytotoxicity and resistance in human prostate cancer cell lines. Prostate 29: 296–302.

Nakamura BN, Lawson G, Chan JY, Banuelos J, Cortes MM, et al. (2010) Knockout of the transcription factor NRF2 disrupts spermatogenesis in an age-dependent manner. Free Radic Biol Med 49:1368–79.

Nakamura H, Nakamura K, Yodoi J (1997) Redox regulation of cellular activation. Annu Rev Immunol 15: 351–69.

Nakamura M, Fujiwara A, Yasumasu I, Okinaga S, Arai K (1982) Regulation of glucose metabolism by adenine nucleotides in round spermatids from rat testes. J Biol Chem 257: 13945–50.

Nakamura M, Okinaga S, Arai K (1984a) Metabolism of round spermatids: evidence that lactate is preferred substrate. Am J Physiol 247: E234–E242.

Nakamura M, Okinaga S, Arai K (1984b) Metabolism of pachytene primary spermatocytes from rat testes: pyruvate maintenance of adenosine triphosphate level. Biol Reprod 30: 1187–97.

Nakanishi Y, Shiratsuchi A (2004) Phagocytic removal of apoptotic spermatogenic cells by Sertoli cells: mechanisms and consequences. Biol Pharm Bull 27: 13-6.

Nakatsuru K, Kramer DL (1982) Is sperm cheap? Limited fertility and female choice in the lemon tetra (Pisces, Characidae). Science 216: 753-5.

Nakayama H, Ru XM, Fujita J, Kasugai T, Onoue H, et al. (1990) Growth competition between W mutant and wild type cells in mouse aggregation chimeras. Dev Growth Differ 32: 255-61.

Nakayashiki H (2005) RNA silencing in fungi: mechanisms and applications. FEBS Lett 579: 5950–7.

Nakayashiki H (2011) The Trickster in the genome: contribution and control of transposable elements. Genes Cells 16: 827-41.

Nakazaki T, Okumoto Y, Horibata A, Yamahira S, Teraishi M, et al. (2003) Mobilization of a transposon in the rice genome. Nature 421: 170-2.

Nakshatri H, Bhat–Nakshatri P, Currie RA (1996) Subunit association and DNA binding activity of the heterotrimeric transcription factor NF-Y is regulated by cellular redox. J Biol Chem 271: 28784–91.

Nalbandian A, Dettin L, Dym M, Ravindranath N (2003) Expression of vascular endothelial growth factor receptors during male germ cell differentiation in the mouse. Biol Reprod 69: 985–94.

Nam SY, Sabapathy K (2011) p53 promotes cellular survival in a context-dependent manner by directly inducing the expression of haeme-oxygenase-1. Oncogene 30: 4476-86.

Nanogaki T, Noda Y, Narimoto K, Shiotani M, Mon T, et al. (1992) Localization of Cu/Zn-superoxide dismutase in the human male genital organs. Hum Reprod 7: 81-5.

Napolitano R, Janel-Bintz R, Wagner J, Fuchs RP (2000) All three SOS-inducible DNA polymerases (Pol II, Pol IV and Pol V) are involved in induced mutagenesis. EMBO J 19: 6259–65.

Nardinocchi L, Puca R, D’Orazi G (2011) HIF-1α antagonizes p53-mediated apoptosis by triggering HIPK2 degradation. Aging (Albany NY) 3: 33-43.

Narendra KS, Thathachar MAL (1974) Learning automata-a survey. IEEE Trans Syst Man Cybern SMC-4: 323-34.

Narendra KS, Thathachar MAL (1989) Learning automata: An introduction. Englewood Cliffs: Prentice-Hall.

Narisawa S, Hecht NB, Goldberg E, Boatright KM, Reed JC, Millan JL (2002) Testis-specific cytochrome c-null mice produce functional sperm but undergo early testicular atrophy. Mol Cell Biol 22: 5554–62.

Narra HP, Ochman H (2006) Of what use is sex to bacteria? Curr Biol 16: R705–R710.

Narsinh KH, Sun N, Sanchez-Freire V, Lee AS, Almeida P, et al. (2011) Single cell transcriptional profiling reveals heterogeneity of human induced pluripotent stem cells. J Clin Invest 121: 1217-21.

Nasution O, Srinivasa K, Kim M, Kim YJ, Kim W, et al. (2008) Hydrogen peroxide induces hyphal differentiation in Candida albicans. Eukaryot Cell 7: 2008-11.

Nathanson M (1975) The effect of resource limitation on competing  populations of flour beetles, Tribolium spp. (Coleoptera, Tenebrionidae). Bull Entomol Res 65: 1–12.

Nätt D (2011) Heritable epigenetic responses to environmental challenges: Effects on behaviour, gene expression and DNA-methylation in the chicken. Thesis. University of Linköping, Sweden.

Navarro RE, Shim EY, Kohara Y, Singson A, Blackwell TK (2001) cgh-1, a conserved predicted RNA helicase required for gametogenesis and protection from physiological germline apoptosis in C. elegans. Development 128: 3221-32.

Naviaux RK (2008) Mitochondrial control of epigenetics. Cancer Biol Ther 7: 1191-3.

Nedelcu AM (2005) Sex as a response to oxidative stress: stress genes co-opted for sex. Proc Biol Sci 272: 1935-40.

Nedelcu AM, Michod RE (2003) Sex as a response to oxidative stress: the effect of antioxidants on sexual induction in a facultatively sexual lineage. Proc R Soc Lond B 270: S136–S139.

Nedelcu AM, Marcu O, Michod RE (2004) Sex as a response to oxidative stress: a twofold increase in cellular reactive oxygen species activates sex genes. Proc R Soc Lond B 271: 1591–6.

Nee S (1989) Antagonistic co-evolution and the evolution of genotypic randomization. J Theor Biol 140: 499-518.

Neel JV (1941) A relation between larval nutrition and the frequency of crossing over in the third chromosome of Drosophila melanogaster. Genetics 26: 506–16.

Neel JV (1942) A study of a case of high mutation rate in Drosophila melanogaster. Genetics 27: 519-36.

Neel JV (1962) Diabetes mellitus: a "thrifty" genotype rendered detrimental by "progress"? Am J Hum Genet 14: 353-62.

Neeman M, Abramovitch R, Schiffenbauer YS, Tempel C (1997) Regulation of angiogenesis by hypoxic stress: from solid tumours to the ovarian follicle. Int J Exp Pathol 78: 57-70.

Neff BD, Pitcher TE (2005) Genetic quality and sexual selection: an integrated framework for good genes and compatible genes. Mol Ecol 14: 19–38.

Negovetic S, Jokela J (2001) Life-history variation, phenotypic plasticity, and subpopulation structure in a freshwater snail. Ecology 82: 2805-15.

Nègre-Salvayre A, Hirtz C, Carrera G, Cazenave R, Troly M, et al. (1997) A role for uncoupling protein-2 as a regulator of mitochondrial hydrogen peroxide generation. FASEB J 11: 809–15.

Nehar D, Mauduit C, Boussouar F, Benahmed M (1997) Tumor necrosis factor-α-stimulated lactate production is linked to lactate dehydrogenase A expression and activity increase in porcine cultured Sertoli cells. Endocrinology 138: 1964–71.

Neher RA, Shraiman BI, Fisher DS (2010) Rate of adaptation in large sexual populations. Genetics 184: 467-81.

Nei M (2005) Selectionism and neutralism in molecular evolution. Mol Biol Evol 22: 2318–42.

Nei M, Graur D (1984) Extent of protein polymorphism and the neutral mutation theory. Evol Biol 17: 73–118.

Neill S, Desikan R, Hancock J (2002) Hydrogen peroxide signalling. Curr Opin Plant Biol 5: 388–95.

Neiman M, Koskella B (2009) Sex and the Red Queen. In: Schön I, Martens K, van Dijk P, eds. Lost sex. Amsterdam, The Netherlands: Springer. pp 133–159.

Neiman M, Taylor DR (2009) The causes of mutation accumulation in mitochondrial genomes. Proc Biol Sci 276: 1201-9.

Neiman M, Meirmans S, Meirmans PG (2009a) What can asexual lineage age tell us about the maintenance of sex? Ann NY Acad Sci 1168: 185–200.

Neiman M, Theisen KM, Mayry ME, Kay AD (2009b) Can phosphorus limitation contribute to the maintenance of sex? A test of a key assumption. J Evol Biol 22: 1359-63.

Neiman M, Hehman G, Miller JT, Logsdon JM Jr, Taylor DR (2010) Accelerated mutation accumulation in asexual lineages of a freshwater snail. Mol Biol Evol 27: 954-63.

Neiman M, Schwander T (2011) Using parthenogenetic lineages to identify advantages of sex. Evol Biol 38: 115–23.

Neiman M, Kay AD, Krist AC (2013) Can resource costs of polyploidy provide an advantage to sex? Heredity (Edinb) 110: 152-9.

Nelson JB (1978) The Sulidae: gannets and boobies. Oxford, UK: Oxford University Press.

Nelson MA (1996) Mating systems in ascomycetes: a romp in the sac. Trends Genet 12: 69-74.

Nelson RJ, Drazen DL (1999) Melatonin mediates seasonal adjustments in immune function. Reprod Nutr Dev 39: 383-98.

Nelson RL, Davis FG, Sutter E, Sobin LH, Kikendall JW, Bowen P (1994) Body iron stores and risk of colonic neoplasia. J Natl Cancer Inst 86: 455–60.

Nelson SK, Bose SK, McCord JM (1994) The toxicity of high-dose superoxide dismutase suggests that superoxide can both initiate and terminate lipid peroxidation in the reperfused heart. Free Radic Biol Med 16: 195-200.

Nelson VR, Nadeau JH (2010) Transgenerational genetic effects. Epigenomics 2: 797–806.

Nelson VR, Heaney JD, Tesar PJ, Davidson NO, Nadeau JH (2012) Transgenerational epigenetic effects of Apobec1 cytidine deaminase deficiency on testicular germ cell tumor susceptibility and embryonic viability. Proc Natl Acad Sci USA 109: E2766–73.

Nemetschke L, Eberhardt AG, Hertzberg H, Streit A (2010) Genetics, chromatin diminution, and sex chromosome evolution in the parasitic nematode genus Strongyloides. Curr Biol 20: 1687–96.

Nepomnaschy PA, Welch KB, McConnell DS, Low BS, Strassmann BI, England BG (2006) Cortisol levels and very early pregnancy loss in humans. Proc Natl Acad Sci USA 103: 3938-42.

Nesbit CE, Tersak JM, Prochownik EV (1999) MYC oncogenes and human neoplastic disease. Oncogene 18: 3004–16.

Nessler SH, Uhl G, Schneider JM (2007) Genital damage in the orbweb spider Argiope bruennichi (Araneae: Araneidae) increases paternity success. Behav Ecol 18: 174–81.

Neubauer K, Jewgenow K, Blottner S, Wildt DE, Pukazhenthi BS (2004) Quantity rather than quality in teratospermic males: a histomorphometric and flow cytometric evaluation of spermatogenesis in the domestic cat (Felis catus). Biol Reprod 71: 1517–24.

Neubert MG, Caswell H (2000) Density-dependent vital rates and their population dynamic consequences. J Math Biol 41: 103–21.

Neuer A, Spandorfer SD, Giraldo P, Dieterle S, Rosenwaks Z, Witkin SS (2000) The role of heat shock proteins in reproduction. Hum Reprod Update 6: 149–59.

Neuhäuser M, Krackow S (2007) Adaptive-filtering of trisomy 21: risk of Down syndrome depends on family size and age of previous child. Naturwissenschaften 94: 117-21.

Nevers P, Saedler H (1977) Transposable genetic elements as agents of gene instability and chromosomal rearrangements. Nature 268: 109-15.

Nevo E (1988) Genetic diversity in nature: Patterns and theory. Evol Biol 23: 217–47.

Nevo E (1989) Modes of specialisation: the nature and role of peripheral isolates in the origin of species. In: Giddings LV, Kaneshiro KY, Anderson WW, eds. Genetics, speciation, and the founder principle. Oxford, UK: Oxford University Press. pp 206-236.

Nevo E (1991) Evolutionary theory and processes of active speciation and adaptive radiation in subterranean mole rats, Spalax ehrenbergy superspecies in Israel. Evol Biol 25: 1-125.

Nevo E (1998) Molecular evolution and ecological stress at global, regional and local scales: The Israeli perspective. J Exp Zool 282: 95–119.

Nevo E (1999) Mosaic evolution of subterranean mammals. Regression, progression and global convergence. Oxford, UK: Oxford University Press.

Nevo E (2001) Evolution of genome–phenome diversity under environmental stress. Proc Natl Acad Sci USA 98: 6233-40.

Nevo E (2011) Evolution under environmental stress at macro- and microscales. Genome Biol Evol 2: 1039–52.

Nevo E, Kim YJ, Shaw CR, Thaeler CS Jr (1974) Genetic variation, selection and speciation in Thomomys talpoides pocket gophers. Evolution 28: 1-23.

Nevo E, Zohary D, Brown AHD, Haber M (1979) Genetic diversity and environmental associations of wild barley, Hordeum spontaneum, in Israel. Evolution 33: 815–33.

Nevo E, Beiles A, Ben-Shlomo R (1984) The evolutionary significance of genetic diversity: ecological, demographic and life history correlates. In: Mani GS, ed. Evolutionary dynamics of genetic diversity. Berlin, Germany: Springer-Verlag. pp 13–213.

Nevo E, Filippucci MG, Beiles A (1994) Genetic polymorphisms in subterranean mammals (Spalax ehrenbergi superspecies) in the Near East revisited: Patterns and theory. Heredity 72: 465-87.

Nevo E, Kirzhner V, Beiles A, Korol A (1997a) Selection versus random drift: long term polymorphism persistence in small populations (evidence and modelling). Philos Trans R Soc Lond Ser B 352: 381–9.

Nevo E, Apelbaum-El-Kaher I, Garty J, Beiles A (1997b) Natural selection causes microscale allozyme diversity in wild barley and lichen at Evolution Canyon, Mt. Carmel, Israel. Heredity 78: 373–82.

Nevo E, Baum B, Beiles A, Johnson DA (1998) Ecological correlates of RAPD DNA diversity of wild barley Hordeum spontaneum, in the Fertile Crescent. Genet Res Crop Evol 45: 151–9.

Nevoux M, Forcada J, Barbraud C, Croxall J, Weimerskirchi H (2010) Bet-hedging response to environmental variability, an intraspecific comparison. Ecology 91: 2416–27.

Newcomer SD, Zeh JA, Zeh DW (1999) Genetic benefits enhance the reproductive success of polyandrous females. Proc Natl Acad Sci USA 96: 10236–41.

Newman RA (1992) Adaptive plasticity in amphibian metamorphosis. BioScience 42: 671–8.

Newman RA (1994) Genetic variation for phenotypic plasticity in the larval life history of spadefoot toads (Scaphiopus couchii). Evolution 48: 1773-85.

Nezis IP, Stravopodis DJ, Papassideri I, Robert-Nicoud M, Margaritis LH (2000) Stage-specific apoptotic patterns during Drosophila oogenesis. Eur J Cell Biol 79: 610-20.

Nezis IP, Stravopodis DJ, Papassideri I, Margaritis LH (2001) Actin cytoskeleton reorganization of the apoptotic nurse cells during the late developmental stages of oogenesis in Dacus oleae. Cell Motil Cytoskeleton 48: 224-33.

Nezis IP, Stravopodis DJ, Papassideri I, Robert-Nicoud M, Margaritis LH (2002) The dynamics of apoptosis in the ovarian follicle cells during the late stages of Drosophila oogenesis. Cell Tissue Res 307: 401-9.

Nezis IP, Modes V, Mpakou V, Stravopodis DJ, Papassideri IS, et al. (2003) Modes of programmed cell death during Ceratitis capitata oogenesis. Tissue Cell 35: 113-9.

Nezis IP, Stravopodis DJ, Margaritis LH, Papassideri IS (2006a) Follicular atresia during Dacus oleae oogenesis. J Insect Physiol 52: 282-90.

Nezis IP, Stravopodis DJ, Margaritis LH, Papassideri IS (2006b) Programmed cell death of follicular epithelium during the late developmental stages of oogenesis in the fruit flies Bactrocera oleae and Ceratitis capitata (Diptera, Tephritidae) is mediated by autophagy. Dev Growth Differ 48:189-98.

Nezis IP, Stravopodis DJ, Margaritis LH, Papassideri IS (2006c) Autophagy is required for the degeneration of the ovarian follicular epithelium in higher Diptera. Autophagy 2: 297-8.

Ng JMY, Vrieling H, Sugasawa K, Ooms MP, Grootegoed JA, et al. (2002) Developmental defects and male sterility in mice lacking the ubiquitin-like DNA repair gene mHR23B. Mol Cell Biol 22: 1233-45.

Nguyen T, Nioi P, Pickett CB (2009) The Nrf2-antioxidant response element signaling pathway and its activation by oxidative stress. J Biol Chem 284: 13291–5.

Nichol ST, Rowe JE, Fitch WM (1993) Punctuated equilibrium and positive Darwinian evolution in vesicular stomatitis virus. Proc Natl Acad Sci USA 90: 10424-8.

Nicholls TJ, Goldsmith AR, Dawson A (1988a) Photorefractoriness in birds and comparison with mammals. Physiol Rev 68: 133–76.

Nicholls TJ, Follett BK, Goldsmith AR, Pearson H (1988b) Possible homologies between photorefractoriness in sheep and birds - the effect of thyroidectomy on the length of the ewes breeding-season. Reprod Nutr Dev 28: 375–85.

Nichols J, Zevnik B, Anastassiadis K, Niwa H, Klewe-Nebenius D, et al. (1998) Formation of pluripotent stem cells in the mammalian embryo depends on the POU transcription factor Oct4. Cell 95: 379–91.

Nickoloff JA (1992) Transcription enhances intrachromosomal homologous recombination in mammalian cells. Mol Cell Biol 12: 5311–8.

Nicoletto PF (1993) Female sexual response to condition-dependent ornaments in the guppy, Poecilia reticulata. Anim Behav 46: 441-50.

Nie Z, Hu G, Wei G, Cui K, Yamane A, et al. (2012) c-Myc is a universal amplifier of expressed genes in lymphocytes and embryonic stem cells. Cell 151: 68–79.

Nielsen KM (1998) Barriers to horizontal gene transfer by natural transformation in soil bacteria. APMIS 106 (Suppl 84): 77-84.

Nielsen R (2001) Mutations as missing data: inferences on the ages and distributions of nonsynonymous and synonymous mutations. Genetics 159: 401–11.

Nielsen R, Weinreich DM (1999) The age of nonsynonymous and synonymous mutations in animal mtDNA and implications for the mildly deleterious theory. Genetics 153: 497–506.

Nielsen R, Hubisz MJ, Hellmann I, Torgerson D, Andrés AM, et al. (2009) Darwinian and demographic forces affecting human protein coding genes. Genome Res 19: 838-49.

Niemeitz A, Kreutzfeldt R, Schartl M, Parzefall J, Schlupp I (2002) Male mating behaviour of a molly, Poecilia latipunctata: a third host for the sperm-dependent Amazon molly, Poecilia formosa. Acta Ethol 5: 45–9.

Niemi M, Sharpe RM, Brown WRA (1986) Macrophages in the interstitial tissue of the rat testis. Cell Tissue Res 243: 337-44.

Niess AM, Passek F, Lorenz I, Schneider EM, Dickhuth HH, et al. (1999) Expression of the antioxidant stress protein heme oxygenase-1 (HO-1) in human leukocytes. Free Radic Biol Med 26: 184–92.

Nigro L, Prout T (1990) Is there selection on RFLP differences in mitochondrial DNA? Genetics 125: 551-5.

Nii T, Marumoto T, Tani K (2012) Roles of p53 in various biological aspects of hematopoietic stem cells. J Biomed Biotechnol 2012: 903435.

Nijhout HF (1999) Control mechanisms of polyphenic development in insects—in polyphenic development, environmental factors alter same aspects of development in an orderly and predictable way. Bioscience 49: 181–92.

Niklas KJ (1997) The evolutionary biology of plants. Chicago, IL: University of Chicago Press.

Niklasson M, Parker ED Jr (1994) Fitness variation in an invading parthenogenetic cockroach. Oikos 71: 47–54.

Niklasson M, Tomiuk J, Parker ED Jr (2004) Maintenance of clonal diversity in Dipsa bifurcate (Fallén, 1810) (Diptera: Lonchopteridae). I. Fluctuating seasonal selection moulds long-term coexistence. Heredity 93: 62–71.

Nilsen J (2008) Estradiol and neurodegenerative oxidative stress. Front Neuroendocrinol 29: 463-75.

Nilsson M, Snoad N (2002) Optimal mutation rates in dynamic environments. Bull Math Biol 64: 1033-43.

Nirupama M, Devaki M, Nirupama R, Yajurvedi HN (2013) Chronic intermittent stress-induced alterations in the spermatogenesis and antioxidant status of the testis are irreversible in albino rat. J Physiol Biochem 69: 59-68.

Nishi T, Shimizu N, Hiramoto M, Sato I, Yamaguchi Y, et al. (2002) Spatial redox regulation of a critical cysteine residue of NF-kappa B in vivo. J Biol Chem 277: 44548–56.

Nishimura Y, Yoshinari T, Naruse K, Yamada T, Sumi K, et al. (2006) Active digestion of sperm mitochondrial DNA in single living sperm revealed by optical tweezers. Proc Natl Acad Sci USA 103: 1382–7.

Nishizawa Y, Fushiki S, Amakata Y, Nishizawa Y (1998) Thyroxine-induced production of superoxide anion by human alveolar neutrophils and macrophages: a possible mechanism for the exacerbation of bronchial asthma with the development of hyperthyroidism. In Vivo 12: 253–7.

Noble SM, Johnson AD (2007) Genetics of Candida albicans, a diploid human fungal pathogen. Annu Rev Genet 41: 193–211.

Noce T, Okamoto-Ito S, Tsunekawa N (2001) Vasa homolog genes in mammalian germ cell development. Cell Struct Funct 26: 131–6.

Noel F, Machon N, Porcher E (2007) No genetic diversity at molecular markers and strong phenotypic plasticity in populations of Ranunculus nodiflorus, an endangered plant species in France. Ann Bot 99: 1203-12.

Nogrady T, Wallace RL, Snell TW (1993) Rotifera 1: biology, ecology and systematics. In: Nogrady T, ed. Guides to the Identification of the Microinvertebrates of the Continental Waters of theWorld. Amsterdam, The Netherlands: SPB Academic Publishing BV. pp 1–137.

Nohl H (1994) Generation of superoxide radicals as byproducts of cellular respiration. Ann Biol Clin 52: 199-204.

Nohl H, Staniek K, Sobhian B, Bahrami S, Redl H, Kozlov AV (2000) Mitochondria recycle nitrite back to the bioregulator nitric monoxide. Acta Biochim Pol 47: 913–21.

Nohmi T (2006) Environmental stress and lesion-bypass DNA polymerases. Annu Rev Microbiol 60: 231–53.

Nollen EA, Morimoto RI (2002) Chaperoning signaling pathways: molecular chaperones as stress-sensing ‘heat shock’ proteins. J Cell Sci 115: 2809–16.

Noma K-I, Sugiyama T, Cam H, Verdel A, Zofall M, et al. (2004) RITS acts in cis to promote RNA interference-mediated transcriptional and post-transcriptional silencing. Nat Genet 36: 1174–80.

Nonoguchi K, Tokuchi H, Okuno H, Watanabe H, Egawa H, et al. (2001) Expression of Apg-1, a member of the Hsp110 family, in the human testis and sperm. Int J Urol 8: 308–14.

Noor MA(2008) Connecting recombination, nucleotide diversity and species divergence in Drosophila. Fly 2: 1–2.

Nordström KJV, Mirza MAI, Sällman Almén M, Gloriam DE, Fredriksson R, Schiöth HB (2009) Critical evaluation of the FANTOM3 non-coding RNA transcripts. Genomics 94: 169–76.

Norimura T, Nomoto S, Katsuki M, Gondo Y, Kondo S (1996) p53-dependent apoptosis suppresses radiation-induced teratogenesis. Nat Med 2: 577–80.

Norman M, Wisniewska KA, Lawrenson K, Garcia-Miranda P, Tada M, et al. (2012) Loss of Scribble causes cell competition in mammalian cells. J Cell Sci 125: 59-66.

Normand S, Treier UA, Randin C, Vittoz P, Guisan A, Svenning JC (2009) Importance of abiotic stress as a range-limit determinant for European plants: insights from species responses to climatic gradients. Global Ecol Biogeogr 18: 437-49.

Normark BB (1999) Evolution in a putatively ancient asexual aphid lineage: recombination and rapid karyotype change. Evolution 53: 1458–69.

Normark BB (2003) The evolution of alternative genetic systems in insects. Annu Rev Entomol 48: 397–423.

Normark BB, Moran NA (2000) Testing for the accumulation of deleterious mutations in asexual eukaryote genomes using molecular sequences. J Nat Hist 34: 1719–29.

Norris K (1993) Heritable variation in a plumage indicator of viability in male great tits Parus major. Nature 362: 537-9.

Norris KH, Hornsby PJ (1990) Cytotoxic effects of expression of human superoxide dismutase in bovine adrenocortical cells. Mutat Res 237: 95-106.

Norwitz ER, Schust DJ, Fisher SJ (2001) Implantation and the survival of early pregnancy. N Engl J Med 345: 1400–8.

Nosil P, Crespi BJ (2006) Experimental evidence that predation promotes divergence in adaptive radiation. Proc Natl Acad Sci USA 103: 9090–5.

Nöthel H (1987) Adaptation of Drosophila melanogaster populations to high mutation pressure: Evolutionary adjustment of mutation rates. Proc Natl Acad Sci USA 84: 1045–9.

Notley-McRobb L, Pinto R, Seeto S, Ferenci T (2002a) Regulation of mutY and nature of mutator mutations in Escherichia coli populations under nutrient limitation. J Bacteriol 184: 739–45.

Notley-McRobb L, Seeto S, Ferenci T (2002b) Enrichment and elimination of mutY mutators in Escherichia coli populations. Genetics 162: 1055–62.

Novella IS (2004) Negative effect of genetic bottlenecks on the adaptability of vesicular stomatitis virus. J Mol Biol 336: 61–7.

Novella IS, Elena SF,Moya A, Domingo E,Holland JJ (1995) Size of genetic bottlenecks leading to virus fitness loss is determined by mean initial population fitness. J Virol 69: 2869–72.

Novella IS, Quer J, Domingo E, Holland JJ (1999) Exponential fitness gains of RNA virus populations are limited by bottleneck effects. J Virol 73: 1668–71.

Novella IS, Ebendick-Corpus BE (2004) Molecular basis of fitness loss and fitness recovery in vesicular stomatitis virus. J Mol Biol 342: 1423–30.

Novella IS, Zárate S, Metzgar D, Ebendick-Corpus BE (2004) Positive selection of synonymous mutations in Vesicular stomatitis virus. J Mol Biol 342: 1415-21.

Nover L, Scharf KD (1997) Heat stress proteins and transcription factors. Cell Mol Life Sci 53: 80–103.

Novick A, Szilard L (1950) Experiments with the chemostat on spontaneous mutation of bacteria. Proc Natl Acad Sci USA 36: 708–19.

Nowacki M, Shetty K, Landweber LF (2011) RNA-mediated epigenetic programming of genome rearrangements. Annu Rev Genomics Hum Genet 12: 367–89.

Nowak MA (1992) What is a quasi-species? Trends Ecol Evol 7: 118-21.

Nowak MA, Anderson RM, Mc Lean AR, Wolfs TF, Goudsmit J, May RM (1991) Antigenic diversity thresholds and the development of AIDS. Science 254: 963–9.

Nowak MA, Bonhoeffer S, May RM (1994) Spatial games and the maintenance of cooperation. Proc Natl Acad Sci USA 91: 4877–81.

Nowell PC (1976) The clonal evolution of tumor cell populations. Science 194: 23-8.

Nowicki MO, Falinski R, Koptyra M, Slupianek A, Stoklosa T, et al. (2004) BCR/ABL oncogenic kinase promotes unfaithful repair of the reactive oxygen species-dependent DNA double strand breaks. Blood 104: 3746–53.

Nuismer SL, Thompson JN (2001) Plant polyploidy and non-uniform effects on insect herbivores. Proc R Soc Lond B 268: 1937–40.

Nuismer SL, Otto SP (2004) Host–parasite interactions and the evolution of ploidy. Proc Natl Acad Sci USA 101: 11036–9.

Nunn GB, Stanley SE (1998) Body size effects and rates of cytochrome b evolution in tube-nosed seabirds. Mol Biol Evol 15: 1360–71.

Nunney L (1990) Drosophila on oranges: colonization, competition and coexistence. Ecology 71: 1904–15.

Nunney L (1993) The influence of mating system and overlapping generations on effective population size. Evolution 47: 1329–41.

Nunney L (1996) The influence of variation in female fecundity on effective population size. Biol J Linn Soc 59: 411–25.

Nunney L (2002) The effective size of annual plant populations: the interaction of a seed bank with fluctuating population size in maintaining genetic variation. Am Nat 160: 195–204.

Nunoshiba T (1996) Two-stage gene regulation of the superoxide stress response soxRS system in Escherichia coli. Crit Rev Eukaryot Gene Expr 6: 377-89.

Nursall JR (1959) Oxygen as a prerequisite to the origin of the Metazoa. Nature 183: 1170–2.

Nussey DH, Postma E, Gienapp P, Visser ME (2005) Selection on heritable phenotypic plasticity in a wild bird population. Science 310: 304-6.

Nutt LK, Margolis SS, Jensen M, Herman CE, Dunphy WG, et al. (2005) Metabolic regulation of oocyte cell death through the CaMKII-mediated phosphorylation of caspase-2. Cell 123: 89–103.

Oakberg EF (1956) A description of spermiogenesis in the mouse and its use in analysis of the cycle of the seminiferous epithelium and germ cell renewal. Am J Anat 99: 391–413.

Oakberg EF (1959) Initial depletion and subsequent recovery of spermatogonia of the mouse after 20 R of gamma rays and 100, 300, and 600 R of X-rays. Radiat Res 11: 700–19.

Oakland E (1956) A description of spermatogenesis in the mouse and its use in analysis of the cycle of seminiferous epithelium and germ cell renewal. Am J Anat 99: 391-413.

Obbard DJ, Jiggins FM, Halligan DL, Little TJ (2006) Natural selection drives extremely rapid evolution in antiviral RNAi genes. Curr Biol 16: 580–5.

Obbard DJ, Gordon KH, Buck AH, Jiggins FM (2009) The evolution of RNAi as a defence against viruses and transposable elements. Philos Trans R Soc Lond B Biol Sci 364: 99-115.

Ober C (1999) Studies of HLA, fertility and mate choice in a human isolate. Hum Reprod Update 5: 103–7.

Ober C, Elias S, Kostyu DD, Hauck WW (1992) Decreased fecundability in Hutterite couples sharing HLA-DR. Am J Hum Genet 50: 6–14.

Oberley LW, Oberley TD, Buettner GR (1981) Cell division in normal and transformed cells: the possible role of superoxide and hydrogen peroxide. Med Hypotheses 7: 21–42.

Obermann WM, Sondermann H, Russo AA, Pavletich NP, Hartl FU (1998) In vivo function of Hsp90 is dependent on ATP binding and ATP hydrolysis. J Cell Biol 143: 901-10.

O’Brien EL, Dawson RD (2007) Context-dependent genetic benefits of extra-pair mate choice in a socially monogamous passerine. Behav Ecol Sociobiol 61: 775–82.

O’Bryan MK, Schlatt S, Gerdprasert O, Phillips DJ, de Kretser DM, Hedger MP (2000) Inducible nitric oxide synthase in the rat testis: evidence from potential roles of both normal function and inflammation-mediated infertility. Biol Reprod 63: 1285–93.

Ocal P, Aydin S, Cepni I, Idil S, Idil M, et al. (2004) Follicular fluid concentrations of vascular endothelial growth factor, inhibin A and inhibin B in IVF cycles: are they markers for ovarian response and pregnancy outcome? Eur J Obstet Gynecol Reprod Biol 115: 194–9.

Ochman H, Lawrence JG, Groisman EA (2000) Lateral gene transfer and the nature of bacterial innovation. Nature 405: 299–304.

Ochman HB, Stille B, Niklasson M, Selander RK (1980) Evolution of clonal diversity in the parthenogenetic fly Lonchoptera dubia. Evolution 34: 539–47.

Ochoa G, Jaffé K (1999) On sex, mate selection and the red queen. J Theor Biol 199: 1-9.

O’Connor CM, Gilmour KM, Arlinghaus R, Matsumura S, Suski CD, et al. (2011) The consequences of short-term cortisol elevation on individual physiology and growth rate in wild largemouth bass (Micropterus salmoides). Can J Fish Aquat Sci 68: 693–705.

O'Connor RJ (1978) Brood reduction in birds: Selection for fratricide, infanticide and suicide? Anim Behav 26: 79-96.

O’Donnell KA, Wentzel EA, Zeller KI, Dang CV, Mendell JT (2005) c-Myc-regulated microRNAs modulate E2F1 expression. Nature 435: 839–43.

O’Donnell L, McLachlan RI, Wreford NG, Robertson DM (1994) Testosterone promotes the conversion of round spermatids between stages VII and VIII of the rat spermatogenic cycle. Endocrinology 135: 2608–14.

O’Donnell L, McLachlan RI, Wreford NG, de Kretser DM, Robertson DM (1996) Testosterone withdrawal promotes stage-specific detachment of round spermatids from the rat seminiferous epithelium. Biol Reprod 55: 895–901.

Odorisio T, Rodriguez TA, Evans EP, Clarke AR, Burgoyne PS (1998) The meiotic checkpoint monitoring synapsis eliminates spermatocytes via p53-independent apoptosis. Nat Genet 18: 257–61.

Oehlen B, Cross FR (1994) Signal transduction in the budding yeast Saccharomyces cerevisiae. Curr Opin Cell Biol 6: 836-41.

Oertel M, Menthena A, Dabeva MD, Shafritz DA (2006) Cell competition leads to a high level of normal liver reconstitution by transplanted fetal liver stem/progenitor cells. Gastroenterology 130: 507–20.

Ogasawara K, Yabe R, Uchikawa M, Nakata K, Watanabe J, et al. (2001) Recombination and gene conversion-like events may contribute to ABO gene diversity causing various phenotypes. Immunogenetics 53: 190–9.

Ogunseitan OA (1995) Bacterial genetic exchange in nature. Sci Prog 78: 183-204.

Oh KP, Badyaev AV (2006) Adaptive genetic complementarity in mate choice coexists with selection for elaborate sexual traits. Proc R Soc Lond B Biol Sci 273: 1913–9.

Oh W, Gelardi NL, Cha CJ (1991) The cross-generation effect of neonatal macrosomia in rat pups of streptozotocin-induced diabetes. Pediatr Res 29: 606–10.

O’Halloran TV (1993) Transition metals in control of gene expression. Science 261: 715-25.

Ohinata Y, Seki Y, Payer B, O'Carroll D, Surani MA, Saitou M (2006)Germline recruitment in mice: a genetic program for epigenetic reprogramming.Ernst Schering Res Found Workshop 2006: 143-74.

Ohnishi O (1977) Spontaneous and ethyl methanesulfonate-induced mutations controlling viability in Drosophila melanogaster. II. Homozygous effect of polygenic mutations. Genetics 87: 529-45.

Ohnishi T, Mori E, Takahashi A (2009) DNA double-strand breaks: their production, recognition, and repair in eukaryotes. Mutat Res 669: 8-12.

Ohno M, Miura T, Furuichi M, Tominaga Y, Tsuchimoto D, et al. (2006) A genome-wide distribution of 8-oxoguanine correlates with the preferred regions for recombination and single nucleotide polymorphism in the human genome. Genome Res 16: 567–75.

Ohno S (1970) Evolution by gene duplication. New York, NY: Springer Verlag.

Ohno S (1996) The notion of the Cambrian pananimalia genome. Proc Natl Acad Sci USA 93: 8475-8.

Ohno S (1997a) The one ancestor per generation rule and three other rules of mitochondrial inheritance. Proc Natl Acad Sci USA 94: 8033-5.

Ohno S (1997b) The reason for as well as the consequence of the Cambrian explosion in animal evolution. J Mol Evol 44 (Suppl. 1): 823-7.

Ohsako S, Bunick D, Hayashi Y (1995) Immunocytochemical observation of the 90-kD heat shock protein (HSP90): high expression in primordial and pre-meiotic germ cells of male and female rat gonads. J Histochem Cytochem 43: 67-76.

Ohta M, Kadota C, Konishi H (1989) A role of melatonin in the initial stage of photoperiodism in the Japanese quail. Biol Reprod 40: 935–41.

Ohta T (1973) Slightly deleterious mutant substitutions in evolution. Nature 246: 96.

Ohta T (1992) The nearly neutral theory of molecular evolution. Annu Rev Ecol Syst 23: 263–86.

Ohta T (1993) Amino acid substitution at the Adh locus of Drosophila is facilitated by small population size. Proc Natl Acad Sci USA 90: 4548–51.

Ohta T (2011) Near-neutrality, robustness, and epigenetics. Genome Biol Evol 3: 1034–8.

Ohta T, Kimura M (1971) On the constancy of the evolutionary rate of cistrons. J Mol Evol 1:18–25.

Ohye T, Inagaki H, Kogo H, Tsutsumi M, Kato T, et al. (2010) Paternal origin of the de novo constitutional t(11;22)(q23;q11). Eur J Hum Genet 18: 783–7.

Oikawa S, Kawanishi S (1996) Site-specific DNA damage induced by NADH in the presence of copper(II): role of active oxygen species. Biochemistry 35: 4584-90.

Ojeda NB, Hennington BS, Williamson DT, Hill ML, Betson NE, et al. (2012) Oxidative stress contributes to sex differences in blood pressure in adult growth-restricted offspring. Hypertension 60: 114-22.

Ojosnegros S, Perales C, Mas A, Domingo E (2011) Quasispecies as a matter of fact: viruses and beyond. Virus Res 162: 203-15.

Okada H, Hirota Y, Moriyama R, Saga Y, Yanagisawa K (1986) Nuclear fusion in multinucleated giant cells during the sexual development of Dictyostelium discoideum. Dev Biol 118: 95-102.

Okada K, Blount JD, Sharma MD, Snook RR, Hosken DJ (2011) Male attractiveness, fertility and susceptibility to oxidative stress are influenced by inbreeding in Drosophila simulans. J Evol Biol 24: 363–71.

Okada T, Noji S, Goto Y, Iwata T, Fujita T, et al. (2005) Immune responses to DNA mismatch repair enzymes hMSH2 and hPMS1 in patients with pancreatic cancer, dermatomyositis and polymyositis. Int J Cancer 116: 925-33.

Okada Y, Yamagata K, Hong K, Wakayama T, Zhang Y (2010) A role for the elongator complex in zygotic paternal genome demethylation. Nature 463: 554-8.

Okamoto K, Tanaka H, Ogawa H, Makino Y, Eguchi H, et al. (1999) Redox-dependent regulation of nuclear import of the glucocorticoid receptor. J Biol Chem 274: 10363–71.

Okano M, Xie S, Li E (1998a) Dnmt2 is not required for de novo and maintenance methylation of viral DNA in embryonic stem cells. Nucleic Acids Res 26: 2536–40.

Okano M, Xie S, Li E (1998b) Cloning and characterization of a family of novel mammalian DNA (cytosine-5) methyltransferases. Nat Genet 19: 219–20.

Okano M, Bell DW, Haber DA, Li E (1999) DNA methyltransferases Dnmt3a and Dnmt3b are essential for de novo methylation and mammalian development. Cell 99: 247-57.

Okasha S (2006) Evolution and the levels of selection. New York: Oxford University Press.

Okazaki N, Okazaki K, Watanabe Y, Kato-Hayashi M, Yamamoto M, Okayama H (1998) Novel factor highly conserved among eukaryotes controls sexual development in fission yeast. Mol Cell Biol 18: 887-95.

Okita K, Yamanaka S (2011) Induced pluripotent stem cells: opportunities and challenges. Philos Trans R Soc Lond B Biol Sci 366: 2198–207.

Okladnova O, Syagailo YV, Tranitz M, Stöber G, Riederer P, et al. (1998) A promoter-associated polymorphic repeat modulates PAX-6 expression in human brain. Biochem Biophys Res Comm 248: 402-5.

Old LJ (2001) Cancer/testis (CT) antigens - a new link between gametogenesis and cancer. Cancer Immun 1: 1.

Oldenburg J, Schwaab R, Grimm T, Zerres K, Hakenberg P, et al. (1993) Direct and indirect estimation of the sex ratio of mutation frequencies in hemophilia A. Am J Hum Genet 53: 1229-38.

Oldereid NB, De Angelis P, Wiger R, Clausen OP (2001) Expression of Bcl-2 family proteins and spontaneous apoptosis in normal human testis. Mol Hum Reprod 7: 403–8.

Oleinick NL, Balasubramaniam U, Xue L, Chiu S (1994) Nuclear structure and the microdistribution of radiation damage in DNA. Int J Radiat Biol 66: 523-9.

Olinski R, Rozalski R, Gackowski D, Foksinski M, Siomek A, Cooke MS (2006) Urinary measurement of 8-OxodG, 8-OxoGua, and 5HMUra: a noninvasive assessment of oxidative damage to DNA. Antioxid Redox Signal 8: 1011-9.

Oliveira CA, Carnes K, Franca LR, Hess RA (2001) Infertility and testicular atrophy in the antiestrogen-treated adult male rat. Biol Reprod 65: 913–20.

Oliveira-Marques V, Marinho HS, Cyrne L, Antunes F (2009) Role of hydrogen peroxide in NF-kappaB activation: from inducer to modulator. Antioxid Redox Signal 11: 2223-43.

Oliver A, Canton R, Campo P, Baquero F, Blazquez J (2000) High frequency of hypermutable Pseudomonas aeruginosa in cystic fibrosis lung infection. Science 288: 1251–4.

Oliver ER, Saunders TL, Tarle SA, Glaser T (2004) Ribosomal protein L24 defect in belly spot and tail (Bst), a mouse Minute. Development 131: 3907–20.

Oliver JH Jr, Herrin CS (1976) Differential variation of parthenogenetic and bisexual Haemaphysalis longicornis (Acari: Ixodidae). J Parasitol 62: 475-84.

Oliver KR, Greene WK (2009a) Transposable elements: powerful facilitators of evolution. Bioessays 31: 703-14.

Oliver KR, Greene WK (2009b) The Genomic Drive hypothesis and punctuated evolutionary taxonations, or radiations. J R Soc WA 92: 447–51.

Oliver KR, Greene WK (2011) Mobile DNA and the TE-Thrust Hypothesis: supporting evidence from the primates. Mobile DNA 2: 8.

Oliver KR, Greene WK (2012) Transposable elements and viruses as factors in adaptation and evolution: an expansion and strengthening of the TE-Thrust hypothesis. Ecol Evol 2: 2912-33.

Oliver PL, Goodstadt L, Bayes JJ, Birtle Z, Roach KC, et al. (2009) Accelerated evolution of the Prdm9 speciation gene across diverse metazoan taxa. PLoS Genet 5: e1000753.

Olivieri D, Sykora MM, Sachidanandam R, Mechtler K, Brennecke J (2010) An in vivo RNAi assay identifies major genetic and cellular requirements for primary piRNA biogenesis in Drosophila. EMBO J 29: 3301–17.

Olivo PD, Van de Walle MJ, Laipis PJ, Hauswirth WW (1983) Nucleotide sequence evidence for rapid genotypic shifts in the bovine mitochondrial DNA D-loop. Nature 306: 400–2.

Ollmann M, Young LM, Di Como CJ, Karim F, Belvin M, et al. (2000) Drosophila p53 is a structural and functional homolog of the tumor suppressor p53. Cell 101: 91–101.

Olmsted SS, Dubin NH, Cone RA, Moench TR (2000) The rate at which human sperm are immobilized and killed by mild acidity. Fertil Steril 73: 687–93.

Olofsson H, Ripa J, Jonzén N (2009) Bet-hedging as an evolutionary game: the trade-off between egg size and number. Proc Biol Sci 276: 2963-9.

Olovnikov IA, Kravchenko JE, Chumakov PM (2009) Homeostatic functions of the p53 tumor suppressor: regulation of energy metabolism and antioxidant defense. Semin Cancer Biol 19: 32-41.

Olsen A-K, Bjørtuft H, Wiger R, Holme JA, Seeber EC, et al. (2001) Highly efficient base excision repair (BER) in human and rat male germ cells. Nucleic Acids Res 29: 1781–90.

Olsen A-K, Lindeman B, Wiger R, Duale N, Brunborg G (2005) How do male germ cells handle DNA damage? Toxicol Appl Pharmacol 207: 521–31.

Olshan AF, van Wijngaarden E (2003) Paternal occupation and childhood cancer. Adv Exp Med Biol 518: 147–61.

Olsson M, Shine R, Madsen T, Gullberg A, Tegelstrom H (1996) Sperm selection by females. Nature 383: 585.

Olsson M, Madsen T, Nordby J, Wapstra E, Ujvari B, Wittsell H (2003) Major histocompatibility complex and mate choice in sand lizards. Proc R Soc Lond B 270: S254-S256.

Olsson M, Wilson M, Uller T, Mott B, Isaksson C, et al. (2008) Free radicals run in lizard families. Biol Lett 4: 186–8.

Olszewska M, Bujarski JJ, Kurpisz M (2012) P-bodies and their functions during mRNA cell cycle: mini-review. Cell Biochem Funct 30: 177-82.

Omichinski JG, Trainor C, Evans T, Gronenborn AM, Clore GM, Felsenfeld G (1993) A small single-“finger” peptide from the erythroid transcription facor GATA-1 binds specifically to DNA as a zinc or iron complex. Proc Natl Acad Sci USA 90: 1676-80.

Omilian AR, Cristescu ME, Dudycha JL, Lynch M (2006) Ameiotic recombination in asexual lineages of Daphnia. Proc Natl Acad Sci USA 103:18638-43.

O’Neill JP, Finette BA (1998) Transition mutations at CpG dinucleotides are the most frequent in vivo spontaneous single-based substitution mutation in the human HPRT gene. Environ Mol Mutagen 32: 188–91.

Ono T, Okada S (1977) Radiation-induced DNA single-strand scission and its rejoining in spermatogonia and spermatozoa of mouse. MutatRes 43: 25-36.

Ono T, Ikehata H, Nakamura S, Saito Y, Hosoi Y, et al. (2000) Age-associated increase of spontaneous mutant frequency and molecular nature of mutation in newborn and old lacZ-transgenic mouse. Mutat Res 447: 165–77.

Ooi K, Yahara T (1999) Genetic variation of geminiviruses: comparison between sexual and asexual host plant populations. Mol Ecol 8: 89-97.

Oppenheimer JH, Schwartz HL, Strait KA (1996) The molecular basis of thyroid hormone action. In: Braverman LE, Utiger RD, eds. Werner and Ingbar’s The Thyroid. A Fundamental and Clinical Text. 7th edn. New York, NY: Lippincott-Raven. pp 162-184.

Orazizadeh M, Khorsandi LS, Hashemitabar M (2010) Toxic effects of dexamethasone on mouse testicular germ cells. Andrologia 42: 247-53.

Orgel LE, Crick FH (1980) Selfish DNA: the ultimate parasite. Nature 284: 604–7.

Orlando C, Caldini AL, Barni T, Wood WG, Strasburger CJ, Natali A, Maver A, Forti G, Serio M (1985) Ceruloplasmin and transferrin in human seminal plasma: are they an index of seminiferous tubular function? Fertil Steril 43: 290-4.

Orlando EF, Guillette JL Jr (2001) A re-examination of variation associated with environmentally stressed organisms. APMIS 109: S178–S186.

O’Roak BJ, Vives L, Girirajan S, Karakoc E, Krumm N, et al. (2012) Sporadic autism exomes reveal a highly interconnected protein network of de novo mutations. Nature 485: 246–50.

Orom UA, Nielsen FC, Lund AH (2008) MicroRNA-10a binds the 5'UTR of ribosomal protein mRNAs and enhances their translation. Mol Cell 30: 460-71.

Orozco TJ, Wang JF, Keen CL (2003) Chronic consumption of a flavanol- and procyanindin-rich diet is associated with reduced levels of 8-hydroxy-2’-deoxyguanosine in rat testes. J Nutr Biochem 14: 104-10.

Orr HA (1998) The population genetics of adaptation: the distribution of factors fixed during adaptive evolution. Evolution 52: 935–49.

Orr HA (1999) The evolutionary genetics of adaptation: a simulation study. Genet Res 74: 207–14.

Orr HA (2000a) The rate of adaptation in asexuals. Genetics 155: 961-8.

Orr HA (2000b) Adaptation and the cost of complexity. Evolution 54: 13–20.

Orr HA (2003) The distribution of fitness effects among beneficial mutations. Genetics 163: 1519-26.

Orr HA (2005a) The genetic theory of adaptation: a brief history. Nat Rev Genet 6: 119-27.

Orr HA (2005b) The genetic basis of reproductive isolation: insights from Drosophila. Proc Natl Acad Sci USA 102 Suppl 1: 6522-6.

Orr HA (2009) Fitness and its role in evolutionary genetics. Nat Rev Genet 10: 531-9.

Orr HA, Otto SP (1994) Does ploidy increase the rate of adaptation? Genetics 136: 1475–80.

Orr HA, Betancourt AJ (2001) Haldane’s sieve and adaptation from the standing genetic variation. Genetics 157: 875–84.

Orr HT (2009) Unstable nucleotide repeat minireview series: a molecular biography of unstable repeat disorders. J Biol Chem 284: 7405.

Orr HT, Zoghbi HY (2007) Trinucleotide repeat disorders. Annu Rev Neurosci 30: 575-621.

Orr TE, Mann DR (1992) Role of glucocorticoid in the stress-induced suppression of testicular steroidogenesis in adult male rats. Horm Behav 26: 350–63.

Orr TE, Taylor MF, Bhattacharyya AK, Collins DC, Mann DR (1994) Acute immobilization stress disrupts testicular steroidogenesis in adult male rats by inhibiting the activities of 17 alpha-hydroxylase and 17,20-lyase without affecting the binding of LH/hCG receptors. J Androl 15: 302-8.

Orrenius S (2007) Reactive oxygen species in mitochondria-mediated cell death. Drug Metab Rev 39: 443-55.

Orrenius S, Gogvadze V, Zhivotovsky B (2007) Mitochondrial oxidative stress: implications for cell death. Annu Rev Pharmacol Toxicol 47: 143–83.

Orsenigo S, Ricci C, Caprioli M (1998) The paradox of bdelloid egg size. Hydrobiologia 387/388: 317–20.

Ortega-Camarillo C, Guzman-Grenfell AM, Hicks JJ (1999) Oxidation of gonadotrophin (PMSG) by oxygen free radicals alters its structure and hormonal activity. Mol Reprod Dev 52: 264–8.

Orth JM, Gunsalus GL, Lamperti AA (1988) Evidence from Sertoli cell-depleted rats indicates that spermatid number in adults depends on numbers of Sertoli cells produced during perinatal development. Endocrinology 122: 787–94.

Ortiz A, Espino J, Bejarano I, Lozano GM, Monllor F, et al. (2011) High endogenous melatonin concentrations enhance sperm quality and short-term in vitro exposure to melatonin improves aspects of sperm motility. J Pineal Res 50: 132-9.

Ortogero N, Hennig GW, Langille C, Ro S, McCarrey JR, Yan W (2013) Computer-assisted annotation of murine Sertoli cell small RNA transcriptome. Biol Reprod 88: 3.

Orzack SH (1985) Population dynamics in variable environments. V. The genetics of homeostasis revisited. Am Nat 125: 550–72.

Orzack SH, Tuljapurkar S (1989) Population dynamics in variable environments. VII. The demography and evolution of iteroparity. Am Nat 133: 901–23.

Orzechowski A, Grizard J, Jank M, Gajkowska B, Lokociejewska M, et al. (2002) Dexamethasone-mediated regulation of death and differentiation of muscle cells. Is hydrogen peroxide involved in the process? Reprod Nutr Dev 42: 197–216.

Osborn HF (1896) A mode of evolution requiring neither natural selection nor the inheritance of acquired characteristics. Trans NY Acad Sci 15: 141–2.

Osborn TC, Pires JC, Birchler JA, Auger DL, Chen ZJ, et al. (2003) Understanding mechanisms of novel gene expression in polyploids. Trends Genet 19: 141-7.

Ossowski S, Schneeberger K, Lucas-Lledo JI, Warthmann N, Clark RM, et al. (2010) The rate and molecular spectrum of spontaneous mutations in Arabidopsis thaliana. Science 327: 92–4.

Östling P, Björk JK, Roos-Mattjus P, Mezger V, Sistonen L (2007) Heat shock factor 2 (HSF2) contributes to inducible expression of hsp genes through interplay with HSF1. J Biol Chem 282: 7077-86.

Östman O, Stuart-Fox D (2011) Sexual selection is positively associated with ecological generalism among agamid lizards. J Evol Biol 24:733-40.

Oswald BP, Nuismer SL (2007) Neopolyploidy and pathogen resistance. Proc Biol Sci 274: 2393-7.

Oswald J, Engemann S, Lane N, Mayer W, Olek A, et al. (2000) Active demethylation of the paternal genome in the mouse zygote. Curr Biol 10: 475–8.

Ota A, Tagawa H, Karnan S, Tsuzuki S, Karpas A, et al. (2004) Identification and characterization of a novel gene, C13orf25, as a target for 13q31-q32 amplification in malignant lymphoma. Cancer Res 64: 3087-95.

Otala M, Suomalainen L, Pentikäinen MO, Kovanen P, Tenhunen M, et al. (2004) Protection from radiation-induced male germ cell loss by sphingosine-1-phosphate. Biol Reprod 70: 759-67.

Otsuka M, Zheng M, Hayashi M, Lee JD, Yoshino O, et al. (2008) Impaired microRNA processing causes corpus luteum insufficiency and infertility in mice. J Clin Invest [Erratum (2008) 118: 2366] 118: 1944–54.

Otto CRV, Snodgrass JW, Forester DC, Mitchell JC, Miller RW (2007) Climatic variation and the distribution of an amphibian polyploid complex. J Anim Ecol 76: 1053–61.

Otto SP (2004) Two steps forward, one step back: the pleiotropic effects of favoured alleles. Proc Biol Sci 271: 705-14.

Otto SP (2007) Unravelling the evolutionary advantage of sex: a commentary on ‘‘Mutation-selection balance and the evolutionary advantage of sex and recombination’’ by Brian Charlesworth. Genet Res 89: 447–9.

Otto SP (2009) The evolutionary enigma of sex. Am Nat 174 Suppl 1: S1-S14.

Otto SP, Walbot V (1990) DNA methylation in eukaryotes: Kinetics of demethylation and de novo methylation during the life cycle. Genetics 124: 429–37.

Otto SP, Orive ME (1995) Evolutionary consequences of mutation and selection within an individual. Genetics 141: 1173–87.

Otto SP, Feldman MW (1997) Deleterious mutations, variable epistatic interactions, and the evolution of recombination. Theor Popul Biol 51: 134–47.

Otto SP, Barton NH (1997) The evolution of recombination: removing the limits to natural selection. Genetics 147: 879-906.

Otto SP, Hastings IM (1998) Mutation and selection within the individual. Genetics 102/103: 507–24.

Otto SP, Michalakis Y (1998) The evolution of recombination in changing environments. Trends Ecol Evol 13: 145–51.

Otto SP, Whitton J (2000) Polyploid incidence and evolution. Annu Rev Genet 34: 401–37.

Otto SP, Barton NH (2001) Selection for recombination in small populations. Evolution 55: 1921–31.

Otto SP, Lenormand T (2002) Resolving the paradox of sex and recombination. Nat Rev Genet 3: 252–61.

Otto SP, Nuismer SL (2004) Species interactions and the evolution of sex. Science 304: 1018–20.

Ottow B (1955) Biologische Anatomie der Genitalorgane und der Fortpflanzung der Säugetiere. Jena, Germany: Fischer.

Ou XM, Chen K, Shih JC (2006) Glucocorticoid and androgen activation of monoamine oxidase A is regulated differently by R1 and Sp1. J Biol Chem 281: 21512-25.

Outten FW (2007) Iron-sulfur clusters as oxygen-responsive molecular switches. Nat Chem Biol 3: 206–7.

Overington J, Donnelly D, Johnson MS, Sali A, Blundell TL (1992) Environment-specific amino acid substitution tables: Tertiary templates and prediction of protein folds. Protein Sci 1: 216–26.

Owen-Schaub LB, Zhang W, Cusack JC (1995) Wild-type human p53 and a temperature-sensitive mutant induce Fas/APO-1 expression. Mol Cell Biol 15: 3032–40.

Owusu-Ansah E, Banerjee U (2009) Reactive oxygen species prime Drosophila haematopoietic progenitors for differentiation. Nature 461: 537-41.

Ozaki M, Suzuki S, Irani K (2002) Redox factor-1/APE suppresses oxidative stress by inhibiting the rac1 GTPase. FASEB J 16: 889–90.

Ozanne SE, Sandovici I, Constância M (2011) Maternal diet, aging and diabetes meet at a chromatin loop. Aging (Albany NY) 3: 548-54.

Ozawa N, Goda N, Makino N, Yamaguchi T, Yoshimura Y, Suematsu M (2002) Leydig cell-derived heme oxygenase-1 regulates apoptosis of premeiotic germ cells in response to stress. J Clin Invest 109: 457-67.

Paaby AB, Schmidt PS (2008) Functional significance of allelic variation at methuselah, an aging gene in Drosophila. PLoS ONE 3: e1987.

Pabst DA, Rommel SA, McLellan WA, Williams TM, Rowles TK (1995) Thermoregulation of the intra-abdominal testes of the bottlenose dolphin (Tursiops truncatus) during exercise. J Exp Biol 198:221–6.

Pacchierotti F, Adler ID, Eichenlaub-Ritter U, Mailhes JB (2007) Gender effects on the incidence of aneuploidy in mammalian germ cells. Environ Res 104: 46-69.

Pace JK, Gilbert C, Clark MS, Feschotte C (2008) Repeated horizontal transfer of a DNA transposon in mammals and other tetrapods. Proc Natl Acad Sci USA 105: 17023–8.

Paciolla M, Boni R, Fusco F, Pescatore A, Poeta L, et al. (2011) Nuclear factor-kappa-B-inhibitor alpha (NFKBIA) is a developmental marker of NF-κB/p65 activation during in vitro oocyte maturation and early embryogenesis. Hum Reprod 26: 1191–201.

Packer MA, Porteous CM, Murphy MP (1996) Mitochondrial superoxide production in the presence of nitric oxide leads to the formation of peroxynitrite. Biochem Mol Biol Int 40: 527–34.

Padilla DK, Adolph SC (1996) Plastic inducible morphologies are not always adaptive: the importance of time delays in a stochastic environment. Evol Ecol 10: 105–17.

Paenke I, Kawecki TJ, Sendhoff B (2009) The influence of learning on evolution: A mathematical framework. Artif Life 15: 227–45.

Pagel M, Meade A (2004) A phylogenetic mixture model for detecting pattern heterogeneity in gene sequence or character-state data. Syst Biol 53: 571–81.

Pagel M, Venditti C, Meade A (2006) Large punctuational contribution of speciation to evolutionary divergence at the molecular level. Science 314: 119-21.

Pahl-Wostl C (1993) The hierarchical organization of the aquatic ecosystem: an outline how reductionism and holism may be reconciled. Ecol Model 66: 81–100.

Pál C, Miklós I (1999) Epigenetic inheritance, genetic assimilation and speciation. J Theor Biol 200: 19-37.

Pál C, Papp B, Hurst LD (2001) Highly expressed genes in yeast evolve slowly. Genetics 158: 927–31.

Pál C, Papp B, Lercher MJ (2006) An integrated view of protein evolution. Nat Rev Genet 7: 337-48.

Pal C, Maciá MD, Oliver A, Schachar I, Buckling A (2007) Coevolution with viruses drives the evolution of bacterial mutation rates. Nature 450: 1079-81.

Palacios-Callender M, Quintero M, Hollis V, Springett RJ, Moncada S (2004) Endogenous NO regulates superoxide production at low oxygen concentrations by modifying the redox state of cytochrome c oxidase. Proc Natl Acad Sci USA 101: 7630–5.

Palaima A (2007) The fitness cost of generalization: present limitations and future possible solutions. Biol J Linn Soc 90: 583–90.

Palaima A, Spitze K (2004) Is a jack-of-all-temperatures a master of none? an experimental test with Daphnia pulicaria (Crustacea: Cladocera). Evol Ecol Res 6: 215–25.

Paland S, Lynch M (2006) Transitions to asexuality result in excess amino acid substitutions. Science 311: 990–2.

Pal-Bhadra M, Leibovitch BA, Gandhi SG, Chikka MR, Bhadra U, et al. (2004) Heterochromatic silencing and HP1 localization in Drosphila are dependent on the RNAi machinery. Science 303: 669–72.

Palchevskiy V, Finkel SE (2006) Escherichia coli competence gene homologs are essential for competitive fitness and the use of DNA as a nutrient. J Bacteriol 188: 3902-10.

Palladino MA, Pirlamarla PR, McNamara J, Sottas CM, Korah N, et al. (2011) Normoxic expression of hypoxia-inducible factor 1 in rat Leydig cells in vivo and in vitro. J Androl 32: 307-23.

Pallardó FV, Degan P, d’Ischia M, Kelly FJ, Zatterale A, et al. (2006) Multiple evidence for an early age pro-oxidant state in Down syndrome patients. Biogerontology 7: 211-20.

Pallepati P, Averill-Bates D (2010) Mild thermotolerance induced at 40 degrees C increases antioxidants and protects HeLa cells against mitochondrial apoptosis induced by hydrogen peroxide: Role of p53. Arch Biochem Biophys 495: 97-111.

Palmen R, Vosman B, Buijsman P, Breek CK, Hellingwerf KJ (1993) Physiological characterization of natural transformation in Acinetobacter calcoaceticus. J Gen Microbiol 139: 295-305.

Palmer AR (2004) Symmetry breaking and the evolution of development. Science 306: 828-33.

Palmer SC, Norton RA (1992) Genetic diversity in thelytokous oribatid mites (Acari; Acariformes: Desmonomata). Biochem Syst Ecol 20: 219–31.

Palumbi SR (2001) Humans as the world’s greatest evolutionary force. Science 293: 1786–90.

Pamilo F, Nei M, Li WH (1987) Accumulation of mutations in sexual and asexual populations. Genet Res 49: 135–46.

Pan X, Zhou G, Wu J, Bian G, Lu P, et al. (2012) Wolbachia induces reactive oxygen species (ROS)-dependent activation of the Toll pathway to control dengue virus in the mosquito Aedes aegypti. Proc Natl Acad Sci USA 109: E23-31.

Panaretou B, Prodromou C, Roe SM, O'Brien R, Ladbury JE, et al. (1998) ATP binding and hydrolysis are essential to the function of the Hsp90 molecular chaperone in vivo. EMBO J 17: 4829-36.

Panaretou B, Siligardi G, Meyer P, Maloney A, Sullivan JK, et al. (2002) Activation of the ATPase activity of hsp90 by the stress-regulated cochaperone aha1. Mol Cell 10: 1307–18.

Panayiotidis MI, Rancourt RC, Allen CB, Riddle SR, Schneider BK, Ahmad S, White CW (2004) Hyperoxia-induced DNA damage causes decreased DNA methylation in human lung epithelial-like A549 cells. Antioxid Redox Signal 6: 129-36.

Pandey RR, Mondal T, Mohammad F, Enroth S, Redrup L, et al. (2008) Kcnq1ot1 antisense noncoding RNA mediates lineage-specific transcriptional silencing through chromatin-level regulation. Mol Cell 32: 232-46.

Pandi-Perumal SR, Srinivasan V, Maestroni GJM, Cardinali DP, Poeggeler B, Hardeland R (2006) Melatonin – Nature’s most versatile biological signal? FEBS J 273: 2813-38.

Pandian TJ, Koteeswaran R (1998) Ploidy induction and sex control in fish. Hydrobiologia 384: 167–243.

Pane A, Wehr K, Schüpbach T (2007) zucchini and squash encode two putative nucleases required for rasiRNA production in the Drosophila germline. Dev Cell 12: 851–62.

Pang AL, Johnson W, Ravindranath N, Dym M, Rennert OM, Chan WY (2006) Expression profiling of purified male germ cells: stage-specific expression patterns related to meiosis and postmeiotic development. Physiol Genomics 24: 75-85.

Pang S, Schlesinger Y, Daar ES, Moudgil T, Ho DD, Chen ISY (1992) Rapid generation of sequence variation during primary HIV-infection. AIDS 6: 453–60.

Pang SF, Li L, Ayre EA, Pang CS, Lee PP, et al. (1998) Neuroendocrinology of melatonin in reproduction: recent developments. J Chem Neuroanat 14: 157-66.

Panhuis TM, Butlin R, Zuk M, Tregenza T (2001) Sexual selection and speciation. Trends Ecol Evol 16: 364–71.

Pani G, Galeotti T (2011) Role of MnSOD and p66shc in mitochondrial response to p53. Antioxid Redox Signal 15:1715-27.

Pankhurst NW, Van Der Kraak G (2000) Evidence that acute stress inhibits ovarian steroidogenesis in rainbow trout in vivo, through the action of cortisol. Gen Comp Endocrinol 117: 225-37.

Pankow S, Bamberger C (2007) The p53 tumor suppressor-like protein nvp63 mediates selective germ cell death in the sea anemone Nematostella vectensis. PLoS ONE 2: e782.

Pannebakker BA, Pijnacker LP, Zwaan BJ, Beukeboom LW (2004) Cytology of Wolbachia-induced parthenogenesis in Leptopilina clavipes (Hymenoptera: Figitidae). Genome 47: 299–303.

Panopoulos A, Harraz M, Engelhardt JF, Zandi E (2005) Iron-mediated H2O2 production as a mechanism for cell type-specific inhibition of tumor necrosis factor alpha-induced but not interleukin-1beta-induced IkappaB kinase complex/nuclear factor-kappaB activation. J Biol Chem 280: 2912–23.

Pansarasa O, D’Antona G, Gualea MR, Marzani B, Pellegrino MA, Marzatico F (2002) “Oxidative stress”: effects of mild endurance training and testosterone treatment on rat gastrocnemius muscle. Eur J Appl Physiol 87: 550–5.

Pantano C, Shrivastava P, McElhinney B, Janssen-Heininger Y (2003) Hydrogen peroxide signaling through tumor necrosis factor receptor 1 leads to selective activation of c-Jun N-terminal kinase. J Biol Chem 278: 44091–6.

Panthier JJ, Guenet J, Condamine H, Jacob F (1990) Evidence for mitotic recombination in Wei/+ heterozygous mice. Genetics 125: 175-82.

Pantopoulos K, Hentze MW (1995) Rapid responses to oxidative stress mediated by iron regulatory protein. EMBO J 14: 2917-24.

Papaj DR (1994) Optimizing learning and its effect on evolutionary change in behavior. In: Real LA, ed. Behavioral mechanisms in evolutionary ecology. Chicago, IL:  University of Chicago Press. pp 133–153.

Papaioannou MD, Pitetti JL, Ro S, Park C, Aubry F, et al. (2009) Sertoli cell Dicer is essential for spermatogenesis in mice. Dev Biol 326: 250-9.

Papaioannou MD, Lagarrigue M, Vejnar CE, Rolland AD, Kuhne F, et al. (2011) Loss of Dicer in Sertoli cells has a major impact on the testicular proteome of mice. Mol Cell Proteomics 10: M900587–MCP200.

Papandreou I, Cairns RA, Fontana L, Lim AL, Denko NC (2006) HIF-1 mediates adaptation to hypoxia by actively downregulating mitochondrial oxygen consumption. Cell Metab 3: 187–97.

Papini A, Mosti S, Milocani E, Tani G, Di Falco P, Brighigna L (2011) Megasporogenesis and programmed cell death in Tillandsia (Bromeliaceae). Protoplasma 248: 651-62.

Papke RT, Koenig JE, Rodriguez-Valera F, Doolittle WF (2004) Frequent recombination in a saltern population of Halorubrum. Science 306: 1928–9.

Papp B, Teusink B, Notebaart RA (2009) A critical view of metabolic network adaptations. HFSP J 3: 24–35.

Pâques F, Leung WY, Haber JE (1998) Expansions and contractions in a tandem repeat induced by double-strand break repair. Mol Cell Biol 18: 2045–54.

Pâques F, Haber JE (1999) Multiple pathways of recombination induced by double-strand breaks in Saccharomyces cerevisiae. Microbiol Mol Biol Rev 63:349-404.

Pâques F, Richard GF, Haber JE (2001) Expansions and contractions in 36-bp minisatellites by gene conversion in yeast. Genetics 158: 155–66.

Paquin CE, Williamson VM (1984) Temperature effects on the rate of Ty transposition. Science 226: 53–5.

Paradkar PN, Zumbrennen KB, Paw BH, Ward DM, Kaplan J (2009) Regulation of mitochondrial iron import through differential turnover of mitoferrin 1 and mitoferrin 2. Mol Cell Biol 29: 1007-16.

Paranko J, Seitz J, Meinhardt A (1996) Developmental expression of heat shock protein-60 (HSP60) in the rat testis and ovary. Differentiation 60: 159-67.

Pariente N, Sierra S, Lowenstein PR, Domingo E (2001) Efficient virus extinction by combination of a mutagen and antiviral inhibitors. J Virol 75: 9723–30.

Parisi D, Nolfi S, Cecconi F (1992) Learning, behavior and evolution. In: Varela FJ, Bourgine P, eds. Toward a Practice of Autonomous Systems: Proceedings of the First European Conference on Artificial Life. Cambridge, MA: MIT Press/Bradford Books. pp 207-216.

Park AW, Jokela J, Michalakis Y (2010) Parasites and deleterious mutations: interactions influencing the evolutionary maintenance of sex. J Evol Biol 23: 1013-23.

Park H, Pontius W, Guet CC, Marko JF, Emonet T, Cluzel P (2010) Interdependence of behavioural variability and response to small stimuli in bacteria. Nature 468:819-23.

Park HO, Craig EA (1989) Positive and negative regulation of basal expression of a yeast HSP70 gene. Mol Cell Biol 9: 2025–33.

Park IH, Zhao R, West JA, Yabuuchi A, Huo H, et al. (2008) Reprogramming of human somatic cells to pluripotency with defined factors. Nature 451: 141-6.

Park SC, Krug J (2007) Clonal interference in large populations. Proc Natl Acad Sci USA 104: 18135–40.

Park SC, Simon D, Krug J (2010) The speed of evolution in large asexual populations. J Stat Phys 138: 381–410.

Park SG, Choi SS (2010) Expression breadth and expression abundance behave differently in correlations with evolutionary rates. BMC Evol Biol 10:241.

Park SY, Lee JH, Ha M, Nam JW, KimVN (2009) miR-29 miRNAs activate p53 by targeting p85α and CDC42. Nat Struct Mol Biol 16: 23–9.

Parker AR (1998) Colour in the Burgess shale animals and the effect of light on the evolution in the Cambrian. Proc R Soc Lond Ser B 265: 967–72.

Parker ED Jr (1979a) Ecological implications of clonal diversity in parthenogenetic morphospecies. Am Zool 19: 753–62.

Parker ED Jr (1979b) Phenotypic consequences of parthenogenesis in Cnemidophorus lizards. I. Variation in parthenogenetic and sexual populations. Evolution 33: 1150-66.

Parker ED, Selander RK, Hudson RO, Lester LJ (1977) Genetic diversity in colonizing parthenogenetic cockroaches. Evolution 31: 836–42.

Parker ED Jr, Niklasson M (1995) Desiccation resistance in invading parthenogenetic cockroaches: a search for the general purpose genotype. J Evol Biol 8: 331-7.

Parker ED Jr, Forbes VE, Nielsen SL, Ritter C, Barata C, et al. (1999) Stress in ecological systems. Oikos 86: 179–84.

Parker ED Jr, Niklasson M (2000) Genetic structure and evolution in parthenogenetic animals. In: Singh RS, Krimbas CB, eds. Evolutionary genetics: from molecules to morphology. Cambridge, UK: Cambridge University Press. pp 456-474.

Parker GA (1970) Sperm competition and its evolutionary consequences in the insects. Biol Rev 45: 525–67.

Parker GA (1982) Why are there so many tiny sperm? Sperm competition and the maintenance of two sexes. J Theor Biol 96: 281–94.

Parker GA (1990) Sperm competition games: raffles and roles. Proc R Soc Lond B 242: 120-6.

Parker GA (1998) Sperm competition and the evolution of ejaculates: towards a theory base. In: Birkhead TR, Møller AP, eds. Sperm competition and sexual selection. London, UK: Academic Press. pp 3-54.

Parker GA, Partridge L (1998) Sexual conflict and speciation. Phil Trans R Soc Lond Ser B 353: 261–74.

Parker GA (2006) Sexual conflict over mating and fertilization: an overview. Philos Trans R Soc B Biol Sci 361: 235–59.

Parker GA, Immler S, Pitnick S, Birkhead TR (2010) Sperm competition games: sperm size (mass) and number under raffle and displacement, and the evolution of P2. J Theor Biol 264: 1003-23.

Parker MA (1994) Pathogens and sex in plants. Evol Ecol 8: 560–84.

Parkhill J, Achtman M, James KD, Bentley SD, Churcher C, et al. (2000) Complete DNA sequence of a serogroup A strain of Neisseria meningitides Z2491. Nature 404: 502-6.

Parmakelis A, Kotsakiozi P, Rand D (2013) Animal mitochondria, positive selection and cyto-nuclear coevolution: insights from pulmonates. PLoS One 8: e61970.

Parmley JL, Urrutia AO, Potrzebowski L, Kaessmann H, Hurst LD (2007) Splicing and the evolution of proteins in mammals. PLoS Biol 5: e14.

Parsons KJ, Robinson BW (2006) Replicated evolution of integrated plastic responses during early adaptive divergence. Evolution 60: 801–13.

Parsons PA (1983) The Evolutionary Biology of Colonizing Species. New York, NY: Cambridge University Press.

Parsons PA (1987) Evolutionary rates under environmental stress. Evol Biol 21: 311–47.

Parsons PA (1988) Evolutionary rates: effects of stress upon recombination. Biol J Linn Soc 35: 49–68.

Parsons PA (1993) Developmental variability and the limits of adaptation: interactions with stress. Genetica 89: 245–53.

Parsons PA (1994) The energetic cost of stress. Can biodiversity be preserved? Biodivers Lett 2: 11-5.

Parsons PA (1997) Stress-resistance genotypes, metabolic efficiency and interpreting evolutionary change. EXS 83: 291-305.

Parsons PA (2005) Environments and evolution: interactions between stress, resource inadequacy and energetic efficiency. Biol Rev Camb Philos Soc 80: 589-610.

Parsons PA (2007) Energetic efficiency under stress underlies positive genetic correlations between longevity and other fitness traits in natural populations. Biogerontology 8: 55-61.

Parsons TJ, Muniec DS, Sullivan K, Woodyatt N, Alliston-Greiner R, et al. (1997) A high observed substitution rate in the human mitochondrial DNA control region. Nat Genet 15: 363–8.

Pärtel M, Hiiesalu I, Öpik M, Wilson SD (2012) Below-ground plant species richness: new insights from DNA-based methods. Funct Ecol 26: 775–82.

Parter M, Kashtan N, Alon U (2008) Facilitated variation: how evolution learns from past environments to generalize to new environments. PLoS Comput Biol 4: e1000206.

Partridge L, Barton NH (2000) Evolving evolvability. Nature 407: 457-8.

Partridge L, Gems D, Withers DJ (2005) Sex and death: what is the connection? Cell 120: 461–72.

Parvanov ED, Petkov PM, Paigen K (2010) Prdm9 controls activation of mammalian recombination hotspots. Science 327: 835.

Parvinen M, Söder O, Mali P, Froysa B, Ritzén EM (1991) In vitro stimulation of stage-specific deoxyribonucleic acid synthesis in rat seminiferous tubule segments by interleukin-1 alpha. Endocrinology 129: 1614-20.

Parzer HF, Moczek AP (2008) Rapid antagonistic coevolution between primary and secondary sexual characters in horned beetles. Evolution 62: 2423-8.

Pasi CE, Dereli-Öz A, Negrini S, Friedli M, Fragola G, et al. (2011) Genomic instability in induced stem cells. Cell Death Differ 18: 745-53.

Pasqualini C, Bojda F, Kerdelhue B (1986) Direct effect of estradiol on the number of dopamine receptors in the anterior pituitary of ovariectomized rats. Endocrinology 119: 2484-9.

Pasquinelli AE, Ruvkun G (2002) Control of developmental timing by microRNAs and their targets. Annu Rev Cell Dev Biol 18: 495–513.

Passos JF, Saretzki G, Ahmed S, Nelson G, Richter T, et al. (2007) Mitochondrial dysfunction accounts for the stochastic heterogeneity in telomere-dependent senescence. PLoS Biol 5: e110.

Pastwa E, Blasiak J (2003) Nonhomologous DNA end joining. Acta Biochim Pol 50: 891-908.

Patade V, Suprasanna P (2008) Radiation induced in vitro mutagenesis for sugarcane improvement. Sugar Tech 10: 14-19.

Patel MS, Srinivasan M, Laychock SG (2009) Metabolic programming: Role of nutrition in the immediate postnatal life. J Inherit Metab Dis 32: 218-28.

Patel MS, Srinivasan M (2011) Metabolic programming in the immediate postnatal life. Ann Nutr Metab 58 Suppl 2: 18-28.

Paterson S, Wilson K, Pemberton JM (1998) Major histocompatibility complex (MHC) variation associated with juvenile survival and parasite resistance in a large ungulate population (Ovis aries L.). Proc Natl Acad Sci USA 95: 3714–9.

Paterson S, Vogwill T, Buckling A, Benmayor R, Spiers AJ, et al. (2010) Antagonistic coevolution accelerates molecular evolution. Nature 464:275-8.

Patil VS, Kai T (2010) Repression of retroelements in Drosophila germline via piRNA pathway by the Tudor domain protein Tejas. Curr Biol 20: 724–30.

Patten BC, Odum EP (1981) The cybernetic nature of ecosystems. Am Nat 118: 886–95.

Patterson C, Williams DM, Humphries CJ (1993) Congruence between molecular and morphological phylogenies. Annu Rev Ecol Syst 24: 153–88.

Patwa Z, Wahl LM (2008) The fixation probability of beneficial mutations. J R Soc Interface 5: 1279–89

Paun O, Fay MF, Soltis DE, Chase MW (2007) Genetic and epigenetic alterations after hybridization and genome doubling. Taxon 56: 649–56.

Paul C, Murray AA, Spears N, Saunders PT (2008a) A single, mild, transient scrotal heat stress causes DNA damage, subfertility and impairs formation of blastocysts in mice. Reproduction 136: 73–84.

Paul C, Melton DW, Saunders PT (2008b) Do heat stress and deficits in DNA repair pathways have a negative impact on male fertility? Mol Hum Reprod 14:1-8.

Paul C, Teng S, Saunders PT (2009) A single, mild, transient scrotal heat stress causes hypoxia and oxidative stress in mouse testes, which induces germ cell death. Biol Reprod 80: 913–9.

Pauli S, von Velsen N, Burfeind P, Steckel M, Mänz J, et al. (2012) CHD7 mutations causing CHARGE syndrome are predominantly of paternal origin. Clin Genet 81: 234-9.

Paull TT, Rogakou EP, Yamazaki V, Kirchgessner CU, Gellert M, Bonner WM (2000) A critical role for histone H2AX in recruitment of repair factors to nuclear foci after DNA damage. Curr Biol 10: 886-95.

Pawlowska TE, Taylor JW (2004) Organization of genetic variation in individuals of arbuscular mycorrhizal fungi. Nature 427: 733–7.

Payabvash S, Kiumehr S, Tavangar SM, Dehpour AR (2008) Ethyl pyruvate reduces germ cell-specific apoptosis and oxidative stress in rat model of testicular torsion/detorsion. J Pediatr Surg 43: 705-12.

Payseur BA, Jing P, Haasl RJ (2011) A genomic portrait of human microsatellite variation. Mol Biol Evol 28: 303–12.

Pearl LH, Prodromou C (2006) Structure and mechanism of the Hsp90 molecular chaperone machinery. Annu Rev Biochem 75: 271–94.

Pearl LH, Prodromou C, Workman P(2008) The Hsp90 molecular chaperone: an open and shut case for treatment. Biochem J 410: 439–53.

Pearlstein DP, Ali MH, Mungai PT, Hynes KL, Gewertz BL, Schumacker PT (2002) Role of mitochondrial oxidant generation in endothelial cell responses to hypoxia. Arterioscler Thromb Vasc Biol 22: 566–73.

Pearson CE, Edamura KN, Cleary JD (2005) Repeat instability: mechanisms of dynamic mutations. Nat Rev Genet 6: 729-42.

Peaston AE, Evsikov AV, Graber JH, de Vries WN, Holbrook AE, et al. (2004) Retrotransposons regulate host genes in mouse oocytes and preimplantation embryos. Dev Cell 7: 597–606.

Peaston AE, Whitelaw E (2006) Epigenetics and phenotypic variation in mammals. Mamm Genome 17: 365-74.

Pecci L, Montefoschi G, Cavallini D (1997) Some new details of the copper-hydrogen peroxide interaction. Biochem Biophys Res Commun 235: 264-7.

Peck JR (1993) Frequency-dependent selection, beneficial mutations, and the evolution of sex. Proc R Soc Lond B Biol Sci 254: 87-92.

Peck JR (1994) A ruby in the rubbish: beneficial mutations, deleterious mutations and the evolution of sex. Genetics 137: 597-606.

Peck JR, Barreau G, Heath SC (1997) Imperfect genes, Fisherian mutation and the evolution of sex. Genetics 145: 1171-99.

Peck JR, Yearsley JM, Waxman D (1998) Explaining the geographic distributions of sexual and asexual populations. Nature 391: 889–92.

Peck JR, Yearsley J, Barreau G (1999) The maintenance of sexual reproduction in a structured population. Proc R Soc Lond B 266: 1857–63.

Peck JR, Waxman D (2000) Mutation and sex in a competitive world. Nature 406: 399-404.

Pecon Slattery J, O’Brien SJ (1998) Patterns of Y and X chromosome DNA sequence divergence during the Felidae radiation. Genetics 148: 1245–55.

Pecoraro V, Zerulla K, Lange C, Soppa J (2011) Quantification of ploidy in proteobacteria revealed the existence of monoploid, (mero-)oligoploid and polyploid species. PLoS ONE 6: e16392.

Pecqueur C, Alves-Guerra MC, Gelly C, Levi-Meyrueis C, Couplan E, et al. (2001) Uncoupling protein 2, in vivo distribution, induction upon oxidative stress, and evidence for translational regulation. J Biol Chem 276: 8705-12.

Pedersen LD, Pedersen AR, Bijlsma R, Bundgaard J (2011) The effects of inbreeding and heat stress on male sterility inDrosophila melanogaster. Biol J Linn Soc 104: 432–42.

Pedraza-Reyes M, Yasbin RE (2004) Contribution of the mismatch DNA repair system to the generation of stationary-phase-induced mutants of Bacillus subtilis. J Bacteriol 186: 6485–91.

Pek JW, Anand A, Kai T (2012a) Tudor domain proteins in development. Development 139: 2255–66.

Pek JW, Patil VS, Kai T (2012b) piRNA pathway and the potential processing site, the nuage, in the Drosophila germline. Dev Growth Differ 54: 66-77.

Peled-Kamar M, Lotem J, Okon E, Sachs L, Groner Y (1995) Thymic abnormalities and enhanced apoptosis of thymocytes and bone marrow cells in transgenic mice overexpressing Cu/Zn-superoxide dismutase: Implications for Down Syndrome. EMBO J 14: 4985–93.

Peled-Kamar M, Lotem J, Wirguin I, Weiner L, Hermalin A, Groner Y (1997) Oxidative stress mediates impairment of muscle formation in transgenic mice with elevated level of wild-type Cu/Zn superoxide dismutase. Proc Natl Acad Sci USA 94: 3883–7.

Pelengaris S, Littlewood T, Khan M, Elia G, Evan G (1999) Reversible activation of c-Myc in skin: induction of a complex neoplastic phenotype by a single oncogenic lesion. Mol Cell 3: 565–77.

Pelengaris S, Khan M, Evan G (2002)c-MYC: more than just a matter of life and death. Nat Rev Cancer 2: 764-76.

Pélisson A, Payen-Groschêne G, Terzian C, Bucheton A (2007) Restrictive flamenco alleles are maintained in Drosophila melanogaster population cages, despite the absence of their endogenous gypsy retroviral targets. Mol Biol Evol 24: 498–504.

Pelletier F, Garant D, Hendry AP (2009) Eco-evolutionary dynamics. Philos Trans R Soc B Biol Sci 364: 1483-9.

Peltola V, Huhtaniemi I, Ahotupa M (1992) Antioxidant enzyme activity in the maturing rat testis. J Androl 13: 450-5.

Peltola V, Huhtaniemi I, Ahotupa M (1995) Abdominal position of the rat testis is associated with high level of lipid peroxidation. Biol Reprod 53: 1146–50.

Peltola V, Huhtaniemi I, Metsa-Ketela T, Ahotupa M (1996) Induction of lipid peroxidation during steroidogenesis in the rat testis. Endocrinology 137: 105–12.

Pembrey M (1996) Imprinting and transgenerational modulation of gene expression; human growth as a model. Acta Genet Med Gemellol (Roma) 45: 111–25.

Pembrey ME, Bygren LO, Kaati G, Edvinsson S, Northstone K, Sjostrom M, Golding J (2006) Sex-specific, male-line transgenerational responses in humans. Eur J Hum Genet 14: 159–66.

Peng J, Zhang L, Drysdale L, Fong GH (2000) The transcription factor EPAS-1/hypoxia-inducible factor 2alpha plays an important role in vascular remodeling. Proc Natl Acad Sci USA 97: 8386–91.

Peng TI, Jou MJ (2010) Oxidative stress caused by mitochondrial calcium overload. Ann NY Acad Sci 1201: 183-8.

Penn D, Potts W (1999) The evolution of mating preferences and major histocompatibility genes. Am Nat 153: 145-64.

Pennisi E (2008) Modernizing the modern synthesis. Science 321: 196–7.

Penrose LS (1934) The relative aetiological importance of birth order and maternal age in mongolism. Proc R Soc Lond B 115: 431–50.

Penrose LS (1955) Parental age and mutation. Lancet II: 312–3.

Pentikäinen V, Erkkilä K, Dunkel L (1999) Fas regulates germ cell apoptosis in the human testis in vitro. Am J Physiol 276: E310–E316.

Pentikäinen V, Erkkilä K, Suomalainen L, Parvinen M, Dunkel L (2000) Estradiol acts as a germ cell survival factor in the human testis in vitro. J Clin Endocrinol Metab 85: 2057–67.

Pentikäinen V, Erkkilä K, Suomalainen L, Otala M, Pentikäinen MO, et al. (2001) TNFalpha down-regulates the Fas ligand and inhibits germ cell apoptosis in the human testis. J Clin Endocrinol Metab 86: 4480-8.

Pentikäinen V, Suomalainen L, Erkkilä K, Martelin E, Parvinen M, et al. (2002) Nuclear factor-kappaB activation in human testicular apoptosis. Am J Pathol 160: 205-18.

Pepling ME, Spradling AC (2001) Mouse ovarian germ cell cysts undergo programmed breakdown to form primordial follicles. Dev Biol 234: 339-51.

Pepling ME, Wilhelm JE, O'Hara AL, Gephardt GW, Spradling AC (2007) Mouse oocytes within germ cell cysts and primordial follicles contain a Balbiani body. Proc Natl Acad Sci USA 104: 187-92.

Perego M (1998) Kinase-phosphatase competition regulates Bacillus subtilis development. Trends Microbiol 6: 366-70.

Pereira B, Rosa L, Safi DA, Bechara EJH, Curi R (1995) Hormonal regulation of superoxide dismutase, catalase, and glutathione peroxidase activities in rat macrophases. Biochem Pharmacol 50: 2093-8.

Pereira L, Soares P, Radivojac P, Li B, Samuels DC (2011) Comparing phylogeny and the predicted pathogenicity of protein variations reveals equal purifying selection across the global human mtDNA diversity. Am J Hum Genet 88: 433–9.

Pereira TM, Carlstedt-Duke J, Lechner MC, Gustafsson JA (1998) Identification of a functional glucocorticoid response element in the CYP3A1/IGC2 gene. DNA Cell Biol 17: 39-49.

Perera D, Pizzey A, Campbell A, Katz M, Porter J, et al. (2002) Sperm DNA damage in potentially fertile homozygous beta-thalassaemia patients with iron overload. Hum Reprod 17: 1820-5.

Perez GI, Robles R, Knudson CM, Flaws JA, Korsmeyer SJ, Tilly JL (1999) Prolongation of ovarian lifespan into advanced chronological age by Bax-deficiency. Nat Genet 21: 200–3.

Perez GI, Trbovich AM, Gosden RG, Tilly JL (2000) Mitochondria and the death of oocytes. Nature 403: 500–1.

Perez GI, Acton BM, Jurisicova A, Perkins GA, White A, et al. (2007) Genetic variance modifies apoptosis susceptibility in mature oocytes via alterations in DNA repair capacity and mitochondrial ultrastructure. Cell Death Differ 14: 524–33.

Perez-Campo R, Lopez-Torres M, Cadenas S, Rojas C, Barja G (1998) The rate of free radical production as a determinant of the rate of aging: evidence from the comparative approach. J Comp Physiol [B] 168: 149–58.

Pérez-Crespo M, Pintado B, Gutiérrez-Adán A (2008) Scrotal heat stress effects on sperm viability, sperm DNA integrity, and the offspring sex ratio in mice. Mol Reprod Dev 75: 40–7.

Perez-Hormaeche J, Potet F, Beauclair L, Le Masson I, Courtial B, et al. (2008) Invasion of the Arabidopsis genome by the tobacco retrotransposon Tnt1 is controlled by reversible transcriptional gene silencing. Plant Physiol 147: 1264-78.

Pérez-Martín J, Uría JA, Johnson AD (1999) Phenotypic switching in Candida albicans is controlled by a SIR2 gene. EMBO J 18: 2580-92.

Perfeito L, Fernandes L, Mota C, Gordo I (2007) Adaptive mutations in bacteria: high rate and small effects. Science 317: 813–5.

Perler F, Efstratiadis A, Lomedico P, Gilbert W, Kolodner R, Dodgson J (1980) The evolution of genes: the chicken preproinsulin gene. Cell 20: 555-66.

Perluigi M, Butterfield DA (2011) The identification of protein biomarkers for oxidative stress in Down syndrome. Expert Rev Proteomics 8: 427–9.

Perrimon N, Engstrom L, Mahowald AP (1984) The effects of zygotic lethal mutations on germline functions in Drosophila. Dev Biol 105: 404-14.

Perrins CM (1965) Population fluctuations and clutch size in the Great Tit, Parus major L. J Anim Ecol 34: 601-47.

Perrins CM, Moss D (1975) Reproductive rates in the Great Tit. J Anim Ecol 44: 695–706.

Perry AN, Grober MS (2003) A model for social control of sex change: interactions of behavior, neuropeptides, glucocorticoids, and sex steroids. Horm Behav 43: 31–8.

Pesah Y, Pham T, Burgess H, Middlebrooks B, Verstreken P, et al. (2004) Drosophila parkin mutants have decreased mass and cell size and increased sensitivity to oxygen radical stress. Development 131: 2183–94.

Pesaresi P, Schneider A, Kleine T, Leister D (2007) Interorganellar communication. Curr Opin Plant Biol 10: 600–6.

Peschel K, Norton RA, Scheu S, Maraun M (2006) Do oribatid mites live in enemy-free space? Evidence from feeding experiments with the predatory mite Pergamasus septentrionalis. Soil Biol Biochem 38: 2985–9.

Pesgraves DC, Baker RH, Wilkinson GS (1999) Coevolution of sperm and female reproductive tract morphology in stalk-eyed flies. Proc R Soc Lond B 266: 1041–7.

Pesheva M, Krastanova O, Staleva L, Dentcheva V, Hadzhitodorov M, Venkov P (2005) The Ty1 transposition assay: a new short-term test for detection of carcinogens. J Microb Meth 61: 1–8.

Pesheva M, Krastanova O, Stamenova R, Kantardjiev D, Venkov P (2008) The response of Ty1 test to genotoxins. Arch Toxicol 82: 779–85.

Peskin AV, Low FM, Paton LN, Maghzal GJ, Hampton MB, Winterbourn CC (2007) The high reactivity of peroxiredoxin 2 with H2O2 is not reflected in its reaction with other oxidants and thiol reagents. J Biol Chem 282: 11885–92.

Peters AD, Lively CM (1999) The Red Queen and fluctuating epistasis: a population genetic analysis of antagonistic coevolution. Am Nat 54: 393–405.

Peters AD, Keightley PD (2000) A test for epistasis among induced mutations in Caenorhabditis elegans. Genetics 156: 1635–47.

Peters AD, Lively CM (2007) Short- and long-term benefit and detriments to recombination under antagonistic coevolution. J Evol Biol 20: 1206–17.

Peters H, Himelstein-Braw R, Faber M (1976) The normal development of the ovary in childhood. Acta Endocrinol (Copenh) 82: 617–30.

Peters RH (1983) The Ecological Implications of Body Size. Cambridge, UK: Cambridge University Press.

Petersen C, Boitani C, Fröysa B, Söder O (2002) Interleukin-1 is a potent growth factor for immature rat Sertoli cells. Mol Cell Endocrinol 186: 37-47.

Petersen C, Fröysa B, Söder O (2004) Endotoxin and proinflammatory cytokines modulate Sertoli cell proliferation in vitro. J Reprod Immunol 61: 13-30.

Petersen C, Söder O (2006) The sertoli cell--a hormonal target and ‘super’ nurse for germ cells that determines testicular size. Horm Res 66: 153-61.

Peterson FC, Baden EM, Owen BA, Volkman BF, Ramirez-Alvarado M (2010) A single mutation promotes amyloidogenicity through a highly promiscuous dimer interface. Structure 18: 563-70.

Peterson KJ, McPeek MA, Evans DAD (2005) Tempo and mode of early animal evolution: inferences from rocks, Hox and molecular clocks. Paleobiology 31 (2, supplement): 36–55.

Peterson KJ, Cotton JA, Gehling JG, Pisani D (2008) The Ediacaran emergence of bilaterians: congruence between the genetic and the geological fossil records. Phil Trans R Soc B 363: 1435–43.

Petes TD (2001) Meiotic recombination hot spots and cold spots. Nat Rev Genet 2: 360–9.

Petit C, Sancar A (1999) Nucleotide excision repair: from E. coli to man. Biochimie 81: 15-25.

Petit MA, Bedale W, Osipiuk J, Lu C, Rajagopalan M, et al. (1994) Sequential folding of UmuC by the Hsp70 and Hsp60 chaperone complexes of Escherichia coli. J Biol Chem 269: 23824-9.

Petitte JN, Etches RJ (1988) The effect of corticosterone on the photoperiodic response of immature hens. Gen Comp Endocrinol 69: 424-30.

Petitte JN, Etches RJ (1991) Daily infusion of corticosterone and reproductive function in the domestic hen (Gailus domesticus). Gen Comp Endocrinol 83: 397-405.

Petraglia F, Sutton S, Vale W, Plotsky P (1987) Corticotropin-releasing factor decreases plasma luteinizing hormone levels in female rats by inhibiting gonadotropin-releasing hormone release into hypophysial-portal circulation. Endocrinology 120: 1083-8.

Petrie M (1994) Improved growth and survival of offspring of peacocks with more elaborate trains. Nature 371: 598-9.

Petrie M, Williams A (1993) Peahens lay more eggs for peacocks with larger trains. Proc R Soc Lond B 251: 127–31.

Petrie M, Doums C, Møller AP (1998) The degree of extra-pair paternity increases with genetic variance. Proc Natl Acad Sci 95: 9390–5.

Petrie M, Kempenaers B (1998) Extra-pair paternity in birds: explaining variation between species and populations. Trends Ecol Evol 13: 52–8.

Petrie M, Roberts G (2007) Sexual selection and the evolution of evolvability. Heredity (Edinb) 98: 198-205.

Petronis A (2010) Epigenetics as a unifying principle in the aetiology of complex traits and diseases. Nature 465: 721-7.

Petronis A, Gottesman II, Kan P, Kennedy JL, Basile VS, et al. (2003) Monozygotic twins exhibit numerous epigenetic differences: clues to twin discordance? Schizophr Bull 29: 169-78.

Petrossian TC, Clarke SG (2009) Multiple motif scanning to identify methyltransferases from the yeast proteome. Mol Cell Proteomics 8: 1516–26.

Petrov DA, Hartl DL (1999) Patterns of nucleotide substitution in Drosophila and mammalian genomes. Proc Natl Acad Sci USA 96: 1475–9.

Petrova E, Soldini D, Moreno E (2011) The expression of SPARC in human tumors is consistent with its role during cell competition. Commun Integr Biol 4: 171-4.

Pettit NE, Froend RH (2001) Availability of seed for recruitment of riparian vegetation: a comparison of a tropical and a temperate river ecosystem in Australia. Aust J Bot 49: 515–28.

Pevzner P, Tesler G (2003a) Human and mouse genomic sequences reveal extensive breakpoint reuse in mammalian evolution. Proc Natl Acad Sci USA 100: 7672–7.

Pevzner P, Tesler G (2003b) Genome rearrangements in mammalian evolution: lessons from human and mouse genomes. Genome Res 13: 37–45.

Pfeffer S, Zavolan M, Grässer FA, Chien M, Russo JJ, et al. (2004) Identification of virus-encoded microRNAs. Science 304: 734-6.

Pfeifer GP (2000) p53 mutational spectra and the role of methylated CpG sequences. Mutat Res 450: 155–66.

Pfeifer GP (2006) Mutagenesis at methylated CpG sequences. Curr Topics Microbiol Immunol 301: 259–281.

Pfeiffer JK, Kirkegaard K (2006) Bottleneck-mediated quasispecies restriction during spread of an RNA virus from inoculation site to brain. Proc Natl Acad Sci USA 103: 5520–5.

Pfeilschifter J, Eberhardt W, Beck KF (2001) Regulation of gene expression by nitric oxide. Pflügers Arch 442: 479–86.

Pfister CA (1998) Patterns of variance in stage-structured populations: evolutionary predictions and ecological implications. Proc Natl Acad Sci USA 95: 213–8.

Pfitzer P, Gilbert P, Rolz G, Vyska K (1982) Flow cytometry of human testicular tissue. Cytometry 3: 116–22.

Pfrender ME, Lynch M (2000) Quantitative genetic variation in Daphnia: temporal changes in genetic architecture. Evolution 54: 1502-9.

Pham P, Rangarajan S, Woodgate R, Goodman MF (2001) Roles of DNA polymerases V and II in SOS-induced error-prone and error-free repair in Escherichia coli. Proc Natl Acad Sci USA 98: 8350-4.

Phifer-Rixey M, Bonhomme F, Boursot P, Churchill GA, Piálek J, et al. (2012) Adaptive evolution and effective population size in wild house mice. Mol Biol Evol 29: 2949-55.

Philippe N, Pelosi L, Lenski RE,Schneider D (2009) Evolution of penicillin-binding protein 2 concentration and cell shape in a long term experiment with Escherichia coli. J Bacteriol 191: 909–21.

Philippi T, Seger J (1989) Hedging one's evolutionary bets, revisited. Trends Evol Ecol 4: 41-4.

Phillips JP, Campbell SD, Michaud D, Charbonneau M, Hilliker AJ (1989) Null mutation of copper/zinc superoxide dismutase in Drosophila confers hypersensitivity to paraquat and reduced longevity. Proc Natl Acad Sci USA 86: 2761-5.

Phillips JR, Dalmay T, Bartels D (2007) The role of small RNAs in abiotic stress. FEBS Lett 581: 3592-7.

Phillips PC (2008) Epistasis–the essential role of gene interactions in the structure and evolution of genetic systems. Nat Rev Genet 9: 855–67.

Phillips PC, Arnold SJ (1989) Visualizing multivariate selection. Evolution 43: 1209–22.

Phoa N, Epe B (2002) Influence of nitric oxide on the generation and repair of oxidative DNA damage in mammalian cells. Carcinogenesis 23: 469-75.

Phoenix C (2009) Cellular differentiation as a candidate “new technology” for the Cambrian Explosion. J Evol Technol 20: 43–8.

Pianka ER (1966) Latitudinal gradients in species diversity: a review of concepts. Am Nat 100: 33–46.

Pianka ER (1970) On r- and K-selection. Am Nat 104: 592–7.

Pianka ER (1974a) Evolutionary ecology. New York, NY: Harper and Row.

Pianka ER (1974b) Niche overlap and diffuse competition. Proc Natl Acad Sci USA 71: 2141-5.

Pickering AD, Pottinger TG, Carragher JF, Sumpter JP (1987) The effects of acute and chronic stress on the levels of reproductive hormones in the plasma of mature male brown trout, Salmo trutta L. Gen Comp Endocrinol 68: 249-59.

Pieau C, Dorizzi M, Richard-Mercier N (1999) Temperature-dependent sex determination and gonadal differentiation in reptiles. Cell Mol Life Sci 55: 887–900.

Pierre JL, Fontecave M (1999) Iron and activated oxygen species in biology: the basic chemistry. Biometals 12: 195-9.

Piganeau G, Gardner M, Eyre-Walker A (2004) A broad survey of recombination in animal mitochondria. Mol Biol Evol 21: 2319–25.

Piganeau G, Eyre-Walker (2009) Evidence for variation in the effective population size of animal mitochondrial DNA. PLoS One 4: e4396.

Pigliucci M (2001) Phenotypic plasticity: beyond nature and nurture. Baltimore, MD: Johns Hopkins University Press.

Pigliucci M (2003) Epigenetics is back! Hsp90 and phenotypic variation. Cell Cycle 2: 34-5.

Pigliucci M (2007) Do we need an extended evolutionary synthesis? Evolution 61: 2734–49.

Pigliucci M (2008) Is evolvability evolvable? Nat Rev Genet 9: 75–82.

Pigliucci M (2009) An extended synthesis for evolutionary biology. Ann NY Acad Sci 1168: 218–28.

Pigliucci M, Kaplan J (2000) The fall and rise of Dr Pangloss: Adaptationism and the spandrels paper 20 years later. Trends Ecol Evol 15: 66–70.

Pigliucci M, Murren CJ (2003) Genetic assimilation and a possible evolutionary paradox: can macroevolution sometimes be so fast as to pass us by? Evolution 57: 1455–64.

Pigliucci M, Murren CJ, Schlichting CD (2006) Phenotypic plasticity and evolution by genetic assimilation. J Exp Biol 209: 2362–7.

Pigozzi MI, Solari AJ (2005) The germ-line-restricted chromosome in the zebra finch: recombination in females and elimination in males. Chromosoma 114: 403–9.

Pihl L (1985) Food selection and consumption of mobile epibenthic fauna in shallow marine areas. Mar Ecol-Prog Ser 22: 169–79.

Pike TW, Blount JD, Bjerkeng B, Lindström J, Metcalfe NB (2007) Carotenoids, oxidative stress and female mating preference for longer lived males. Proc R Soc B 274: 1591–6.

Pillai RS, Chuma S (2012) piRNAs and their involvement in male germline development in mice. Dev Growth Differ54: 78-92.

Pineau C, Le Magueresse B, Courtens JL, Jégou B (1991) Study in vitro of the phagocytic function of Sertoli cells in the rat. Cell Tissue Res 264: 589–98.

Pineda-Krch F, Fagerström T (1999) On the potential for evolutionary change in meristematic cell lineages through intraorganismal selection. J Evol Biol 12: 681–8.

Pineda-Krch M, Lehtilä K (2002) Cell lineage dynamics in stratified shoot apical meristems. J Theor Biol 219: 495–505.

Pineda-Krch M, Lehtilä K (2004) Costs and benefits of genetic heterogeneity within organisms. J Evol Biol 17: 1167–77.

Pines A, Bivi N, Romanello M, Damante G, Kelley MR, et al. (2005) Cross-regulation between Egr-1 and APE/Ref-1 during early response to oxidative stress in the human osteoblastic HOBIT cell line: Evidence for an autoregulatory loop. Free Radic Res 39: 269–81.

Pinheiro AR, Salvucci ID, Aguila MB, Mandarim-de-Lacerda CA (2008) Protein restriction during gestation and/or lactation causes adverse transgenerational effects on biometry and glucose metabolism in F1 and F2 progenies of rats. Clin Sci (Lond) 114: 381-92.

Pink CJ, Swaminathan SK, Dunham I, Rogers J, Ward A, Hurst LD (2009) Evidence that replication-associated mutation alone does not explain between-chromosome differences in substitution rates. Genome Biol Evol 1:13–22.

Pink CJ, Hurst LD (2010) Timing of replication is a determinant of neutral substitution rates but does not explain slow Y chromosome evolution in rodents. Mol Biol Evol 27: 1077–86.

Pinkerton JH, McKay DG, Adams EC, Hertig AT (1961) Development of the human ovary--a study using histochemical technics. Obstet Gynecol 18: 152-81.

Pinto RL, Rocha CEF, Martens K (2007) Early release of eggs and embryos in a brooding ancient asexual ostracod: brood selection or bet-hedging to increase fecundity? Hydrobiologia 585: 249–53.

Piret JP, Mottet D, Raes M, Michiels C (2002) Is HIF-1α a pro- or an antiapoptotic protein? Biochem Pharmacol 64: 889–892.

Piriyapongsa J, Jordan IK (2007) A family of human microRNA genes from miniature inverted-repeat transposable elements. PLoS ONE 2: e203.

Piriyapongsa J, Rutledge MT, Patel S, Borodovsky M, Jordan IK (2007a) Evaluating the protein coding potential of exonized transposable element sequences. Biol Direct 2: 31.

Piriyapongsa J, Marino-Ramirez L, Jordan IK (2007b) Origin and evolution of human microRNAs from transposable elements. Genetics 176: 1323–37.

Piriyapongsa J, Jordan IK (2008) Dual coding of siRNAs and miRNAs by plant transposable elements. RNA 14: 814-21.

Pirkkala L, Nykänen P, Sistonen L (2001) Roles of the heat shock transcription factors in regulation of the heat shock response and beyond. FASEB J 15: 1118-31.

Pischedda A, Chippindale AK (2006) Intralocus sexual conflict diminishes the benefits of sexual selection. PLoS Biol 4: e356.

Pischedda A, Stewart AD, Little MK, Rice WR (2011) Male genotype influences female reproductive investment in Drosophila melanogaster. Proc Biol Sci 278: 2165-72.

Pitnick S, Hosken DJ, Birkhead TR (2009) Sperm morphological diversity. In: Birkhead TR, Hosken DJ, Pitnick S, eds. Sperm Biology, an Evolutionary Perspective. San Diego, CA: Academic Press. pp 69–149.

Pitt JN, Schisa JA, Priess JR (2000) P granules in the germ cells of Caenorhabditis elegans adults are associated with clusters of nuclear pores and contain RNA. Dev Biol 219: 315–33.

Piva R, Belardo G, Santoro MG (2006) NF-kappaB: a stress-regulated switch for cell survival. Antioxid Redox Signal 8: 478-86.

Pizzari T, Worley K, Burke T, Froman DP (2008) Sperm competition dynamics: ejaculate fertilising efficiency changes differentially with time. BMC Evol Biol 8: 332–8.

Pizzari T, Foster KR (2008) Sperm sociality: Cooperation, altruism, and spite. PLoS Biol 6: e130.

Planchet E, Jagadis GK, Sonoda M, Kaiser WM (2005) Nitric oxide emission from tobacco leaves and cell suspensions: rate limiting factors and evidence for the involvement of mitochondrial electron transport. Plant J 41: 732–43.

Plasterk RH (2002) RNA silencing: The genome’s immune system. Science 296: 1263–5.

Plath M (2008) Male mating behavior and costs of sexual harassment for females in cavernicolous and extremophile populations of Atlantic mollies (Poecilia mexicana). Behaviour 145:73–98.

Plendl J (2000) Angiogenesis and vascular regression in the ovary. Anat Histol Embryol 29: 257-66.

Plesset J, Ludwig J, Cox B, McLaughlin C (1987) Effect of cell cycle position on thermotolerance in Saccharomyces cerevisiae. J Bacteriol 169: 779-84.

Plotsky PM, Vale W (1984) Hemorrhage-induced secretion of corticotropin-releasing factor like immunoreactivity into the rat hypophyseal portal circulation and its inhibition by glucocorticoid. Endocrinology 114: 164-9.

Plough HH (1917) The effect of temperature on crossing over in Drosophila. J Exp Biol 24: 147–209.

Plough HH (1921) Further studies on the effect of temperature on crossing over. J Exp Biol 32: 187–202.

Plyusnin A, Cheng Y, Lehvaslaiho H, Vaheri A (1996) Quasispecies in wild-type Tula hantavirus populations. J Virol 70: 9060–3.

Pocock R (2011) Invited review: decoding the microRNA response to hypoxia. Pflugers Arch 461: 307-15.

Podrabsky JE, Clelen D, Crawshaw LI (2008) Temperature preference and reproductive fitness of the annual killifish Austrofundulus limnaeus exposed to constant and fluctuating temperatures. J Comp Physiol A 194: 385–93.

Poelwijk FJ, Kiviet DJ, Weinreich DM, Tans SJ (2007) Empirical fitness landscapes reveal accessible evolutionary paths. Nature 445: 383–6.

Poelwijk FJ, Tanase-Nicola S, Kiviet DJ, Tans SJ (2011) Reciprocal sign epistasis is a necessary condition for multi-peaked fitness landscapes. J Theor Biol 272: 141–4.

Poinar GO, Ricci C (1982) Bdelloid rotifers in Dominican amber: evidence for parthenogenetic continuity. Experientia 48: 408-10.

Poinar G Jr, Poinar R (1998) Parasites and pathogens of mites. Annu Rev Entomol 43: 449–69.

Polis GA (1981) The evolution and dynamics of intraspecific predation. Annu Rev Ecol Syst 12: 225-51.

Polo SE, Jackson SP (2011) Dynamics of DNA damage response proteins at DNA breaks: a focus on protein modifications. Genes Dev 25: 409-33.

Pollanen P, Söder O, Parvinen M (1989) Interleukin-1 alpha stimulation of spermatogonial proliferation in vivo. Reprod Fertil Dev 1: 85-7.

Pollard KS, Salama SR, Lambert N, Lambot MA, Coppens S, et al. (2006a) An RNA gene expressed during cortical development evolved rapidly in humans. Nature 443: 167–172.

Pollard KS, Salama SR, King B, Kern AD, Dreszer T, et al. (2006b) Forces shaping the fastest evolving regions in the human genome. PLoS Genet 2: e168.

Pollard PJ, Loenarz C, Mole DR, McDonough MA, Gleadle JM, et al. (2008) Regulation of Jumonji-domain containing histone demethylases by hypoxia inducible factor (HIF) 1-alpha. Biochem J 416: 387–94.

Pollock DD, Thiltgen G, Goldstein RA (2012) Amino acid coevolution induces an evolutionary Stokes shift. Proc Natl Acad Sci USA 109: E1352-9.

Poltoratsky V, Woo CJ, Tippin B, Martin A, Goodman MF, Scharff MD (2001) Expression of error-prone polymerases in BL2 cells activated for Ig somatic hypermutation. Proc Natl Acad Sci USA 98: 7976-81.

Polyak K, Xia Y, Zweier JL, Kinzler KW, Vogelstein B (1997) A model for p53-induced apoptosis. Nature 389: 300–5.

Pomeraniec Y, Grion N, Gadda L, Pannunzio V, Podesta EJ, Cymeryng CB (2004) Adrenocorticotropin induces heme oxygenase-1 expression in adrenal cells. J Endocrinol 180: 113–24.

Pomerantz MM, Ahmadiyeh N, Jia L, Herman P, Verzi MP, et al. (2009) The 8q24 cancer risk variant rs6983267 shows long-range interaction with MYC in colorectal cancer. Nat Genet 41: 882–4.

Pomiankowski A, Møller AP (1995) A resolution of the lek paradox. Proc R Soc Lond B 260: 21-29.

Pomiankowski A, Bridle J (2004) No sex please we’re at QLE (quasi-linkage equilibrium). Heredity 93: 407.

Ponder RG, Fonville NC, Rosenberg SM (2005) A switch from high-fidelity to error-prone DNA double-strand break repair underlies stress-induced mutation. Mol Cell 19: 791-804.

Ponjavic J, Ponting CP, Lunter G (2007) Functionality or transcriptional noise? Evidence for selection within long noncoding RNAs. Genome Res 17: 556–65.

Pontecorvo G (1956) The parasexual cycle in fungi. Annu Rev Microbiol 10: 393–400.

Pontecorvo G, Käfer E (1958) Genetic analysis based on mitotic recombination. Adv Genet 9:71–104.

Ponting CP, Oliver PL, Reik W (2009) Evolution and functions of long noncoding RNAs. Cell 136: 629–41.

Poon A, Chao L (2004) Drift increases the advantage of sex in RNA bacteriophage Phi 6. Genetics 166:19–24.

Poon WL, Hung CY, Randall DJ (2001) The effect of aquatic hypoxia on fish. In: Thurston RV ed. Fish Physiology, Toxicology and Water Quality. Environmental Protection Agency, USA. pp 31–50.

Poorter H, Niklas KJ, Reich PB, Oleksyn J, Poot P, Mommer L (2012) Biomass allocation to leaves, stems and roots: meta-analyses of interspecific variation and environmental control. New Phytol 193: 30–50.

Popadin K, Polishchuk LV, Mamirova L, Knorre D, Gunbin K (2007) Accumulation of slightly deleterious mutations in mitochondrial protein-coding genes of large versus small mammals. Proc Natl Acad Sci USA 104: 13390-5.

Popowich DA, Vavra AK, Walsh CP, Bhikhapurwala HA, Rossi NB, et al. (2010) Regulation of reactive oxygen species by p53: implications for nitric oxide-mediated apoptosis.Am J Physiol Heart Circ Physiol 298: H2192-200.

Popp C, Dean W, Feng S, Cokus SJ, Andrews S, et al. (2010) Genome-wide erasure of DNA methylation in mouse primordial germ cells is affected by AID deficiency. Nature 463: 1101-5.

Pornthanakasem W, Kongruttanachok N, Phuangphairoj C, Suyarnsestakorn C, Sanghangthum T, et al. (2008) LINE-1 methylation status of endogenous DNA double-strand breaks. Nucleic Acids Res 36: 3667–75.

Portela M, Casas-Tinto S, Rhiner C, López-Gay JM, Domínguez O, et al. (2010) Drosophila SPARC is a self-protective signal expressed by loser cells during cell competition. Dev Cell 19: 562-73.

Porter ML, Crandall KA (2003) Lost along the way: the significance of evolution in reverse. Trends Ecol Evol 18: 541-7.

Porter RK, Hulbert AJ, Brand MD (1996) Allometry of mitochondrial proton leak: influence of membrane surface area and fatty acid composition. Am J Physiol 271: R1550-60.

Porter SM (2007) Seawater chemistry and early carbonate biomineralization. Science 316: 1302.

Pörtner HO (2006) Climate dependent evolution of Antarctic ectotherms: an integrative analysis. Deep Sea Res II 53: 1071–104.

Posnick LM, Samson LD (1999) Imbalanced base excision repair increases spontaneous mutation and alkylation sensitivity in Escherichia coli. J Bacteriol 181: 6763–71.

Pothof J, Verkaik NS, Hoeijmakers JH, van Gent DC (2009) MicroRNA responses and stress granule formation modulate the DNA damage response. Cell Cycle 8: 3462–8.

Potts RJ, Bespalov IA, Wallace SS, Melamede RJ, Hart BA (2001) Inhibition of oxidative DNA repair in cadmium-adapted alveolar epithelial cells and the potential involvement of metallothionein. Toxicology 161: 25–38.

Potts RJ, Watkin RD, Hart BA (2003) Cadmium exposure downregulates 8-oxoguanine DNA glycosylase expression in rat lung and alveolar epithelial cells. Toxicology 184: 189–202.

Potts WK, Manning CJ, Wakeland EK (1991) Mating patterns in seminatural populations of mice influenced by MHC genotype. Nature 352: 619-21.

Pouchkina-Stantcheva NN, McGee BM, Boschetti C, Tolleter D, Chakrabortee S, et al. (2007) Functional divergence of former alleles in an ancient asexual invertebrate. Science 318: 268–271.

Poulsen M, Erhardt DP, Molinaro DJ, Lin T-L, Currie CR (2007) Antagonistic bacterial interactions help shape host-symbiont dynamics within the fungus-growing ant-microbe mutualism. PLoS ONE 2: e960.

Poulton J, Chiaratti MR, Meirelles FV, Kennedy S, Wells D, et al. (2010) Transmission of mitochondrial DNA diseases and ways to prevent them. PLoS Genet 6: e1001066.

Pound GE, Doncaster CP, Cox SJ (2002) A Lotka–Volterra model of coexistence between a sexual population and multiple asexual clones. J Theor Biol 217: 535–45.

Pound GE, Cox SJ, Doncaster CP (2004) The accumulation of deleterious mutations within the frozen niche variation hypothesis. J Evol Biol 17: 651–62.

Pouteau S, Grandbastien MA, Boccara M (1994) Microbial elicitors of plant defense responses activate transcription of a retrotransposon. Plant J 5: 535-42.

Poveda AM, Le Clech M, Pasero P (2010) Transcription and replication: breaking the rules of the road causes genomic instability. Transcription 1:  99-102.

Powell CL, Swenberg JA, Rusyn I (2005) Expression of base excision DNA repair genes as a biomarker of oxidative DNA damage. Cancer Lett 229: 1-11.

Powell JD, Elshtein R, Forest DJ, Palladino MA (2002) Stimulation of hypoxia-inducible factor-1 alpha (HIF-1alpha) protein in the adult rat testis following ischemic injury occurs without an increase in HIF-1alpha messenger RNA expression. Biol Reprod 67: 995–1002.

Powell SJ, Bale JS (2004) Cold shock injury and ecological costs of rapid cold hardening in the grain aphid Sitobion avenae (Hemiptera: Aphididae). J Insect Physiol 50: 277–84.

Powers PA, Smithies O (1986) Short gene conversions in the human fetal globin gene region: a by-product of chromosome pairing during meiosis? Genetics 112: 343–58.

Poyton RO, Castello PR, Ball KA, Woo DK, Pan N (2009a) Mitochondria and hypoxic signaling: a new view. Ann NY Acad Sci 1177: 48-56.

Poyton RO, Ball KA, Castello PR (2009b) Mitochondrial generation of free radicals and hypoxic signaling. Trends Endocrinol Metab 20: 332–40.

Prabhakar S, Noonan JP, Paabo S, Rubin EM (2006) Accelerated evolution of conserved noncoding sequences in humans. Science 314: 786.

Prahlad V, Morimoto RI (2009) Integrating the stress response: lessons for neurodegenerative diseases from C. elegans. Trends Cell Biol19: 52-61.

Prakash S, Johnson RE, Prakash L (2005) Eukaryotic translesion synthesis DNA polymerases: Specificity of structure and function. Annu Rev Biochem 74: 317-53.

Prasad A, Croydon-Sugarman MJ, Murray RL, Cutter AD (2011) Temperature-dependent fecundity associates with latitude in Caenorhabditis briggsae. Evolution 65:52-63.

Prasad S, Kalra N, Shukla Y (2006) Modulatory effects of diallyl sulfide against testosterone- induced oxidative stress in Swiss albino mice. Asian J Androl 8: 719-23.

Prasad S, Kalra N, Singh M, Shukla Y (2008) Protective effects of lupeol and mango extract against androgen induced oxidative stress in Swiss albino mice. Asian J Androl 10: 313-8.

Prasad SM, Czepiel M, Cetinkaya C, Smigielska K, Weli SC, et al. (2009) Continuous hypoxic culturing maintains activation of Notch and allows long-term propagation of human embryonic stem cells without spontaneous differentiation. Cell Prolif 42: 63–74.

Pratt WB (1992) Control of steroid receptor function and cytoplasmic-nuclear transport by heat shock proteins. Bioessays 14: 841-8.

Pratt WB, Gehring U, Toft DO (1996) Molecular chaperoning of steroid hormone receptors. EXS 77: 79-95.

Precht H, Christophersen J,Hensel H (1955) Temperatur und Leben. Berlin, Germany: Springer.

Prentice AM, Rayco-Solon P, Moore SE (2005) Insights from the developing world: thrifty genotypes and thrifty phenotypes. Proc Nutr Soc 64: 153-61.

Prescott D (1994) The DNA of ciliated protozoa. Microbiol Rev 58: 233–67.

Presgraves DC (2005) Recombination enhances protein adaptation in Drosophila melanogaster. Curr Biol 15: 1651–6.

Preston CR, Flores CC, Engels WR (2006a) Differential usage of alternative pathways of double-strand break repair in Drosophila. Genetics 172: 1055–68.

Preston CR, Flores C, Engels W (2006b) Age-dependent usage of double-strand-break repair pathways. Curr Biol16: 2009-15.

Price GR (1972) Extension of covariance selection mathematics. Ann Hum Genet 35: 485-90.

Price TD, Schluter D, Heckman NE (1993) Sexual selection when the female directly benefits. Biol J Linn Soc 48: 187–211.

Price TD, Qvarnström A, Irwin DE (2003) The role of phenotypic plasticity in driving genetic evolution. Proc Biol Sci 270: 1433-40.

Print CG, Loveland KL (2000) Germ cell suicide: new insights into apoptosis during spermatogenesis. Bioessays 22: 423-30.

Prisco M, Liguoro A, Comitato R, Cardone A, D’Onghia B, et al. (2003) Apoptosis during spermatogenesis in the spotted ray Torpedo marmorata. Mol Reprod Dev 64: 341–8.

Pritchard JK, Pickrell JK, Coop G (2010) The genetics of human adaptation: hard sweeps, soft sweeps, and polygenic adaptation. Curr Biol 20: R208–215.

Pritchett TL, Tanner EA, McCall K (2009) Cracking open cell death in the Drosophila ovary. Apoptosis 14: 969–79.

Pritham EJ, Feschotte C, Wessler SR (2005) Unexpected diversity and differential success of DNA transposons in four species of Entamoeba Protozoans. Mol Biol Evol 22: 1751–63.

Pritham EJ, Feschotte C (2007) Massive amplification of rolling-circle transposons in the lineage of the bat Myotis lucifugus. Proc Natl Acad Sci USA 104: 1895–900.

Prochnik SE, Rokhsar D, Aboobaker AA (2007) Evidence for a microRNA expansion in the bilaterian ancestor. Dev Genes Evol 217: 73–7.

Prochownik EV, Li Y (2007) The ever expanding role for c-Myc in promoting genomic instability. Cell Cycle 6: 1024-9.

Prolla TA, Baker SM, Harris AC, Tsao JL, Yao X, et al. (1998) Tumour susceptibility and spontaneous mutation in mice deficient in Mlh1, Pms1 and Pms2 DNA mismatch repair. Nat Genet 18: 276–9.

Promislow DEL, Harvey PH (1990) Living fast and dying young: A comparative analysis of life-history variation among mammals. J Zool 220: 417–37.

Promislow D, Clobert J, Barbault R (1992) Life history allometry in mammals and squamate reptiles: Taxon-level effects. Oikos 65: 285–94.

Pröschel M, Zhang Z, Parsch J (2006) Widespread adaptive evolution of Drosophila genes with sex-biased expression. Genetics 174: 893–900.

Proulx SR (1999) Matings systems and the evolution of niche breadth. Am Nat 154: 89–98.

Proulx SR (2001) Female choice via indicator traits easily evolves in the face of recombination and migration. Evolution 55: 2401–11.

Proulx SR (2002) Niche shifts and expansion due to sexual selection. Evol Ecol Res 4: 351–69.

Proulx SR, Day T (2001) What can invasion analyses tell us about evolution under stochasticity? Selection 2: 1–16.

Proulx SR, Adler FR (2010) The standard of neutrality: still flapping in the breeze? J Evol Biol 23: 1339–50.

Provencher LM, Morse GE, Weeks AR, Normark BB (2005) Parthenogenesis in the Aspidiotus nerii complex (Hemiptera: Diaspididae): a single origin of a worldwide, ployphagous lineage associated with Cardinium bacteria. Ann Entomol Soc Am 98: 629–35.

Provine WB (1971) The Origins of Theoretical Population Genetics. Chicago, IL: University of Chicago Press.

Prudhomme M, Attaiech L, Sanchez G, Martin B,Claverys JP (2006) Antibiotic stress induces genetic transformability in the human pathogen Streptococcus pneumoniae. Science 313: 89–92.

Prud’homme N, Gans M, Masson M, Terzian C (1995) Flamenco, a gene controlling the gypsy retrovirus of Drosophila melanagaster. Genetics 139: 697–711.

Prunier AL, Malbruny B, Laurans M, Brouard J, Duhamel JF, Leclercq R (2003) High rate of macrolide resistance in Staphylococcus aureus strains from patients with cystic fibrosis reveals high proportions of hypermutable strains. J Infect Dis 187: 1709–16.

Prusiner SB (1991) Molecular biology of prion diseases. Science 252: 1515-22.

Przeworski M, Coop G, Wall JD (2005) The signature of positive selection on standing genetic variation. Evolution 59: 2312–23.

Puchta H, Dujon B, Hohn B (1993) Homologous recombination in plant cells is enhanced by in vivo induction of double strand breaks into DNA by a site-specific endonuclease. Nucleic Acids Res 21: 5034–40.

Puglisi R, Bevilacqua A, Carlomagno G, Lenzi A, Gandini L, et al. (2007) Mice overexpressing the mitochondrial phospholipid hydroperoxide glutathione peroxidase in male germ cells show abnormal spermatogenesis and reduced fertility. Endocrinology 148: 4302-9.

Puisségur MP, Mazure NM, Bertero T, Pradelli L, Grosso S, et al. (2011) miR-210 is overexpressed in late stages of lung cancer and mediates mitochondrial alterations associated withmodulation of HIF-1activity. Cell Death Differ 18: 465-78.

Pujol B, Pannell JR (2008) Reduced responses to selection after species range expansion. Science 321: 96.

Pulkkinen K, Malm T, Turunen M, Koistinaho J, Ylä-Herttuala S (2008) Hypoxia induces microRNA miR-210 in vitro and in vivo ephrin-A3 and neuronal pentraxin 1 are potentially regulated by miR-210. FEBS Lett 582: 2397–401.

Pumpernik D, Oblak B, Borstnik B (2008) Replication slippage versus point mutation rates in short tandem repeats of the human genome. Mol Genet Genomics 279: 53–61.

Punnett RC (1911) Mendelism. 3rd edn. New York, NY: MacMillan.

Punt PJ, Dingemanse MA, Kuyvenhoven A, Soede RD, Pouwels PH, van den Hondel CA (1990) Functional elements in the promoter region of the Aspergillus nidulans gpdA gene coding glyceradehyde-3-phosphate dehydrogenase. Gene 93: 101-9.

Purrington CB, Bergelson J (1999) Exploring the physiological basis of costs of herbicide resistance in Arabidopsis thaliana. Am Nat 154: S82-S91.

Purves D (1980) Neuronal competition. Nature 287: 585–6.

Pusey A, Wolf M (1996) Inbreeding avoidance in mammals. Trends Ecol Evol 11: 201–6.

Puurtinen M, Ketola T, Kotiaho JS (2005) Genetic compatibility and sexual selection. Trends Ecol Evol 20: 157–8.

Pybus OG, Rambaut A, Belshaw R, Freckleton RP, Drummond AJ, Holmes EC (2007) Phylogenetic evidence for deleterious mutation load in RNA viruses and its contribution to viral evolution. Mol Biol Evol 24: 845-52.

Pybus C, Pedraza-Reyes M, Ross CA, Martin H, Ona K, et al. (2010) Transcription-associated mutation in Bacillus subtilis cells under stress. J Bacteriol 192: 3321-8.

Qian SZ, Wang ZG (1984) Gossypol: a potential antifertility agent for males. Annu Rev Pharmacol Toxicol 24: 329-60.

Qin J, Calabrese P, Tiemann-Boege I, Shinde DN, Yoon SR, et al. (2007) The molecular anatomy of spontaneous germline mutations in human testes. PLoS Biol 5: e224.

Qu J, Liu GH, Huang B, Chen C (2007) Nitric oxide controls nuclear export of APE1/Ref-1 through S-nitrosation of cysteines 93 and 310. Nucleic Acids Res 35: 2522-32.

Quan X, Lim SO, Jung G (2011) Reactive oxygen species downregulate catalase expression via methylation of a CpG island in the Oct-1 promoter. FEBS Lett 585: 3436-41.

Quarmby LM (1994) Signal transduction in the sexual life of Chlamydomonas. Plant Mol Biol 26: 1271-87.

Quattro JM, Avise JC, Vrijenhoek RC (1992) Mode of origin and sources of genotypic diversity in triploid fish clones (Poeciliopsis: Poecilidae). Genetics 130: 621–8.

Quayle AP, Bullock S (2006) Modelling the evolution of genetic regulatory networks. J Theor Biol 238: 737–53.

Queitsch C, Sangster TA, Lindquist S (2002) Hsp90 as a capacitor of phenotypic variation. Nature 417: 618-24.

Queller DC, Strassman JE (2009) Beyond sociality: the evolution of organismality. Philos Trans R Soc B Biol Sci 364: 3143–55.

Quer J, Esteban JI, Cos J, Sauleda S, Ocaña L, et al. (2005) Effect of bottlenecking on evolution of the nonstructural protein 3 gene of hepatitis C virus during sexually transmitted acute resolving infection. J Virol 79: 15131–41.

Quesada H, Wenne R, Skibinski DO (1999) Interspecies transfer of female mitochondrial DNA is coupled with role-reversals and departure from neutrality in the mussel Mytilus trossulus. Mol Biol Evol 16: 655-65.

Quinn PG, Payne AH (1984) Oxygen-mediated damage of microsomal cytochrome P-450 enzymes in cultured Leydig cells. Role in steroidogenic desensitization. J Biol Chem 259: 4130–5.

Quinn PG, Payne AH (1985) Steroid product-induced, oxygen-mediated damage of microsomal cytochrome P-450 enzymes in Leydig cell cultures. Relationship to desensitization. J Biol Chem 260: 2092–9.

Quintana R, Kopcow L, Sueldo C, Marconi G, Rueda NG, Barañao RI (2004) Direct injection of vascular endothelial growth factor into the ovary of mice promotes follicular development. Fertil Steril 82 Suppl 3: 1101-5.

Quinto-Alemany D, Canerina-Amaro A, Hernández-Abad LG, Machín F, Romesberg FE, et al. (2012) Yeasts acquire resistance secondary to antifungal drug treatment by adaptive mutagenesis. PLoS ONE 7: e42279.

Raaphorst GP, Ng CE, Yang DP (1999) Thermal radiosensitization and repair inhibition in human melanoma cells: a comparison of survival and DNA double-strand breaks. Int J Hyperthermia 15: 17-27.

Rabin D, Gold PW, Margioris AN, Chrousos GP (1988) Stress and reproduction: physiologic and pathophysiologic interactions between the stress and reproductive axes. Adv Rap Med Biol 245: 377-87.

Rabosky DL (2012) Positive correlation between diversification rates and phenotypic evolvability can mimic punctuated equilibrium on molecular phylogenies. Evolution 66: 2622-7.

Radeke HH, Meier B, Topley N, Floge J, Habermehl GG, Resch K (1990) Interleukin 1-alpha and tumor necrosis factor-alpha induce oxygen radical production in mesangial cells. Kidney Int 37: 767-75.

Radford EJ, Ferrón SR, Ferguson-Smith AC (2011) Genomic imprinting as an adaptative model of developmental plasticity. FEBS Lett 585: 2059-66.

Radi R, Turrens JF, Chang LY, Bush KM, Crapo JD, Freeman BA (1991) Detection of catalase in rat heart mitochondria. J Biol Chem 266: 22028–34.

Radicella JP, Park PU, Fox MS (1995) Adaptive mutation in Escherichia coli: a role for conjugation. Science 268: 418-20.

Radman M (1999) Enzymes of evolutionary change. Nature 401: 866–9.

Radman M, Matic I, Taddei F (1999) Evolution of evolvability. Ann NY Acad Sci 870: 146-55.

Radman M., Taddei F, Matic I (2000) Evolution-driving genes. Res Microbiol 151: 91–5.

Radwan J (2004) Effectiveness of sexual selection in removing mutations induced with ionizing radiation. Ecol Lett 7: 1149–54.

Radwan J, Tkacz A, Kloch A (2008) MHC and preferences for male odour in the bank vole. Ethology 114: 827-33.

Radzikowski J (2013) Resistance of dormant stages of planktonic invertebrates to adverse environmental conditions.J Plankton Res (2013): 1–17. doi:10.1093/plankt/fbt032.[Epub ahead of print]

Raes M, Michiels C, Remacle J (1987) Comparative study of the enzymatic defense systems against oxygen-derived free radicals: the key role of glutathione peroxidase. Free Radic Biol Med 3: 3-7.

Raff MC (1992) Social controls on cell survival and cell death. Nature 356: 397–400.

Raffel TR, Martin LB, Rohr JR (2008) Parasites as predators: unifying natural enemy ecology. Trends Ecol Evol 23: 610–8.

Raghavan SC, Raman MJ (2004) Nonhomologous end joining of complementary and noncomplementary DNA termini in mouse testicular extracts. DNA Repair 3: 1297-310.

Rahbek C, Graves GR (2001) Multiscale assessment of patterns of avian species richness. Proc Natl Acad Sci USA 98: 4534–9.

Rahman I (2000) Regulation of nuclear factor-kappa B, activator protein-1, and glutathione levels by tumor necrosis factor-alpha and dexamethasone in alveolar epithelial cells. Biochem Pharmacol 60: 1041-9.

Rahman I, Gilmour PS, Jimenez LA, MacNee W (2002) Oxidative stress and TNF-alpha induce histone acetylation and NF-kappaB/AP-1 activation in alveolar epithelial cells: potential mechanism in gene transcription in lung inflammation. Mol Cell Biochem 234-235: 239-48.

Rahman I, Marwick J, Kirkham P (2004) Redox modulation of chromatin remodeling: impact on histone acetylation and deacetylation, NF-kappaB and pro-inflammatory gene expression. Biochem Pharmacol 68: 1255-67.

Rai K, Huggins IJ, James SR, Karpf AR, Jones DA, Cairns BR (2008) DNA demethylation in zebrafish involves the coupling of a deaminase, a glycosylase, and gadd45. Cell 135: 1201–12.

Raineri I, Carlson EJ, Gacayan R, Carra S, Oberley TD, Huang TT, Epstein CJ (2001) Strain-dependent high-level expression of a transgene for manganese superoxide dismutase is associated with growth retardation and decreased fertility. Free Radic Biol Med 31: 1018–30.

Rainey PB (1999) Evolutionary genetics: The economics of mutation. Curr Biol 9: R371–R373.

Rainey PB, Beaumont HJ, Ferguson GC, Gallie J, Kost C, et al. (2011) The evolutionary emergence of stochastic phenotype switching in bacteria. Microb Cell Fact 10 Suppl 1: S14.

Raj A, van Oudenaarden A (2008) Nature, nurture, or chance: stochastic gene expression and its consequences. Cell 135: 216–26.

Rajewsky N (2006) microRNA target predictions in animals. Nat Genet 38(Suppl): S8-13.

Rajon E, Venner S, Menu F (2009) Spatially heterogeneous stochasticity and the adaptive diversification of dormancy. J Evol Biol 22: 2094-103.

Raju VS, McCoubrey WK Jr, Maines MD (1997) Regulation of heme oxygenase-2 by glucocorticoids in neonatal rat brain: characterization of a functional glucocorticoid response element. Biochim Biophys Acta 1351: 89-104.

Rakyan VK, Blewitt ME, Druker R, Preis JI, Whitelaw E (2002) Metastable epialleles in mammals. Trends Genet 18: 348–51.

Rakyan VK, Chong S, ChampME, Cuthbert PC, Morgan HD, et al. (2003) Transgenerational inheritance of epigenetic states at the murine Axin(Fu) allele occurs after maternal and paternal transmission. Proc Natl Acad Sci USA 100: 2538–43.

Rakyan VK, Beck S (2006) Epigenetic variation and inheritance in mammals. Curr Opin Genet Dev 16: 573-7.

Rakyan VK, Down TA, Balding DJ, Beck S (2011) Epigenome-wide association studies for common human diseases. Nat Rev Genet 12: 529–41.

Ralls K, Ballou JD, Templeton A (1988) Estimates of lethal equivalents and the cost of inbreeding in mammals. Conserv Biol 2: 185–93.

Ralph SJ, Rodríguez-Enríquez S, Neuzil J, Saavedra E, Moreno-Sánchez R (2010) The causes of cancer revisited: ‘‘Mitochondrial malignancy” and ROS-induced oncogenic transformation – Why mitochondria are targets for cancer therapy. Mol Aspects Med 31: 145–70.

Ram Y, Hadany L (2012) The evolution of stress-induced hypermutation in asexual populations. Evolution DOI: 10.1111/j.1558-5646.2012.01576.x

Ramakrishna PA, Prasad MR (1967) Changes in the male reproductive organs of Loris tardigradus lydekkerianus (Cabrera). Folia Primatol (Basel) 5: 176-89.

Ramalho-Santos J, Varum S, Amaral S, Mota PC, Sousa AP, Amaral A (2009) Mitochondrial functionality and reproduction: from gonads and gametes to embryos and embryonic stem cells. Hum Reprod Update 15: 553-72.

Ramana CV, Boldogh I, Izumi T, Mitra S (1998) Activation of apurinic/apyrimidinic endonuclease in human cells by reactive oxygen species and its correlation with their adaptive response to genotoxicity of free radicals. Proc Natl Acad Sci USA 95: 5061–6.

Ramesh MA, Malik SB, Logsdon JM Jr (2005) A phylogenomic inventory of meiotic genes; evidence for sex in Giardia and an early eukaryotic origin of meiosis. Curr Biol 15: 185-91.

Ramos-Mejia V, Muñoz-Lopez M, Garcia-Perez JL, Menendez P (2010) iPSC lines that do not silence the expression of the ectopic reprogramming factors may display enhanced propensity to genomic instability. Cell Res 20: 1092-5.

Ramsey J, Schemske DW (2002) Neopolyploidy in flowering plants. Annu Rev Ecol Syst 33: 589–639.

Ramsey M, Vaughton G (2000) Pollen quality limits seed set in Burchardia umbellata (Cholchicaceae). Am J Bot 87: 845–52.

Rana TM (2007) Illuminating the silence: understanding the structure and function of small RNAs. Nat Rev Mol Cell Biol 8: 23–36.

Rand DA, Keeling M, Wilson HB (1995) Invasion, stability and evolution to criticality in spatially extended, artificial host-pathogen ecologies. Proc R Soc Lond Ser B 259: 55–63.

Rand DM (1994) Thermal habit, metabolic rate and the evolution of mitochondrial DNA. Trends Ecol Evol 9: 125–31.

Rand DM (2001) The units of selection of mitochondrial DNA. Annu Rev Ecol Syst 32: 415-48.

Rand DM, Harrison RG (1986) Mitochondrial DNA transmission genetics in crickets. Genetics 114: 955–70.

Rand DM, Dorfsman M, Kann LM (1994) Neutral and non-neutral evolution of Drosophila mitochondrial DNA. Genetics 138: 741-56.

Rand DM, Kann LM (1996) Excess amino acid polymorphism in mitochondrial DNA: contrasts among genes from Drosophila, mice, and humans. Mol Biol Evol 13: 735-48.

Rand DM, Kann LM (1998) Mutation and selection at silent and replacement sites in the evolution of animal mitochondrial DNA. Genetica 102/103: 393-407.

Rand DM, Weinreich DM, Cezairliyan BO (2000) Neutrality tests of conservative-radical amino acid changes in nuclear- and mitochondrially-encoded proteins. Gene 261: 115-25.

Rand DM, Clark AG, Kann LM (2001) Sexually antagonistic cytonuclear fitness effects in Drosophila melanogaster. Genetics 159: 173-87.

Randall MGM (1982) The dynamics of an insect population throughout its altitudinal distribution: Coleophora alticolella (Lepidoptera) in northern England. J Anim Ecol 51: 993–1016.

Rando OJ, Verstrepen KJ (2007) Timescales of genetic and epigenetic inheritance. Cell 128: 655–68.

Rangwala SH, Elumalai R, Vanier C, Ozkan H, Galbraith DW, Richards EJ (2006) Meiotically stable natural epialleles of Sadhu, a novel Arabidopsis retroposon. PLoS Genet 2: e36.

Rankin DJ (2007) Resolving the tragedy of the commons: the feedback between intraspecific conflict and population density. J Evol Biol 20: 173-80.

Rankin DJ (2011) Kin selection and the evolution of sexual conflict. J Evol Biol 24:71–81.

Rankin DJ, López-Sepulcre A (2005) Can adaptation lead to extinction? Oikos 111: 616-9.

Rankin DJ, Kokko H (2006) Sex, death and tragedy. Trends Ecol Evol 21: 225-6.

Rannan-Eliya SV, Taylor IB, De Heer IM, Van Den Ouweland AM, Wall SA, Wilkie AO (2004) Paternal origin of FGFR3 mutations in Muenke-type craniosynostosis. Hum Genet 115: 200–7.

Ranta E (1979) Population biology of Darwinula stevensoni (Crustacea, Ostracoda) in an oligotrophic lake. Ann Zool Fenn 16: 28–35.

Ranz J, Castillo-Davis C, Meiklejohn C, Hartl D (2003) Sex-dependent gene expression and evolution of the Drosophila transcriptome. Science 300: 1742-5.

Rao MW, Gangadharan B (2008) Antioxidant potential of melatonin against mercury induced intoxication in spermatozoa in vitro. Toxicol In Vitro 22: 935–42.

Rao YS, Wang ZF, Chai XW, Wu GZ, Zhou M, et al. (2010) Selection for the compactness of highly expressed genes in Gallus gallus. Biol Direct 5: 35.

Rapp RA,Wendel JF (2005) Epigenetics and plant evolution. New Phytol 168: 81–91.

Raser JM, O’Shea EK (2004) Control of stochasticity in eukaryotic gene expression. Science 304: 1811–4.

Raser JM, O’Shea EK (2005) Noise in gene expression: origins, consequences, and control. Science 309: 2010–3.

Rasmussen AT (1917) Seasonal changes in interstitial cells of the testis in the woodchuck (Marmota monax). Am J Anat 22: 475-515.

Rasmussen JE, Torres-Aleman I, MacLusky NJ, Naftolin F, Robbins RJ (1990) The effects of estradiol on the growth patterns of estrogen receptor-positive hypothalamic cell lines. Endocrinology 126: 235-240.

Rassoulzadegan M, Grandjean V, Gounon P, Vincent S, Gillot I, Cuzin F (2006) RNA-mediated non-Mendelian inheritance of an epigenetic change in the mouse. Nature 441: 469–74.

Ratcliffe PJ (2013) Oxygen sensing and hypoxia signalling pathways in animals: the implications of physiology for cancer. J Physiol 591: 2027-42.

Ratcliff WC, Denison RF (2010) Individual-level bet hedging in the bacterium Sinorhizobium meliloti. Curr Biol 20: 1740-4.

Rathke C, Baarends WM, Jayaramaiah-Raja S, Bartkuhn M, Renkawitz R, Renkawitz-Pohl R (2007) Transition from a nucleosome-based to a protamine-based chromatin configuration during spermiogenesis in Drosophila. J Cell Sci 120: 1689–700.

Ratner VA, Zabanov SA, Kolesnikova OV, Vasilyeva LA (1992) Induction of the mobile genetic element Dm-412 transpositions in the Drosophila genome by heat-shock treatment. Proc Natl Acad Sci USA 89: 5650–4.

Ratnieks FLW, Wenseleers T (2005) Policing insect societies. Science 307: 54–6.

Ratnieks FLW, Foster KR, Wenseleers T (2006) Conflict resolution in insect societies. Annu Rev Entomol 51: 581–608.

Rato L, Alves MG, Socorro S, Duarte AI, Cavaco JE, Oliveira PF (2012) Metabolic regulation is important for spermatogenesis. Nat Rev Urol 9: 330-8.

Rattray AJ, Shafer BK, McGill CB, Strathern JN (2002) The roles of REV3 and RAD57 in double-strandbreak-repair-induced mutagenesis of Saccharomyces cerevisiae. Genetics 162: 1063-77.

Rattray AJ, Strathern JN (2003) Error-prone DNA polymerases: when making a mistake is the only way to get ahead. Annu Rev Genet 37: 31-66.

Rauch EM, Sayama H, Bar-Yam Y (2002) The relationship between measures of fitness and time scale in evolution. Phys Rev Lett 88: 228101/1–228101/4.

Rauch EM, Sayama H, Bar-Yam Y (2003) Dynamics and genealogy of strains in spatially extended host-pathogen systems. J Theor Biol 221: 655-64.

Rausher MD (1988) Is coevolution dead? Ecology 69: 898–901.

Raver-Shapira N, Marciano E, Meiri E, Spector Y, Rosenfeld N, et al. (2007) Transcriptional activation of miR-34a contributes to p53-mediated apoptosis. Mol Cell 26: 731–43.

Ravi R, Mookerjee B, Bhujwalla ZM, Sutter CH, Artemov D, et al. (2000) Regulation of tumor angiogenesis by p53-induced degradation of hypoxia-inducible factor 1alpha. Genes Dev 14: 34-44.

Ravindranath N, Little-Ihrig LL, Phillips HS, Ferrara N, Zeleznik AJ (1992) Vascular endothelial growth factor messenger ribonucleic acid expression in the primate ovary. Endocrinology 131: 254–60.

Ravussin E (2002) Cellular sensors of feast and famine. J Clin Invest 109: 1537–40.

Rawi A, Louvet-Vallée S, Djeddi A, Sachse M, Culetto E, et al. (2011) Postfertilization autophagy of sperm organelles prevents paternal mitochondrial DNA transmission. Science 334: 1144–7.

Ray DA, Pagan HJT, Thompson ML, Stevens RD (2007) Bats with hATs: evidence for recent DNA transposon activity in genus Myotis. Mol Biol Evol 24: 632–9.

Ray S, Lee C, Hou T, Bhakat KK, Brasier AR (2010) Regulation of signal transducer and activator of transcription 3 enhanceosome formation by apurinic/apyrimidinic endonuclease 1 in hepatic acute phase response. Mol Endocrinol 24: 391-401.

Raymond J, Segre D (2006) The effect of oxygen on biochemical networks and the evolution of complex life. Science 311: 1764–7.

Raz E (2000) The function and regulation of vasa-like genes in germ-cell development. Genome Biol 1: 1017.1–1017.6.

Razin A, Riggs AD (1980) DNA methylation and gene function. Science 210: 604–10.

Rebollo R, Horard B, Hubert B, Vieira C (2010) Jumping genes and epigenetics: towards new species. Gene 454: 1-7.

Rebollo R, Romanish MT, Mager DL (2012) Transposable elements: an abundant and natural source of regulatory sequences for host genes. Annu Rev Genet 46: 21-42.

Redd AD, Collinson-Streng AN, Chatziandreou N, Mullis CE, Laeyendecker O, et al. (2012) Previously transmitted HIV-1 viral strains are preferentially selected during subsequent sexual transmissions. J Infect Dis 206: 1433–42.

Redding GP, Bronlund JE, Hart AL (2007) Mathematical modelling of oxygen transport-limited follicle growth. Reproduction 133: 1095-106.

Redding GP, Bronlund JE, Hart AL (2008) Theoretical investigation into the dissolved oxygen levels in follicular fluid of the developing human follicle using mathematical modeling. Reprod Fertil Dev 20: 408–17.

Reddy N, Kasukurthi KB, Mahla RS, Pawar RM, Goel S (2012) Expression of vascular endothelial growth factor (VEGF) transcript and protein in the testis of several vertebrates, including endangered species. Theriogenology 77: 608-14.

Redfield RJ (1988) Evolution of bacterial transformation: is sex with dead cells ever better than no sex at all? Genetics 119: 213-21.

Redfield RJ (1993a) Evolution of natural transformation: testing the DNA repair hypothesis in Bacillus subtilis and Haemophilus influenzae. Genetics 133: 755-61.

Redfield RJ (1993b) Genes for breakfast: the have-your-cake-and-eat-it-too of bacterial transformation. J Hered 84:400–4.

Redfield RJ (1994) Male mutation rates and the cost of sex for females. Nature 369: 145-7.

Redfield RJ (2001) Do bacteria have sex? Nature Rev Genet 2: 634–9.

Redfield RJ, Schrag MR, Dean AM (1997) The evolution of bacterial transformation: sex with poor relations. Genetics 146: 27-38.

Redon CE, Dickey JS, Nakamura AJ, Kareva IG, Naf D, et al. (2010) Tumors induce complex DNA damage in distant proliferative tissues in vivo. Proc Natl Acad Sci USA 107: 17992–7.

Redondo P, Jimenez E, Perez A, Garcia-Foncillas J (2000) N-acetylcysteine downregulates vascular endothelial growth factor production by human keratinocytes in vitro. Arch Dermatol Res 292: 621-8.

Reed DH, Frankham R (2003) Correlation between fitness and genetic diversity. Conserv Biol 17: 230–7.

Rees M, Jessica C, Metcalf E, Childs DZ (2010) Bet-hedging as an evolutionary game: the trade-off between egg size and number. Proc Biol Sci 277: 1149-51.

Reese HW (2005) A conceptual analysis of selectionism: Part I & II. Behav Dev Bull 1: 8-16.

Reetz MT, Bocola M, Carballeira JD, Zha D, Vogel A (2005) Expanding the range of substrate acceptance of enzymes: Combinatorial active-site saturation test. Angew Chem Int Ed 44: 4192–6.

Reeve HK, Hölldobler B (2007) The emergence of a superorganism through intergroup competition. Proc Natl Acad Sci USA 104: 9736–40.

Regueira E, Sassone AG, Scaia MF, Volonteri MC, Ceballos NR (2013) Seasonal changes and regulation of the glucocorticoid receptor in the testis of the toad Rhinella arenarum. J Exp Zool A Ecol Genet Physiol 319: 39-52.

Rehman J (2010) Empowering self-renewal and differentiation: the role of mitochondria in stem cells. J Mol Med (Berl) 88: 981-6.

Reichard M, Spence R, Bryjová A, Bryja J, Smith C (2012) Female rose bitterling prefer MHC-dissimilar males: experimental evidence. PLoS One 7: e40780.

Reichmann B, Ephrussi A (2001) Axis formation during Drosophila oogenesis. Curr Opin Genet Dev 11: 374-83.

Reik W, Collick A, Norris ML, Barton SC, Surani MA (1987) Genomic imprinting determines methylation of parental alleles in transgenic mice. Nature 328: 248–51.

Reik W, Dean W, Walter J (2001) Epigenetic reprogramming in mammalian development. Science 293: 1089–93.

Reik W, Santos F, Mitsuya K, Morgan H, Dean W (2003) Epigenetic asymmetry in the mammalian zygote and early embryo: relationship to lineage commitment? Philos Trans R Soc Lond B Biol Sci 358: 1403–9.

Reimers JM, Schmidt KH, Longacre A, Reschke DK, Wright BE (2004) Increased transcription rates correlate with increased reversion rates in leuB and argH Escherichia coli auxotrophs. Microbiology 150: 1457–66.

Reintal M, Tali K, Haldna M, Kull T (2010) Habitat preferences as related to the prolonged dormancy of perennial herbs and ferns. Plant Ecol 210: 111–23.

Reiss D, Mager DL (2007) Stochastic epigenetic silencing of retrotransposons: does stability come with age? Gene 390: 130-5.

Reiss MJ (1989) The allometry of growth and reproduction. Cambridge, UK: Cambridge University Press.

Reiter RJ (1991) Pineal melatonin: cell biology of its synthesis and of its physiological interactions. Endocr Rev 12: 151–80.

Reiter RJ, Tan DX, Osuna C, Gitto E (2000) Actions of melatonin in the reduction of oxidative stress: a review. J Biomed Sci 7: 444–58.

Reiter RJ, Tan DX, Manchester LC, Qi W (2001) Biochemical reactivity of melatonin with reactive oxygen and nitrogen species: a review of the evidence. Cell Biochem Biophys 34: 237–56.

Reiter RJ, Tan DX, Manchester LC, Paredes SD, Mayo JC, Sainz RM (2009) Melatonin and reproduction revisited. Biol Reprod 81: 445-56.

Reitsma PH, Rothberg PG, Astrin SM, Trial J, Bar-Shavit Z, et al. (1983) Regulation of c-Myc gene expression in HL-60 leukaemia cells by a vitamin D metabolite. Nature 306: 492–4.

Remacle J, Raes M, Toussaint O, Renard P, Rao G (1995) Low levels of reactive oxygen species as modulators of cell function. Mutat Res 316: 103-22.

Remold SK, Lenski RE (2004) Pervasive joint influence of epistasis and plasticity on mutational effects in Escherichia coli. Nat Genet 36: 423–6.

Ren R, Xu X, Lin T, Weng S, Liang H, et al. (2011) Cloning, characterization, and biological function analysis of the SidT2 gene from Siniperca chuatsi. Dev Comp Immunol 35: 692-701.

Renault AD, Sigal YJ, Morris AJ, Lehmann R (2004) Soma-germ line competition for lipid phosphate uptake regulates germ cell migration and survival. Science 305: 1963-6.

Resnitzky D, Yarden A, Zipori D, Kimchi A (1986) Autocrine β-related interferon controls c-Myc suppression and growth arrest during hematopoietic cell differentiation. Cell 46: 31–40.

Retterstøl K, Tran TN, Haugen TB, Christophersen BO (2001a) Metabolism of very long chain polyunsaturated fatty acids in isolated rat germ cells. Lipids 36: 601-6.

Retterstøl K, Haugen TB, Tran TN, Christophersen BO (2001b) Studies on the metabolism of essential fatty acids in isolated human testicular cells. Reproduction 121: 881-7.

Reusch TBH, Bostrom C, Stam WT, Olsen JL (1999) An ancient eelgrass clone in the Baltic. Mar Ecol-Prog Ser 183: 301–4.

Reusch TBH, Ehlers A, Haemmerli A, Worm B (2005) Ecosystem recovery after climatic extremes enhanced by genotypic diversity. Proc Natl Acad Sci USA 102: 2826–31.

Reusch TBH, Hughes AR (2006) The emerging role of genetic diversity for ecosystem functioning: estuarine macrophytes as models. Estuar Coast 29: 170–5.

Reuter S, Schnekenburger M, Cristofanon S, Buck I, Teiten MH, et al. (2009) Tumor necrosis factor alpha induces gamma-glutamyltransferase expression via nuclear factor-kappaB in cooperation with Sp1. Biochem Pharmacol 77: 397-411.

Reyer H-U, Frei G, Som C (1999) Cryptic female choice: Frogs reduce clutch size when amplexed by undesired males. Proc R SocBiol Sci B 266: 2101–7.

Reyes JG, Farias JG, Henríquez-Olavarrieta S, Madrid E, Parraga M, et al. (2012) The hypoxic testicle: physiology and pathophysiology.Oxid Med Cell Longev 2012: 929285. doi: 10.1155/2012/929285.

Reylea RA (2002) Costs of phenotypic plasticity. Am Nat 159: 272–82.

Reynier P, Chretien MF, Savagner F, Larcher G, Rohmer V, et al. (1998) Long PCR analysis of human gamete mtDNA suggests defective mitochondrial maintenance in spermatozoa and supports the bottleneck theory for oocytes. Biochem Biophys Res Commun 252: 373-7.

Reynier P, May-Panloup P, Chrétien MF, Morgan CJ, Jean M, et al. (2001) Mitochondrial DNA content affects the fertilizability of human oocytes. Mol Hum Reprod 7: 425-9.

Reynolds AE, Felton J, Wright A (1981) Insertion of DNA activates the cryptic bgl operon in E. coli K12. Nature 293: 625-9.

Reznick D (1985) Costs of reproduction: an evaluation of the empirical evidence. Oikos 44: 257-67.

Reznick D, Endler JA (1982) The impact of predation on life history evolution in Trinidadian guppies (Poecilia reticulata). Evolution 36: 160–77.

Reznick DA, Bryga H, Endler JA (1990) Experimentally induced life-history evolution in a natural population. Nature 346: 357–359.

Reznick DN, Shaw FH, Rodd FH, Shaw RG (1997) Evaluation of the rate of evolution in natural populations of guppies (Poecilia reticulata). Science 275: 1934–7.

Reznick DN, Ghalambor C (1999) Sex and death. Science 286: 2458–9.

Reznick DN, Ghalambor CK (2001) The population ecology of contemporary adaptations: what empirical studies reveal about the conditions that promote adaptive evolution. Genetica 112-13: 183-98.

Reznick D, Bryant MJ, Bashey F (2002) r- and K-selection revisited: the role of population regulation in life-history evolution. Ecology 83: 1509–20.

Reznick DN, Bryant MJ, Roff D, Ghalambor CK, Ghalambor DE (2004a) Effect of extrinsic mortality on the evolution of senescence in guppies. Nature 431: 1095–9.

Reznick D, Rodd H, Nunney L (2004b) Empirical evidence for rapid evolution. In: Ferrière R, Dieckmann U, Couvet D, eds. Evolutionary Conservation Biology. Cambridge, UK: Cambridge University Press. pp 101–118.

Rezvani HR, Mahfouf W, Ali N, Chemin C, Ged C, et al. (2010) Hypoxia-inducible factor-1alpha regulates the expression of nucleotide excision repair proteins in keratinocytes. Nucleic Acids Res 38: 797-809.

Rhee SG (1999) Redox signaling: hydrogen peroxide as intracellular messenger. Exp Mol Med 31: 53-9.

Rhee SG (2006) Cell signaling. H2O2, a necessary evil for cell signaling. Science 312: 1882–3.

Rhee SG, Kang SW, Chang TS, Jeong W, Kim K (2001) Peroxiredoxin, a novel family of peroxidases. IUBMB Life 52: 35–41.

Rhee SG, Yang KS, Kang SW, Woo HA, Chang TS (2005a) Controlled elimination of intracellular H2O2: regulation of peroxiredoxin, catalase, and glutathione peroxidase via post-translational modification. Antioxid Redox Signal 7: 619–26.

Rhee SG, Kang SW, Jeong W, Chang TS, Yang KS, Woo HA (2005b) Intracellular messenger function of hydrogen peroxide and its regulation by peroxiredoxins. Curr Opin Cell Biol 17: 183–9.

Rhiner C, Díaz B, Portela M, Poyatos JF, Fernández-Ruiz I, et al. (2009) Persistent competition among stem cells and their daughters in the Drosophila ovary germline niche. Development 136: 995-1006.

Rhiner C, López-Gay JM, Soldini D, Casas-Tinto S, Martín FA, et al. (2010) Flower forms an extracellular code that reveals the fitness of a cell to its neighbors in Drosophila. Dev Cell 18: 985-98.

Rhoades DF (1979) Evolution of plant chemical defense against herbivores. In: Rosenthal GA, Janzen DH, eds. Herbivores: their interaction with plant secondary metabolites. New York, NY: Academic Press. pp 3–54.

Rhodes D, Klug A (1993) Zinc fingers. Sci Am 268: 56–65.

Rhomberg LR, Joseph S, Singh RS (1985) Seasonal variation and clonal selection in cyclically parthenogenetic rose aphids (Macrosiphum rosae). Can J Genet Cytol 27: 224–32.

Ribera I, Dolédec S, Downie IS, Foster GN (2001) Effect of land disturbance and stress on species traits: a three table analysis of ground beetle assemblages. Ecology 82: 1112–29.

Riccioli A, Starace D, D'Alessio A, Starace G, Padula F, et al. (2000) TNF-α and IFN-γ regulate expression and function of the Fas system in the seminiferous epithelium. J Immunol 165: 743–9.

Riccioli A, Salvati L, D'Alessio A, Starace D, Giampietri C, et al. (2003) The Fas system in the seminiferous epithelium and its possible extra-testicular role. Andrologia 35: 64-70.

Ricci C (1987) The ecology of bdelloids: How to be successful. Hydrobiologia 147: 117–27.

Ricci C (1998) Anhydrobiotic capabilities of bdelloid rotifers. Hydrobiologia 387/388: 321–6.

Ricci C (1991) Comparison of five strains of a parthenogenetic species, Macrotrachela quadricornifera (Rotifera, Bdelloidea). Hydrobiologia 211: 147-55.

Ricci C (2001) Dormancy patterns in rotifers. Hydrobiologia 446/447: 1–11.

Ricci C, Vaghi L, Manzini ML (1987) Desiccation of rotifers (Macrotrachela quadricornifera): survival and reproduction. Ecology 68: 1488–94.

Ricci C, Covino C (2005) Anhydrobiosis of Adineta sp 1: Costs and benefits. Hydrobiologia 546: 307–14.

Ricci C, Caprioli M, Fontaneto D (2007) Stress and fitness in parthenogens: Is dormancy a key feature for bdelloid rotifers? BMC Evol Biol 7 (Suppl 2): S9.

Rice JC, Allis CD (2001) Histone methylation versus histone acetylation: new insights into epigenetic regulation. Curr Opin Cell Biol 13: 263-73.

Rice WR (1984) Sex-chromosomes and the evolution of sexual dimorphism. Evolution 38: 735–42.

Rice WR (1992) Sexually antagonistic genes: experimental evidence. Science 256: 1436–9.

Rice WR (1996) Sexually antagonistic male adaptation triggered by experimental arrest of female evolution. Nature 381: 232–4.

Rice WR (1998a) Male fitness increases when females are eliminated from gene pool: implications for the Y chromosome. Proc Natl Acad Sci USA 95: 6217–21.

Rice WR (1998b) Intergenomic conflict, interlocus antagonistic coevolution, and the evolution of reproductive isolation. In: Howard DJ, Berlocher SH, eds. Endless Forms: Species and Speciation. Oxford, UK: Oxford University Press. pp 261–270.

Rice WR (2002) Experimental tests of the adaptive significance of sexual recombination. Nat Rev Genet 3: 241–51.

Rice WR, Holland BT (1997) The enemies within: intergenomic conflict, interlocus contest evolution (ICE), and the intraspecific Red Queen. Behav Ecol Sociobiol 41: 1-10.

Rice WR, Chippindale AK (2001) Sexual recombination and the power of natural selection. Science 294: 555–9.

Rice WR, Chippindale AK (2002) The evolution of hybrid infertility: perpetual coevolution between gender-specific and sexually antagonistic genes. Genetica 116: 179-88.

Rice WR, Linder JE, Friberg U, Lew TA, Morrow EH, Stewart AD (2005) Inter-locus antagonistic coevolution as an engine of speciation: assessment with hemiclonal analysis. Proc Natl Acad Sci USA 102 (Suppl.): 6527–34.

Rice WR, Friberg U (2007) Genomic clues to an ancient asexual scandal. Genome Biol 8: 232

Richard GF, Kerrest A, Dujon B (2008) Comparative genomics and molecular dynamics of DNA repeats in eukaryotes. Microbiol Mol Biol Rev 72: 686-727.

Richards AJ (1989) A comparison of within-plant karyological heterogeneity between agamospermous and sexual Taraxacum (Compositae) as assessed by the nucleolar organizer chromosome. Plant Syst Evol 163: 177–85.

Richards AJ (1997) Plant breeding systems. 2nd edn. London, UK: Chapman & Hall.

Richards CL, Pennings SC, Donovan LA (2005) Habitat range and phenotypic variation in salt marsh plants. Plant Ecol 176: 263–73.

Richards CL, Bossdorf O, Muth NZ, Gurevitch J, Pigliucci M (2006) Jack of all trades, master of some? On the role of phenotypic plasticity in plant invasions. Ecol Lett 9: 981–93.

Richards CL, Walls RL, Bailey JP, Parameswaran R, George T, Pigliucci M (2008) Plasticity in salt tolerance traits allows for invasion of novel habitat by Japanese knotweed s. l. (Fallopia japonica and F. bohemica, Polygonaceae). Am J Bot 95: 931–42.

Richards EJ (2006) Inherited epigenetic variation—revisiting soft inheritance. Nat Rev Genet 7: 395–401.

Richards EJ (2011) Natural epigenetic variation in plant species: a view from the field. Curr Opin Plant Biol 14: 204-9.

Richards FM, Payne SJ, Zbar B, Affara NA, Ferguson-Smith MA, Maher ER (1995) Molecular analysis of de novo germline mutations in the von Hippel-Lindau disease gene. Hum Mol Genet 4: 2139–43.

Richards RI, Sutherland GR (1994) Simple repeat DNA is not replicated simply. Nat Genet 6: 114-6.

Richardson AR, Yu Z, Popovic T, Stojiljkovic I (2002) Mutator clones of Neisseria meningitidis in epidemic serogroup A disease. Proc Natl Acad Sci USA 99: 6103–7.

Richardson DS, Komdeur J, Burke T, von Schantz T (2005) MHC-based patterns of social and extra-pair mate choice in the Seychelles warbler. Proc R Soc Lond B 272: 759-67.

Richardson SJ, Senikas V, Nelson JF (1987) Follicular depletion during the menopausal transition: Evidence for accelerated loss and ultimate exhaustion. J Clin Endocrinol Metab 65: 1231–7.

Richburg JH, Nanez A, Williams LR, Embree ME , Boekelheide K (2000) Sensitivity of testicular germ cells to toxicant-induced apoptosis in gld mice that express a nonfunctional form of Fas ligand. Endocrinology 141: 787-93.

Richner H, Christe P, Oppliger A (1995) Paternal investment affects prevalence of malaria. Proc Natl Acad Sci USA 92: 1192-4.

Richter C (1997) Reactive oxygen and nitrogen species regulate mitochondrial Ca2+ homeostasis and respiration. Biosci Rep 17: 53–66.

Richter K, Buchner J (2001) Hsp90: chaperoning signal transduction. J Cell Physiol 188: 281–90.

Rico C, Rico I, Hewitt G (1996) 470 million years of conservation of microsatellite loci among fish species. Proc R Soc Lond B Biol Sci 263: 549–57.

Ricote M, Alfaro JM, García-Tuñón I, Arenas MI, Fraile B, et al. (2002) Control of the annual testicular cycle of the marbled-newt by p53, 21, and Rb gene products. Mol Reprod Dev 63: 202–9.

Ricquier D, Bouillaud F (2000) The uncoupling protein homologues: UCP1, UCP2, UCP3, StUCP and AtUCP. Biochem J 345: 161–79.

Riddle NC, Richards EJ (2002) The control of natural variation in cytosine methylation in Arabidopsis. Genetics 162: 355–63.

Riddle NC, Richards EJ (2005) Genetic variation in epigenetic inheritance of ribosomal RNA gene methylation in Arabidopsis. Plant J 41: 524–32.

Ridley M (2003) Evolution. 3rd edn. Blackwell Publishing.

Ries G,Heller W, Puchta H, Sandermann H, Seidlitz HK, Hohn B (2000) Elevated UV-B radiation reduces genome stability in plants. Nature 406: 98–101.

Riesch R, Schlupp I, Plath M (2008) Female sperm limitation in natural populations of a sexual/asexual mating complex (Poecilia latipinna, Poecilia formosa). Biol Lett 4: 266–9.

Riesgo A, Maldonado M, Durfort M (2008) Occurrence of somatic cells within the spermatic cysts of demosponges: a discussion of their role. Tissue Cell 40: 387-96.

Rigby MC, Jokela J (2000) Predator avoidance and immune defence: costs and tradeoffs in snails. Proc R Soc Lond B 267: 171-6.

Riggs AD (1975) X inactivation, differentiation, and DNA methylation. Cytogenet Cell Genet 14: 9–25.

Riggs AD, Xiong Z (2004) Methylation and epigenetic fidelity. Proc Natl Acad Sci USA 101: 4-5.

Rigney LP (1995) Postfertilization causes of differential success of pollen donors in Erythronium grandiflorum (Liliaceae): nonrandom ovule abortion. Am J Bot 82: 578–84.

Rigoutsos I, Huynh T, Miranda K, Tsirigos A, McHardy A, Platt D (2006) Short blocks from the noncoding parts of the human genome have instances within nearly all known genes and relate to biological processes. Proc Natl Acad Sci USA 103: 6605–10.

Riley GM (1937) Experimental studies on spermatogenesis in the house sparrow, Passer domesticus (Linnaeus). Anat Rec 67: 327-51.

Riley GM (1940) Light versus activity as a regulator of the sexual cycle in the house sparrow. Wilson Bull 52: 73-86.

Riley JCM, Behrman HR (1991) Oxygen radicals and reactive oxygen species in reproduction. Proc Soc Exp Biol Med 5: 781-91.

Riley T, Sontag E, Chen P, Levine A (2008) Transcriptional control of human p53-regulated genes. Nat Rev Mol Cell Biol 9: 402-12.

Rimon G, Bazenet CE, Philpott KL, Rubin LL (1997) Increased surface phosphatidylserine is an early marker of neuronal apoptosis. J Neurosci Res 48: 563-70.

Rink H, Partke HJ (1975) Giant cells from Saccharomyces uvarum grown after x-irradiation. Radiat Environ Biophys 12: 119-25.

Rinn JL, Chang HY (2012) Genome regulation by long noncoding RNAs. Annu Rev Biochem 81: 145–66.

Ripa J, Olofsson H, Jonzén N (2010) What is bet-hedging, really? Proc Biol Sci 277: 1153-4.

Ripley LS (1982) Model for the participation of quasi-palindromic DNA sequences in frameshift mutation. Proc Natl Acad Sci USA 79: 4128–32.

Risch N, Reich EW, Wishnick MM, McCarthy JG (1987) Spontaneous mutation and parental age in humans. Am J Hum Genet 41:218-48.

Risch TS, Dobson FS, Murie JO (1995) Is mean litter size the most productive? A test in Columbian ground squirrels. Ecology 76:1643-54.

Rispe C, Pierre JS (1998) Coexistence between cyclical parthenogens, obligate parthenogens, and intermediates in a fluctuating environment. J Theor Biol 195: 97-110.

Rispe C, Pierre JS, Simon JC, Gouyon PH (1998) Models of sexual and asexual coexistence in aphids based on constraints. J Evol Biol 11: 685–701.

Rivero A, Balloux F, West SA (2003) Testing for epistasis between deleterious mutations in a parasitoid wasp. Evolution 57: 1698–703.

Rivest S, Rivier C (1995) The role of corticotropin-releasing factor and interleukin-1 in the regulation of neurons controlling reproductive functions. Endocr Rev 16: 177-99.

Rivier C (2002) Inhibitory effect of neurogenic and immune stressors on testosterone secretion in rats. Neuroimmunomodulation 10: 17-29.

Rivier C, Rivier J, Vale W (1986) Stress-induced inhibition of reproductive functions: role of endogenous corticotropin-releasing factor. Science 231: 607-9.

Rivier C, Rivest S (1991) Effect of stress on the activity of the hypothalamic-pituitary-gonadal axis: peripheral and central mechanisms. Biol Reprod 45: 523-32.

Rivlin J, Mendel J, Rubinstein S, Etkovitz N, Breitbart H (2004) Role of hydrogen peroxide in sperm capacitation and acrosome reaction. Biol Reprod 70: 518-22.

Rizzo A, Roscino M, Binetti F, Sciorsci R (2012) Roles of reactive oxygen species in female reproduction. Reprod Domest Anim 47: 344-52.

Ro S, Park C, Sanders KM, McCarrey JR, Yan W (2007a) Cloning and expression profiling of testis-expressed microRNAs. Dev Biol 311: 592–602.

Ro S, Park C, Song R, Nguyen D, Jin J, et al. (2007b) Cloning and expression profiling of testis-expressed piRNA-like RNAs. RNA 13: 1693-702.

Roach JC, Glusman G, Smit AF, Huff CD, Hubley R, et al. (2010) Analysis of genetic inheritance in a family quartet by whole-genome sequencing. Science 328: 636–9.

Robert KA, Bronikowski AM (2010) Evolution of senescence in nature: physiological evolution in populations of garter snake with divergent life histories. Am Nat 175: E47–E159.

Roberts SC, Gosling LM (2003) Genetic similarity and quality interact in mate choice decisions by female mice. Nat Genet 35: 103–6.

Roberts SC, Hale ML, Petrie M (2006) Correlations between heterozygosity and measures of genetic similarity: implications for understanding mate choice. J Evol Biol 19: 558–69.

Robertson DS (1984) Different frequency in the recovery of crossover products from male and female gametes of plants hypoploid for B-A translocations in maize. Genetics 107: 117-30.

Robertson HM, Lampe DJ (1995) Recent horizontal transfer of a mariner transposable element among and between Diptera and Neuroptera. Mol Biol Evol 12: 850–62.

Robichaud NF, Sassine J, Beaton MJ, Lloyd VK (2012) The epigenetic repertoire of Daphnia magna includes modified histones. Genet Res Int 2012: 174860.

Robinson BW, Wilson DS (1996) Genetic variation and phenotypic plasticity in a trophically polymorphic population of pumpkinseed sunfish (Lepomis gibbosus). Evol Ecol 10: 631–52.

Robinson BW, Dukas R (1999) The influence of phenotypic modifications on evolution: the Baldwin effect and modern perspectives. Oikos 85: 582–9.

Robinson MR, Pilkington JG, Clutton-Brock TH, Pemberton JM, Kruuk LEB (2008) Environmental heterogeneity generates fluctuating selection on a secondary sexual trait. Curr Biol 18: 751–7.

Robinson MT, Weeks AR, Hoffmann AA (2002) Geographic patterns of clonal diversity in the earth mite species Penthaleus major with particular emphasis on species margins. Evolution 56: 1160–7.

Robinson R, Fritz IB (1981) Metabolism of glucose by Sertoli cells in culture. Biol Reprod 24: 1032–41.

Robleto EA, Yasbin R, Ross C, Pedraza-Reyes M (2007) Stationary phase mutagenesis in B. subtilis: a paradigm to study genetic diversity programs in cells under stress. Crit Rev Biochem Mol Biol 42: 327-39.

Robson AJ, Bergstrom CT, Pritchard JK (1999) Risky business: sexual and asexual reproduction in variable environments. J Theor Biol 197: 541-56.

Rocha EPC (2006) The quest for the universals of protein evolution. Trends Genet 22: 412-6.

Rocha EPC, Matic I, Taddei F (2002) Over-representation of repeats in stress response genes: a strategy to increase versatility under stressful conditions? Nucleic Acids Res 30: 1886-94.

Rocha EPC, Danchin A (2004) An analysis of determinants of amino acids substitution rates in bacterial proteins. Mol Biol Evol 21: 108–16.

Roche E, Prentki M (1994) Calcium regulation of immediate-early response genes. Cell Calcium 16: 331-8.

Rocheleau AF, Houle G (2001) Different cost of reproduction for the males and females of the rare dioecious shrub Corema conradii (Empetraceae) Am J Bot 88: 659-66.

Rochester DE, WinerJA, Shah DM (1986) The structure and expression of maize genes encoding the major heat shock protein, hsp70. EMBO J 5: 451-8.

Rockett JC, Mapp FL, Garges JB, Luft JC, Mori C, Dix DJ (2001) Effects of hyperthermia on spermatogenesis, apoptosis, gene expression, and fertility in adult male mice. Biol Reprod 65: 229–39.

Rodrigues AB, Zoranovic T, Ayala-Camargo A, Grewal S, Reyes-Robles T, et al. (2012) Activated STAT regulates growth and induces competitive interactions independently of Myc, Yorkie, Wingless and ribosome biogenesis. Development 139: 4051-61.

Rodrigues AS, Lacerda B, Moreno AJ, Ramalho-Santos J (2010) Proton leak modulation in testicular mitochondria affects reactive oxygen species production and lipid peroxidation. Cell Biochem Funct 28: 224-31.

Rodriguez A, Vigorito E, Clare S, Warren MV, Couttet P, et al. (2007) Requirement of bic/microRNA-155 for normal immune function. Science 316: 608–11.

Rodriguez AM, Carrico PM, Mazurkiewicz JE, Melendez JA (2000) Mitochondrial or cytosolic catalase reverses the MnSOD-dependent inhibition of proliferation by enhancing respiratory chain activity, net ATP production, and decreasing the steady state levels of H2O2. Free Radic Biol Med 29: 801–13.

Rodriguez C, Mayo JC, Sainz RM, Antolin I, Herrera F, et al. (2004) Regulation of antioxidant enzymes: a significant role for melatonin. J Pineal Res 36: 1–9.

Rodriguez H, Holmquist GP, D’Agostino R Jr, Keller J, Akman SA (1997) Metal ion-dependent hydrogen peroxide-induced DNA damage is more sequence specific than metal specific. Cancer Res 57: 2394-2403

Rodriguez H, Valentine MR, Holmquist GP, Akman SA, Termini J (1999) Mapping of peroxyl radical induced damage on genomic DNA. Biochemistry 38: 16578-88.

Rodríguez I, Ody C, Araki K, Garcia I, Vassalli P (1997) An early and massive wave of germinal cell apoptosis is required for the development of functional spermatogenesis. EMBO J 16: 2262-70.

Roeder GS, Bailis JM (2000) The pachytene checkpoint. Trends Genet 16: 395-403.

Roff DA (1992) The evolution of life histories. New York, NY: Chapman & Hall.

Roff DA (1996) The evolution of threshold traits in animals. Q Rev Biol 71: 3–36.

Roff D (1997) Evolutionary Quantitative Genetics. New York, NY: Chapman & Hall.

Roff D (2007) Evolutionary Quantitative Genetics. New York, NY: Chapman & Hall.

Roff DA (1998) The maintenance of phenotypic and genetic variation in threshold traits by frequency-dependent selection. J Evol Biol 11: 513–29.

Roff DA, Heibo E, Vøllestad LA (2006) The importance of growth and mortality costs in the evolution of the optimal life history. J Evol Biol 19: 1920-30.

Roff DA, Fairbairn DJ (2007) The evolution of trade-offs: Where are we? J Evol Biol 20: 433–47.

Rogakou EP, Pilch DR, Orr AH, Ivanova VS, Bonner WM (1998) DNA double-stranded breaks induce histone H2AX phosphorylation on serine 139. J Biol Chem 273: 5858–68.

Roger L, Gadea G, Roux P (2006) Control of cell migration: a tumour suppressor function for p53? Biol Cell 98: 141–152.

Rogers AR (1992) Resources and population dynamics. In: Smith EA, Winterhalder B, eds. Evolutionary ecology and human behavior. Hawthorne, NY: Aldine de Gruyter. pp 375-402.

Rohde K (1978) Latitudinal gradients in species diversity and their causes. I. A review of the hypotheses explaining the gradients. Biol Zentralbl 97: 393–403.

Rohde K (1992) Latitudinal gradients in species diversity: the search for the primary cause. Oikos 65: 514–27.

Rohde M, Daugaard M, Jensen MH, Helin K, Nylandsted J, Jäättelä M (2005) Members of the heat-shock protein 70 family promote cancer cell growth by distinct mechanisms. Genes Dev 19: 570-82.

Rohmer C, David JR, Moreteau B, Joly D (2004) Heat induced male sterility in Drosophila melanogaster: adaptive genetic variations among geographic populations and role of the Y chromosome. J Exp Biol 207: 2735–43.

Rokas A (2008) The origins of multicellularity and the early history of the genetic toolkit for animal development. Annu Rev Genet 42: 235–51.

Rokas A, Ladoukakis ED, Zouros E (2003) Animal mitochondrial DNA recombination revisited. Trends Ecol Evol 18: 411–7.

Rolfe I, Banks G (1986) Induction of yeast Ty element transcription by ultraviolet light. Nature 319: 339–40.

Rollo CD (2007) Multidisciplinary aspects of regulatory systems relevant to multiple stressors: aging, xenobiotics, and radiation. In: Mothersill C, Mosse I, Seymour C, eds. Multiple Stressors: a Challenge for the Future. New York, NY: Springer. pp 185–224.

Romanienko PJ, Camerini-Otero RD (1999) Cloning, characterization, and localization of mouse and human SPO11. Genomics 61: 156-69.

Romanienko PJ, Camerini-Otero RD (2000) The mouse Spo11 gene is required for meiotic chromosome synapsis. Mol Cell 6: 975–87.

Romanish MT, Lock WM, van de Lagemaat LN, Dunn CA, Mager DL (2007) Repeated recruitment of LTR retrotransposons as promoters by the antiapoptotic locus NAIP during mammalian evolution. PLoS Genet 3: e10.

Romer AS (1966) Vertebrate paleontology. Chicago, IL: University of Chicago Press.

Romero LM, Dickens MJ, Cyr NE (2009) The reactive scope model: a new model integrating homeostasis, allostasis, and stress. Horm Behav 55: 375–89.

Romero Y, Meikar O, Papaioannou MD, Conne B, Grey C, et al. (2011) Dicer1 depletion in male germ cells leads to infertility due to cumulative meiotic and spermiogenic defects. PLoS One 6: e25241.

Romiguier J, Ranwez V, Douzery EJP, Galtier N (2010) Contrasting GC-content dynamics across 33 mammalian genomes: Relationship with life-history traits and chromosome sizes. Genome Res 20: 1001–9.

Romme WH, Turner MG, Gardner RH, Hargrove WW, Tuskan GA, et al. (1997) A rare episode of sexual reproduction in aspen (Populus tremuloides Michx.) following the 1988 Yellowstone fires. Nat Area J 17:17–25.

Rommel SA, Pabst DA, McLellan WA, Mead JG, Potter CW (1992) Anatomical evidence for a countercurrent heat exchanger associated with dolphin testes. Anat Rec 232: 150–6.

Rommel SA, Early GA, Matassa KA, Pabst DA, McLellan WA (1995) Venous structures associated with thermoregulation of phocid seal reproductive organs. Anat Rec 243: 390–402.

Rommel SA, Pabst DA, McLellan WA (1998) Reproductive thermoregulation in marine mammals. Am Sci 86:440-50.

Rommel SA, Pabst DA, McLellan WA (2001) Functional morphology of venous structures associated with the male and female reproductive systems in Florida manatees (Trichechus manatus latirostris). Anat Rec 264: 339-47.

Romslo I (1975) Energy-dependent accumulation of iron by isolated rat liver mitochondria. IV. Relationship to the energy state of the mitochondria. Biochim Biophys Acta 387: 69-79.

Roopin M, Levy O (2012) Melatonin distribution reveals clues to its biological significance in basal metazoans. PLoS ONE 7: e52266.

Roopnarine PD (2003) Analysis of rates of morphologic evolution. Annu Rev Ecol Evol Syst 34: 605–32.

Roosen-Runge EC (1953) Postmortem mitotic activity of spermatogonia and spermatocytes in the albino rat. Exp Cell Res 4: 52-9.

Roosen-Runge EC (1973) Germinal-cell loss in normal metazoan spermatogenesis. J Reprod Fertil 35: 339–48.

Roosen-Runge EC (1977) The Process of Spermatogenesis in Animals. Cambridge, UK: Cambridge University Press.

Roossinck MJ, Schneider WL (2006) Mutant clouds and occupation of sequence space in plant RNA viruses. Curr Top Microbiol Immunol 299: 337–48.

Root T (1988) Energy constraints on avian distributions and abundances. Ecology 69: 330–9.

Roqueta-Rivera M, Stroud CK, Haschek WM, Akare SJ, Segre M, et al. (2010) Docosahexaenoic acid supplementation fully restores fertility and spermatogenesis in male delta-6 desaturase-null mice. J Lipid Res 51: 360-7.

Rosa CE, Figueiredo MA, Lanes CF, Almeida DV, Monserrat JM, Marins LF (2008) Metabolic rate and reactive oxygen species production in different genotypes of GH-transgenic zebrafish. Comp Biochem Physiol B Biochem Mol Biol 149: 209-14.

Rosario MO, Perkins SL, O'Brien DA, Allen RL, Eddy EM (1992) Identification of the gene for the developmentally expressed 70 kDa heat-shock protein (P70) of mouse spermatogenic cells. Dev Biol 150: 1-11.

Rosas A, Gordo I, Campos PRA (2005) Scaling, genetic drift, and clonal interference in the extinction pattern of asexual population. Phys Rev E 72: 012901.

Rosche WA, Foster PL (1999) The role of transient hypermutators in adaptive mutation in Escherichia coli. Proc Natl Acad Sci USA 96: 6862–7.

Rosche WA, Foster PL, Cairns J (1999) The role of transient hypermutators in adaptive mutation in Escherichia coli. Proc Natl Acad Sci USA 96: 6862–7.

Rose-Hellekant TA, Libersky-Williamson EA, Bavister BD (1998) Energy substrates and amino acids provided during in vitro maturation of bovine oocytes alter acquisition of developmental competence. Zygote 6: 285–94.

Roselius K, Stephan W, Städler T (2005) The relationship of nucleotide polymorphism, recombination rate and selection in wild tomato species. Genetics 171: 753–63.

Rosenberg SM (1994) In pursuit of a molecular mechanism for adaptive mutation. Genome 37: 893-9.

Rosenberg SM (2011) Stress-induced loss of heterozygosity in Candida: a possible missing link in the ability to evolve. mBio2(5). pii: e00200-11.

Rosenberg SM, Thulin C, Harris RS (1998) Transient and heritable mutators in adaptive evolution in the lab and in nature. Genetics 148: 1559–66.

Rosenberg SM, Hastings PJ (2003) Microbiology and evolution. Modulating mutation rates in the wild. Science 300: 1382-3.

Rosenberg SM, Hastings PJ (2004) Genomes: worming into genetic instability. Nature 430: 625–6.

Rosenbluth R, Cuddeford C, Baillie DL (1983) Mutagenesis in Caenorhabditis elegans. I. A rapid eukaryotic mutagen test system using the reciprocal translocation eT1 (III;V). Mutat Res 110: 39–48.

Rosendaal FR, Brocker-Vriends AH, van Houwelingen JC, Smit C, Varekamp I, et al. (1990) Sex ratio of the mutation frequencies in haemophilia A: estimation and meta-analysis. Hum Genet 86: 139-46.

Rosenzweig ML (1978) Competitive speciation. Biol J Linn Soc 10: 275-89.

Rosenzweig RF, Sharp RR, Treves DS, Adams J (1994) Microbial evolution in a simple unstructured environment: genetic differentiation in Escherichia coli. Genetics 137: 903-17.

Rosner A, Moiseeva E, Rinkevich Y, Lapidot Z, Rinkevich B (2009) Vasa and the germ line lineage in a colonial urochordate. Dev Biol 331: 113-28.

Ross AJ, Waymire KG, Moss JE, Parlow AF, Skinner MK, et al. (1998) Testicular degeneration in Bclw-deficient mice. Nat Genet 18: 251–6.

Ross AJ, Li M, Yu B, Gao MX, Derry WB (2011) The EEL-1 ubiquitin ligase promotes DNA damage-induced germ cell apoptosis in C. elegans. Cell Death Differ 18: 1140-9.

Ross C, Pybus C, Pedraza-Reyes M, Sung HM, Yasbin RE, Robleto E (2006) Novel role of mfd: effects on stationary-phase mutagenesis in Bacillus subtilis. J Bacteriol 188: 7512–20.

Ross JA, Watson NE, Jarow JP (1994) The effect of varicoceles upon testicular blood flow in man. Urology 44: 535–9.

Rossetti V, Schirrmeister BE, Bernasconi M, Bagheri H (2010) The evolutionary path to terminal differentiation and division of labor in cyanobacteria. J Theor Biol 262: 23–34.

Rossi V, Schön I, Butlin RK, Menozzi P (1998) Clonal genetic diversity. In: Martens K, ed. Sex and Parthenogenesis: Evolutionary ecology of reproductive modes in non-marine ostracods. Leiden, The Netherlands: Backhuys Publishers. pp 257–74.

Rossi V, Gandolfi A, Gentile G, Geiger W, Menozzi P (2004) Low genetic variability in the ancient asexual ostracod Darwinula stevensoni. Ital J Zool 71: 135-42.

Rossi V, Menozzi P (2012) Effects of mother presence and photoperiod on egg production and hatching of two asexual lineages of Eucypris virens (Crustacea: Ostracoda). Fundam Appl Limnol 180: 59-68.

Rosty C, Ueki T, Argani P, Jansen M, Yeo CJ, et al. (2002) Overexpression of S100A4 in pancreatic ductal adenocarcinomas is associated with poor differentiation and DNA hypomethylation. Am J Pathol 160: 45–50.

Roth GS, Lane MA, Ingram DK (2005) Caloric restriction mimetics: The next phase. Ann NY Acad Sci 1057: 365-71.

Roth JR (2011) The joys and terrors of fast adaptation: new findings elucidate antibiotic resistance and natural selection. Mol Microbiol 79: 279-82.

Roth JR, Andersson DI (2004) Amplification-mutagenesis--how growth under selection contributes to the origin of genetic diversity and explains the phenomenon of adaptive mutation. Res Microbiol 155: 342-51.

Roth JR, Kugelberg E, Reams AB, Kofoid E, Andersson DI (2006) Origin of mutations under selection: the adaptive mutation controversy. Annu Rev Microbiol 60: 477–501.

Roth LM (1974) Reproductive potential of bisexual Pycnoscelus indicus and clones of its parthenogenetic relative Pycnoscelus surinamensis. Ann Entomol Soc Am 67: 215–23.

Rothstein JD, Bristol LA, Hosler B, Brown RH, Kuncl RW (1994) Chronic inhibition of superoxide dismutase produces apoptotic death of spinal neurons. Proc Natl Acad Sci USA 91: 4155-9.

Rotter V, Schwartz D, Almon E, Goldfinger N, Kapon A, et al. (1993) Mice with reduced levels of p53 protein exhibit the testicular giant-cell degenerative syndrome. Proc Natl Acad Sci USA 90: 9075–9.

Rouet P, Smih F, Jasin M (1994) Expression of a site-specific endonuclease stimulates homologous recombination in mammalian cells. Proc Natl Acad Sci USA 91: 6064–8.

Rouget C, Papin C, Boureux A, Meunier AC, Franco B, et al. (2010) Maternal mRNA deadenylation and decay by the piRNA pathway in the early Drosophila embryo. Nature 467: 1128-32.

Roughgarden J (1971) Density-dependent natural selection. Ecology 52: 453-68.

Roughgarden J (1972) Evolution of niche width. Am Nat 106: 683-718.

Roughgarden J (1979) Theory of Population Genetics and Evolutionary Ecology: an Introduction. New York, NY: Macmillan.

Roughgarden J (1991) The evolution of sex. Am Nat 138: 823-42.

Roush RT, McKenzie JA (1987) Ecological genetics of insecticide and acaricide resistance. Annu Rev Entomol 32: 361–80.

Rousseau F, Bonaventure J, Legeai-Mallet L, Pelet A, Rozet JM, et al. (1996) Mutations of the fibroblast growth factor receptor-3 gene in achondroplasia. Horm Res 45: 108–10.

Roux W (1881) Der Kampf der Theile im Organismus. Leipzig, Germany: W. Engelmann.

Rouzine IM, Wakeley J, Coffin JM (2003) The solitary wave of asexual evolution. Proc Natl Acad Sci USA 100: 587–92.

Rouzine IM, Brunet É, Wilke CO (2008) The traveling wave approach to asexual evolution: Muller’s ratchet and speed of adaptation. Theor Popul Biol 73: 24–46.

Rowe G, Beebee TJ (2003) Population on the verge of a mutational meltdown? Fitness costs of genetic load for an amphibian in the wild. Evolution Int J Org Evolution 57: 177-81.

Rowe GT, Manson NH, Caplan M, Hess ML (1983) Hydrogen peroxide and hydroxyl radical mediation of activated leukocyte depression of cardiac sarcoplasmic reticulum: participation of the cyclooxygenase pathway. Circ Res 59: 612-9.

Rowe L, Houle D (1996) The lek paradox and the capture of genetic variance by condition dependent traits. Proc R Soc Lond B 263: 1415-21.

Rowe M, Pruett-Jones S (2011) Sperm competition selects for sperm quantity and quality in the Australian Maluridae. PLoS ONE 6: e15720.

Roy AK, Vellanoweth RL, Chen S, Supakar PC, Jung MH, et al. (1996) The evolutionary tangle of aging, sex, and reproduction and an experimental approach to its molecular dissection. Exp Gerontol 31: 83-94.

Roy BA, Kirchner JW (2000) Evolutionary dynamics of pathogen resistance and tolerance. Evolution 54: 51–63.

Roy K, Jablonski D, Valentine JW (2004) Beyond species richness: Biogeographic patterns and biodiversity dynamics using other metrics of diversity. In: Lomolino MV, Heaney LR, eds. Frontiers of biogeography: new directions in the geography of nature. Sunderland, MA: Sinauer. pp 151-170.

Royere D, Guérif F, Laurent-Cadoret V, Hochereau de Reviers MT (2004) Apoptosis in testicular germ cells. Int Congr Ser 1266: 170-6.

Royle NJ, Surai PF, Hartley IR (2001) Maternally derived androgens and antioxidants in bird eggs: complementary but opposing effects? Behav Ecol 12: 381–5.

Roze D (2012) Disentangling the benefits of sex. PLoS Biol 10: e1001321.

Rozen DE, de Visser JA, Gerrish PJ (2002) Fitness effects of fixed beneficial mutations in microbial populations. Curr Biol 12: 1040–5.

Rozen DE, McGee L, Levin BR, Klugman KP (2007) Fitness costs of fluoroquinolone resistance in Streptococcus pneumoniae. Antimicrob Agents Chemother 51: 412–6.

Rozen S, Skaletsky H, Marszalek JD, Minx PJ, Cordum HS, et al. (2003) Abundant gene conversion between arms of palindromes in human and ape Y chromosomes. Nature 423: 873–6.

Rozenboim I, Aharony T, Yahav S (2002) The effect of melatonin administration on circulating plasma luteinizing hormone concentration in castrated White Leghorn roosters. Poult Sci 81: 1354–9.

Rübe CE, Zhang S, Miebach N, Fricke A, Rübe C (2011) Protecting the heritable genome: DNA damage response mechanisms in spermatogonial stem cells. DNA Repair (Amst) 10: 159-68.

Rubolini D, Romano M, Boncoraglio G, Ferrari RP, Martinelli R, et al. (2005) Effects of elevated egg corticosterone levels on behavior, growth, and immunity of yellow-legged gull (Larus michahellis) chicks. Horm Behav 47: 592–605.

Ruden DM, Garfinkel MD, Sollars VE, Lu X (2003) Waddington's widget: Hsp90 and the inheritance of acquired characters. Semin Cell Dev Biol 14: 301–10.

Ruden DM, Xiao L, Garfinkel MD, Lu X (2005) Hsp90 and environmental impacts on epigenetic states: a model for the trans-generational effects of diethylstibesterol on uterine development and cancer. Hum Mol Genet 14 Spec No 1: R149-55.

Ruder EH, Hartman TJ, Blumberg J, Goldman MB (2008) Oxidative stress and antioxidants: exposure and impact on female fertility. Hum Reprod Update 14: 345-57.

Rudin N, Haber JE (1988) Efficient repair of HO-induced chromosomal breaks in Saccharomyces cerevisiae by recombination between flanking homologous sequences. Mol Cell Biol 8: 3918–28.

Rudolfsson SH, Wilkström P, Jonsson A, Collin O, Bergh A (2004) Hormonal regulation and functional role of vascular endothelial growth factor A in the rat testis. Biol Reprod 70: 340–47.

Rudra D, Warner JR (2004) What better measure than ribosome synthesis? Genes Dev 18: 2431–6.

Rueffler C, Van Dooren TJ, Leimar O, Abrams PA (2006) Disruptive selection and then what? Trends Ecol Evol 21: 238-45.

Rüegger S, Großhans H (2012) MicroRNA turnover: when, how, and why. Trends Biochem Sci 37: 436-46.

Ruffié J (1986) Le sexe et la mort. Paris, France: Éditions Odile Jacob.

Rugo RE, Mutamba JT, Mohan KN, Yee T, Chaillet JR, et al. (2011) Methyltransferases mediate cell memory of a genotoxic insult. Oncogene 30: 751–6.

Ruiz-Laguna J, Prieto-Alamo MJ, Pueyo C (2000) Oxidative mutagenesis in Escherichia coli strains lacking ROS-scavenging enzymes and/or 8-oxoguanine defenses. Environ Mol Mutagen 35: 22-30.

Ruiz-Pesini E, Lapena AC, Diez C, Alvarez E, Enriquez JA, Lopez-Perez MJ (2000a) Seminal quality correlates with mitochondrial functionality. Clin Chim Acta 300: 97–105.

Ruiz-Pesini E, Lapena AC, Diez-Sanchez C, et al. (2000b) Human mtDNA haplogroups associated with high or reduced spermatozoa motility. Am J Hum Genet 67: 682–96.

Ruiz-Pesini E, Mishmar D, Brandon M, Procaccio V, Wallace DC (2004) Effects of purifying and adaptive selection on regional variation in human mtDNA. Science 303: 223-6.

Ruiz-Pesini E, Wallace DC (2006) Evidence for adaptive selection acting on the tRNA and rRNA genes of human mitochondrial DNA. Hum Mutat 27: 1072-81.

Rundle HD, Chenoweth SF, Blows MW (2006) The roles of natural and sexual selection during adaptation to a novel environment. Evolution 60: 2218–25.

Runnegar B (1986) Molecular palaeontology. Palaeontology 29: 1–24.

Runnegar B (1991) Precambrian oxygen levels estimated from the biochemistry and physiology of early eukaryotes. Global Planet Change 5: 97–111.

Runyan C, Schaible K, Molyneaux K, Wang Z, Levin L, Wylie C (2006) Steel factor controls midline cell death of primordial germ cells and is essential for their normal proliferation and migration. Development 133: 4861–9.

Ruse M (2003) Darwin and Design: Does Evolution have a Purpose? Cambridge, MA: Harvard University Press.

Ruse M (2009) The history of evolutionary thought. In: Ruse M, Travis J, eds. Evolution. The first four billion years. Cambridge, MA: Belknap Press. pp 1-48.

Rush B, Sandver S, Bruer J, Roche R, Wells M, Giebultowicz J (2007) Mating increases starvation resistance and decreases oxidative stress resistance in Drosophila melanogaster females. Aging Cell 6: 723–6.

Rush M, Appanah R, Lee S, Lam LL, Goyal P, Lorincz MC (2009) Targeting of EZH2 to a defined genomic site is sufficient for recruitment of Dnmt3a but not de novo DNA methylation. Epigenetics 4: 404-14.

Russell L (1977) Movement of spermatocytes from the basal to the adluminal compartment of the rat testis. Am J Anat 148: 313–28.

Russell LD, Clermont Y (1977) Degeneration of germ cells in normal, hypophysectomized and hormone treated hypophysectomized rats. Anat Rec 187: 347–66.

Russell L, Frank B (1978) Ultrastructural characterization of nuage in spermatocytes of the rat testis. Anat Rec 190: 79–97.

Russell LD, Peterson RN (1984) Determination of the elongate spermatid-Sertoli cell ratio in various mammals. J Reprod Fertil 70: 635-41.

Russell LD, Peterson RN (1985) Sertoli cell junctions: morphological and functional correlates. Int Rev Cytol 94: 177–211.

Russell LD, Griswold MD (1993) The Sertoli Cell. Clearwater, FL: Cache River Press.

Rusyn I, Denissenko MF, Wong VA, Butterworth BE, Cunningham ML, et al. (2000) Expression of base excision repair enzymes in rat and mouse liver is induced by peroxisome proliferators and is dependent upon carcinogenic potency. Carcinogenesis 21: 2141–5.

Rusyn I, Asakura S, Pachkowski B, Bradford BU, Denissenko MF, et al. (2004) Expression of base excision DNA repair genes is a sensitive biomarker for in vivo detection of chemical-induced chronic oxidative stress: identification of the molecular source of radicals responsible for DNA damage by peroxisome proliferators. Cancer Res 64: 1050-7.

Rutherford SL (2003) Between genotype and phenotype: protein chaperones and evolvability. Nat Rev Genet 4: 263–74.

Rutherford SL, Zuker CS (1994) Protein folding and the regulation of signaling pathways. Cell 79: 1129–32.

Rutherford SL, Lindquist S (1998) Hsp90 as a capacitor for morphological evolution. Nature 396: 336–42.

Rutherford SL, Henikoff S (2003) Quantitative epigenetics. Nat Genet 33: 6-8.

Rutherford S, Knapp JR, Csermely P (2007a) Hsp90 and developmental networks. In: Csermely P, Vigh L, eds. Molecular aspects of the stress response: chaperones, membranes and networks. London, UK: Landes Bioscience. pp 190-198.

Rutherford S, Hirate Y, Swalla BJ (2007b) The Hsp90 capacitor, developmental remodeling, and evolution: the robustness of gene networks and the curious evolvability of metamorphosis. Crit Rev Biochem Mol Biol 42: 355-72.

Rutkowska J, Cichon M, Puerta M, Gil D (2005) Negative effects of elevated testosterone on female fecundity in zebra finches. Horm Behav 47: 585–91.

Rutter GA, Rizzuto R (2000) Regulation of mitochondrial metabolism by ER Ca2+ release: an intimate connection. Trends Biol Sci 25: 215-21.

Ryan FJ (1959) Bacterial mutation in stationary phase and the question of cell turnover. J Gen Microbiol 21: 530–49.

Ryan FJ, Nakada D, Schneider MJ (1961) Is DNA replication a necessary condition for spontaneous mutation? Z Vererbungsl 92:38–41.

Ryan FJ, Okada T, Nagata T (1963) Spontaneous mutation in spheroplasts of Escherichia coli. J Gen Microbiol 30: 193–9.

Ryan HE, Lo J, Johnson RS (1998) HIF-1 alpha is required for solid tumor formation and embryonic vascularization. EMBO J 17: 3005–15.

Ryan MJ, Dries LA, Batra P, Hillis DM (1996) Male mate preferences in a gynogenetic species complex of Amazon mollies. Anim Behav 52: 1225-36.

Ryter SW, Alam J, Choi AM (2006) Heme oxygenase-1/carbon monoxide: from basic science to therapeutic applications. Physiol Rev 86: 583-650.

Sabeur K, Ball BA (2006) Detection of superoxide anion generation by equine spermatozoa. Am J Vet Res 67: 701–6.

Sabeur K, Ball BA (2007) Characterization of NADPH oxidase 5 in equine testis and spermatozoa. Reproduction 134: 263-70.

Sabin LR, Delás MJ, Hannon GJ (2013) Dogma derailed: the many influences of RNA on the genome. Mol Cell 49: 783-94.

Sablina AA, Budanov AV, Ilyinskaya GV, Agapova LS, Kravchenko JE, Chumakov PM (2005) The antioxidant function of the p53 tumor suppressor. Nat Med 11: 1306-13.

Saccheri IJ, Lloyd HD, Helyar SJ, Brakefield PM (2005) Inbreeding uncovers fundamental differences in the genetic load affecting male and female fertility in a butterfly. Proc Biol Sci 272: 39-46.

Saccone C, Gissi C, Reyes A, Larizza A, Sbisà E, Pesole G (2002) Mitochondrial DNA in metazoa: degree of freedom in a frozen event. Gene 286: 3-12.

Sacconi S, Salviati L, Nishigaki Y, Walker WF, Hernandez-Rosa E, et al. (2008) A functionally dominant mitochondrial DNA mutation. Hum Mol Genet 17: 1814–20.

Sachs JL, Mueller UG, Wilcox TP, Bull JJ (2004) The evolution of cooperation. Q Rev Biol 79: 135–60.

Sadras VO, Denison RF (2009) Do plant parts compete for resources? An evolutionary viewpoint. New Phytol 183: 565-74.

Saffman EE, Lasko P (1999) Germline development in vertebrates and invertebrates. Cell Mol Life Sci 55: 1141-63.

Saga Y, Yanagisawa K (1982) Macrocyst development in Dictyostelium discoideum. I. Induction of synchronous development by giant cells and biochemical analysis. J Cell Sci 55: 341–52.

Sagar M (2010) HIV-1 transmission biology: selection and characteristics of infecting viruses. J Infect Dis 202 Suppl 2: S289-96.

Sagar M, Laeyendecker O, Lee S, Gamiel J, Wawer MJ, et al. (2009) Selection of HIV variants with signature genotypic characteristics during heterosexual transmission. J Infect Dis 199: 580–9.

Sahoo DK, Roy A, Bhanja S, Chainy GB (2005) Experimental hyperthyroidism-induced oxidative stress and impairment of antioxidant defence system in rat testis. Indian J Exp Biol 43: 1058–67.

Sahoo DK, Roy A, Bhanja S, Chainy GB (2007) Hypothyroidism impairs antioxidant defence system and testicular physiology during development and maturation. Gen Comp Endocrinol 156: 63–70.

Sahoo DK, Roy A, Chainy GB (2008) Protective effects of vitamin E and curcumin on L-thyroxine-induced rat testicular oxidative stress. Chem Biol Interact 176: 121–8.

Saidapur SK (1978) Follicular atresia in the ovaries of nonmammalian vertebrates. Int Rev Cytol 54: 225–44.

Saini HS (1997) Effect of water stress on male gametophyte development in plants. Sex Plant Reprod 10: 67–73.

Saino N, Romano M, Ferrari RP, Martinellim R, Møller AP (2005) Stressed mothers lay eggs with high corticosterone levels which produce low quality offspring. J Exp Zool 303A: 998–1006.

Saint-Ruf C, Matic I (2006) Environmental tuning of mutation rates. Environ Microbiol 8: 193–9.

Saito K, Nishida KM, Mori T, Kawamura Y, Miyoshi K, et al. (2006) Specific association of Piwi with rasiRNAs derived from retrotransposon and heterochromatic regions in the Drosophila genome. Genes Dev 20: 2214–22.

Saito K, Sakaguchi Y, Suzuki T, Suzuki T, Siomi H, Siomi MC (2007)Pimet, the Drosophila homolog of HEN1, mediates 2'-O-methylation of Piwi- interacting RNAs at their 3' ends. Genes Dev 21: 1603-8.

Saito K, Siomi MC (2010) Small RNA-mediated quiescence of transposable elements in animals. Dev Cell 19: 687-97.

Saito Y, Kanai Y, Sakamoto M, Saito H, Ishii H, Hirohashi S (2002) Overexpression of a splice variant of DNA methyltransferase 3b, DNMT3b4, associated with DNA hypomethylation on pericentromeric satellite regions during human hepatocarcinogenesis. Proc Natl Acad Sci USA 99: 10060–5.

Saitou M (2009) Germ cell specification in mice. Curr Opin Genet Dev 19: 386-95.

Saitou M, Barton SC, Surani MA (2002) A molecular programme for the specification of germ cell fate in mice. Nature 418: 293–300.

Saitou M, Yamaji M (2010) Germ cell specification in mice: signaling, transcription regulation, and epigenetic consequences. Reproduction 139: 931-42.

Sakai Y, Aminaka M, Takata A, Kudou Y, Yamauchi H, Aizawa Y, Sakagami H (2010) Oxidative stress in mature rat testis and its developmental changes. Dev Growth Differ 52: 657-63.

Sakata T, Higashitani A (2008) Male sterility accompanied with abnormal anther development in plants—genes and environmental stresses with special reference to high temperature injury. Intl J Plant Dev Biol 2:42–51.

Sakkas D, Manicardi G, Bianchi PG, Bizzaro D, Bianchi U (1995) Relationship between the presence of endogenous nicks and sperm chromatin packaging in maturing and fertilizing mouse spermatozoa. Biol Reprod 52: 1149–55.

Sakkas D, Mariethoz E, Manicardi G, Bizzaro D, Bianchi PG, Bianchi U (1999) Origin of DNA damage in ejaculated human spermatozoa. Rev Reprod 4: 31-7.

Sakkas D, Alvarez JG (2010) Sperm DNA fragmentation: mechanisms of origin, impact on reproductive outcome, and analysis. Fertil Steril 93: 1027–36.

Salamon JA, Alphei J, Ruf A, Schaefer M, Scheu S, et al. (2006) Transitory dynamic effects in the soil invertebrate community in a temperate deciduous forest: effects of resource quality. Soil Biol Biochem 38: 209–21.

Salathé M, Kouyos RD, Bonhoeffer S (2008) The state of affairs in the kingdom of the Red Queen. Trends Ecol Evol 23: 439–45.

Salathé P, Ebert D (2003) The effects of parasitism and inbreeding on the competitive ability in Daphnia magna: evidence for synergistic epistasis. J Evol Biol 16: 976-85.

Salathia N, Queitsch C (2007) Molecular mechanisms of canalization: Hsp90 and beyond. J Biosci 32: 457–63.

Salceda S, Caro J (1997) Hypoxia-inducible factor 1α (HIF-1α) protein is rapidly degraded by the ubiquitin-proteasome system under normoxic conditions. Its stabilization by hypoxia depends on redox-induced changes. J Biol Chem 272: 22642–7.

Saleh RA, Agarwal A (2002) Oxidative stress and male infertility: from research bench to clinical practice. J Androl 23:737-52.

Salisbury EJ (1942) The reproductive capacity of plants. London, UK: Bell and Sons.

Salmon AB, Marx DB, Harshman LG (2001) A cost of reproduction in Drosophila melanogaster: stress susceptibility. Evolution 55: 1600-8.

Salnikow K, Zhitkovich A (2008) Genetic and epigenetic mechanisms in metal carcinogenesis and cocarcinogenesis: nickel, arsenic, and chromium. Chem Res Toxicol 21: 28–44.

Salt WR (1954) The structure of the cloacal protuberance of the Vesper Sparrow (Pooecetes gramineus) and certain other passerine birds. Auk 71: 64-73.

Salverda MLM, Dellus E, Gorter FA, Debets AJM, Van der Oost J, et al. (2011) Initial mutations direct alternative pathways of protein evolution. PLoS Genet 7: e1001321.

Salverda MLM, de Visser JAGM (2011) Evolutionary genetics: evolution with foresight. Curr Biol 21: R398-400.

Salvi M, Battaglia V, Brunati AM, La Rocca N, Tibaldi E, et al. (2007) Catalase takes part in rat liver mitochondria oxidative stress defense. J Biol Chem 282: 24407–15.

Samain JF, Dégremont L, Soletchnik P, Haure J, Bédier E, et al. (2007) Genetically based resistance to summer mortality in the Pacific oyster (Crassostrea gigas) and its relationship with physiological, immunological characteristics and infection processes. Aquaculture 268: 227–43.

Samali A, Orrenius S (1998) Heat shock proteins: regulators of stress response and apoptosis. Cell Stress Chaperones 3: 228-36.

Samani P, Bell G (2010) Adaptation of experimental yeast populations to stressful conditions in relation to population size. J Evol Biol 23: 791-6.

Sampayo JN, Gill MS, Lithgow GL (2003) Oxidative stress and aging — the use of superoxide dismutase/catalase mimetics to extend lifespan. Biochem Soc Trans 31: 1305–7.

Samson L, Cairns J (1977) A new pathway for DNA repair in Escherichia coli. Nature 267: 281–3.

Sanchez L, Perondini ALP (1999) Sex determination in sciarid flies: a model for the control of differential X-chromosome elimination. J Theor Biol 197: 247–59.

Sanchez Y, Taulien J, Borkovich KA, Lindquist S (1992) Hsp104 is required for tolerance to many forms of stress. EMBO J 11: 2357–64.

Sánchez-Hidalgo M, de la Lastra CA, Carrascosa-Salmoral MP, Naranjo MC, Gomez-Corvera A, et al. (2009a)Age-related changes in melatonin synthesis in rat extrapineal tissues. Exp Gerontol 44: 328-34.

Sánchez-Hidalgo M, Guerrero Montávez JM, Carrascosa-Salmoral MP, Naranjo Gutierrez MC, Lardone PJ, de la Lastra Romero CA (2009b) Decreased MT1 and MT2 melatonin receptor expression in extrapineal tissues of the rat during physiological aging. J Pineal Res 46: 29-35.

Sandaltzopoulos R, Mitchelmore C, Bonte E, Wall G, Becker PB (1995) Dual regulation of the Drosophila hsp26 promoter in vitro. Nucleic Acids Res 23: 2479-87.

Sanders NJ, Lessard J-P, Fitzpatrick MC, Dunn RR (2007) Temperature, but not productivity or geometry, predicts elevational diversity gradients in ants across spatial grains. Global Ecol Biogeogr 16: 640–9.

Sandhu R, Rivenbark AG, Coleman WB (2012) Loss of post-transcriptional regulation of DNMT3b by microRNAs: a possible molecular mechanism for the hypermethylation defect observed in a subset of breast cancer cell lines. Int J Oncol 41: 721-32.

Sandovici I, Kassovska-Bratinova S, Loredo-Osti JC, Leppert M, Suarez A, et al. (2005) Interindividual variability and parent of origin DNA methylation differences at specific human Alu elements. Hum Mol Genet 14: 2135–43.

Sandovici I, Sapienza C (2010) PRDM9 sticks its zinc fingers into recombination hotspots and between species. F1000 Biol Rep 2: 37.

Sandqvist A, Björk JK, Åkerfelt M, Chitikova Z, Grichine A, et al. (2009) Heterotrimerization of heat-shock factors 1 and 2 provides a transcriptional switch in response to distinct stimuli. Mol Biol Cell 20: 1340-7.

Sandstedt S, Tucker P (2005) Male-driven evolution in closely related species of the mouse genus Mus. J Mol Evol 61: 138–44.

Sangster TA, Queitsch C, Lindquist S (2003) Hsp90 and chromatin: where is the link? Cell Cycle 2: 166-8.

Sangster TA, Lindquist S, Queitsch C (2004) Under cover: causes, effects and implications of Hsp90-mediated genetic capacitance. BioEssays 26: 348–62.

Sangster TA, Bahrami A, Wilczek A, Watanabe E, Schellenberg K, et al. (2007) Phenotypic diversity and altered environmental plasticity in Arabidopsis thaliana with reduced Hsp90 levels. PLoS ONE 2: e648.

Sangster TA, Salathia N, Undurraga S, Milo R, Schellenberg K, et al. (2008) HSP90 affects the expression of genetic variation and developmental stability in quantitative traits. Proc Natl Acad Sci USA 105: 2963-8.

Sanjuán R (2012) From molecular genetics to phylodynamics: evolutionary relevance of mutation rates across viruses. PLoS Pathog 8: e1002685.

Sanjuán R, Moya A, Elena SF (2004a) The distribution of fitness effects caused by single-nucleotide substitutions in an RNA virus. Proc Natl Acad Sci USA 101: 8396-401.

Sanjuán R, Moya A, Elena SF (2004b) The contribution of epistasis to the architecture of fitness in an RNA virus. Proc Natl Acad Sci USA 43: 15376–9.

Sanjuán R, Elena SF (2006) Epistasis correlates to genomic complexity. Proc Natl Acad Sci USA 103: 14402-5.

Sanjuán R, Cuevas JM, Furio V, Holmes EC, Moya A (2007) Selection for robustness in mutagenized RNA viruses. PLoS Genet 3: e93.

Sanjuán R, Nebot MR (2008) A network model for the correlation between epistasis and genomic complexity. PLoS One 3: e2663.

SanMiguel P, Tikhonov A, Jin YK, Motchoulskaia N, Zakharov D, et al. (1996) Nested retrotransposons in the intergenic regions of the maize genome. Science 274: 765-8.

Sansom OJ, Mansergh FC, Evans MJ, Wilkins JA, Clarke AR (2007) Deficiency of SPARC suppresses intestinal tumorigenesis in APCMin/+ mice. Gut 56: 1410–4.

Santangelo AM, de Souza FS, Franchini LF, Bumaschny VF, Low MJ, Rubinstein M (2007) Ancient exaptation of a CORESINE retroposon into a highly conserved mammalian neuronal enhancer of the proopiomelanocortin gene. PLoS Genet 3: 1813–26.

Santos C, Montiel R, Sierra B, Bettencourt C, Fernandez E, et al. (2005) Understanding differences between phylogenetic and pedigree-derived mtDNA mutation rate: a model using families from the Azores Islands (Portugal). Mol Biol Evol 22: 1490–505.

Sapienza C, Peterson AC, Rossant J, Balling R (1987) Degree of methylation of transgenes is dependent on gamete of origin. Nature 328: 251–4.

Sapolsky RM (1985) Stress-induced suppression of testicular function in the wild baboon: Role of glucocorticoids. Endocrinology 116: 2273-8.

Sapolsky R (1992) Stress, the Aging Brain and the Mechanisms of Neuron Death. Cambridge, MA: MIT Press.

Sarabia L, Maurer I, Bustos-Obregón E (2009a) Melatonin prevents damage elicited by the organophosphorous pesticide diazinon on mouse sperm DNA. Ecotoxicol Environ Saf 72: 663-8.

Sarabia L, Maurer I, Bustos-Obregón E (2009b) Melatonin prevents damage elicited by the organophosphorous pesticide diazinon on the mouse testis. Ecotoxicol Environ Saf 72: 938-42.

Saran M, Michel C, Bors W (1998) Radical functions in vivo: a critical review of current concepts and hypotheses. Z Naturforsch [C] 53: 210-27.

Sardina JL, López-Ruano G, Sánchez-Sánchez B, Llanillo M, Hernández-Hernández A (2012) Reactive oxygen species: are they important for haematopoiesis? Crit Rev Oncol Hematol 81: 257-74.

Saretzki G, Armstrong L, Leake A, Lako M, von Zglinicki T (2004) Stress defense in murine embryonic stem cells is superior to that of various differentiated murine cells. Stem Cells 22: 962–71.

Saretzki G, Walter T, Atkinson S, Passos JF, Bareth B, et al. (2008) Downregulation of multiple stress defense mechanisms during differentiation of human embryonic stem cells. Stem Cells 26: 455-64.

Sarge KD (1995) Male germ cell-specific alteration in temperature set-point of the cellular stress response. J Biol Chem 270: 18745-8.

Sarge KD, Park-Sarge O-K, Kirby JD, Mayo KE, Morimoto RI (1994) Expression of heat shock factor 2 in mouse testis: potential role as a regulator of heat-shock protein gene expression during spermatogenesis. Biol Reprod 50: 1334–43.

Sarge KD, Bray AE, Goodson ML (1995) Altered stress response in testis. Nature 374: 126.

Sarge KD, Cullen KE (1997) Regulation of hsp expression during rodent spermatogenesis. Cell Mol Life Sci 53: 191–7.

Sargent RG, Brenneman MA, Wilson JH (1997) Repair of site-specific double-strand breaks in a mammalian chromosome by homologous and illegitimate recombination. Mol Cell Biol 17: 267–77.

Saridaki A, Bourtzis K (2010) Wolbachia: more than just a bug in insects genitals. Curr Opin Microbiol 13: 67-72.

Sarkar B (1995) Metal replacement in DNA-binding zinc finger proteins and its relevance to mutagenicity and carcinogenicity through free radical generation. Nutrition 11(5 Suppl.): 646-9.

Sarot E, Payen-Groschêne G, Bucheton A, Pélisson A (2004) Evidence for a piwi-dependent RNA silencing of the gypsy endogenous retrovirus by the Drosophila melanogaster flamenco gene. Genetics 166: 1313–21.

Sarsour EH, Venkataraman S, Kalen AL, Oberley LW, Goswami PC (2008) Manganese superoxide dismutase activity regulates transitions between quiescent and proliferative growth. Aging Cell 7: 405–17.

Sarsour EH, Kumar MG, Chaudhuri L, Kalen AL, Goswami PC (2009) Redox control of the cell cycle in health and disease. Antioxid Redox Signal 11: 2985-3011.

Sasagawa I, Matsuki S, Suzuki Y, Iuchi Y, Tohya K, et al. (2001a) Possible involvement of the membrane-bound form of peroxiredoxin 4 in acrosome formation during spermiogenesis of rats. Eur J Biochem 268: 3053–61.

Sasagawa I, Yazawa H, Suzuki Y, Nakada T (2001b) Stress and testicular germ cell apoptosis. Arch Androl 47: 211-6.

Sasaki A (1994) Evolution of antigen drift/switching – continuously evading pathogens. J Theor Biol 168: 291-308.

Sasaki A, Ellner S (1995) The evolutionarily stable phenotype distribution in a random environment. Evolution 49: 337–50.

Sasaki H, Matsui Y (2008) Epigenetic events in mammalian germ-cell development: reprogramming and beyond. Nat Rev Genet 9: 129–40.

Sasaki M, Lange J, Keeney S (2010) Genome destabilization by homologous recombination in the germ line. Nat Rev Mol Cell Biol 11: 182-95.

Sasaki T, Tokoro M (1997) Adaptation toward changing environments: why Darwinian in nature. In: Husbands P, Harvey I, eds. Proceedings of the Fourth European Conference on Artificial Life. Cambridge, MA: Bradford/MIT Press. pp 145-153.

Sasaki T, Tokoro M (1999) Evolving learnable neural networks under changing environments with various rates of inheritance of acquired characters: comparison between Darwinian and Lamarckian evolution. Artif Life 5: 203-23.

Sasso-Cerri E, Cerri PS, Freymüller E, Miraglia SM (2006) Apoptosis during the seasonal spermatogenic cycle of Rana catesbeiana. J Anat 209: 21-9.

Sathees CR, Raman MJ (1999) Mouse testicular extracts process DNA double-strand breaks efficiently by DNA end-to-end joining. Mutat Res 433: 1-13.

Sato K, Ito Y, Yomo T, Kaneko K (2003) On the relation between fluctuation and response in biological systems. Proc Natl Acad Sci USA 100: 14086–90.

Sato M, Sato K (2011) Degradation of paternal mitochondria by fertilization-triggered autophagy in C. elegans embryos. Science 334: 1141–4.

Sato N, Fukushima N, Matsubayashi H, Goggins M (2004) Identification of maspin and S100P as novel hypomethylation targets in pancreatic cancer using global gene expression profiling. Oncogene 23: 1531–8.

Sattler M, Winkler T, Verma S, Byrne CH, Shrikhande G, et al. (1999) Hematopoietic growth factors signal through the formation of reactive oxygen species. Blood 93: 2928-35.

Sauer H, Wartenberg M, Hescheler J (2001) Reactive oxygen species as intracellular messengers during cell growth and differentiation. Cell Physiol Biochem 11: 173–86.

Saunders MA, Liang H, Li WH (2007) Human polymorphism at microRNAs and microRNA target sites. Proc Natl Acad Sci USA 104: 3300-5.

Saunders NJ, Jeffries AC, Peden JF, Hood DW, Tettelin H, et al. (2000) Repeat-associated phase variable genes in thecomplete genome sequence of Neisseria meningitidis strain MC58. Mol Microbiol 37: 207-15.

Savic DJ, Kanazir DT (1972) The effect of a histidine operator-constitutive mutation on UV-induced mutability within the histidine operon of Salmonella typhimurium. Mol Gen Genet 118: 45–50.

Savill NJ, Hogeweg P (1998) Spatially induced speciation prevents extinction: the evolution of dispersal distance in oscillatory predator-prey models. Proc R Soc Lond Ser B 265: 25–32.

Savitz DA, Sonnenfeld NL, Olshan AF (1994) Review of epidemiologic studies of paternal occupational exposure and spontaneous abortion. Am J Ind Med 25: 361–83.

Savva D (1982) Spontaneous mutation rates in continuous cultures: the effect of some environmental factors. Microbios 33: 81-92.

Sawaya S, Bagshaw A, Buschiazzo E, Kumar P, Chowdhury S, et al. (2013) Microsatellite tandem repeats are abundant in human promoters and are associated with regulatory elements. PLoS ONE 8: e54710.

Sawhney VK, Shukla A (1994) Male sterility in flowering plants: Are plant growth substances involved? Am J Bot 81: 1640–7.

Sawyer DE, Mercer BG, Wiklendt AM, Aitken RJ (2003) Quantitative analysis of gene-specific DNA damage in human spermatozoa. Mutat Res 529: 21–34.

Sawyer LA, Hennessy JM, Peixoto AA, Rosato E, Parkinson H, et al. (1997) Natural variation in a Drosophila clock gene and temperature compensation. Science 278: 2117–20.

Sawyer SA, Kulathinal RJ, Bustamante CD, Hartl DL (2003) Bayesian analysis suggests that most amino acid replacements in Drosophila are driven by positive selection. J Mol Evol 57 Suppl 1: S154–S164.

Sawyer SA, Parsch J, Zhang Z, Hartl DL (2007) Prevalence of positive selection among nearly neutral amino acid replacements in Drosophila. Proc Natl Acad Sci USA 104: 6504–10.

Saxowsky TT, Doetsch PW (2006) RNA polymerase encounters with DNA damage: transcription-coupled repair or transcriptional mutagenesis? Chem Rev 106: 474–88.

Sayres MAW, Makova KD (2011) Genome analyses substantiate male mutation bias in many species. Bioessays 33: 938–45.

Sbilordo SH, Schäfer MA, Ward PI (2009) Sperm release and use at fertilization by yellow dung fly females (Scathophaga stercoraria). Biol J Linn Soc 98: 511–8.

Scacheri PC, Rozenblatt-Rosen O, Caplen NJ, Wolfsberg TG, Umayam L, et al. (2004) Short interfering RNAs can induce unexpected and divergent changes in the levels of untargeted proteins in mammalian cells. Proc Natl Acad Sci USA 101: 1892–7.

Schaack S, Gilbert C, Feschotte C (2010a) Promiscuous DNA: horizontal transfer of transposable elements and why it matters for eukaryotic evolution. Trends Ecol Evol 25: 537–46.

Schaack S, Choi E, Lynch M, Pritham EJ (2010b) DNA transposons and the role of recombination in mutation accumulation in Daphnia pulex. Genome Biol 11: R46.

Schaefer I, Domes K, Heethoff M, Schneider K, Schön I, et al. (2006) No evidence for the ‘Meselson effect’ in parthenogenetic oribatid mites (Oribatida, Acari). J Evol Biol 19: 184–93.

Schafer FQ, Buettner GR (2001) Redox environment of the cell as viewed through the redox state of the glutathione disulfide/glutathione couple. Free Radic Biol Med 30: 1191–212.

Schaible R, Gerloff-Elias A, Colchero F, Schubert H (2012) Two parthenogenetic populations of Chara canescens differ in their capacity to acclimate to irradiance and salinity. Oecologia 168: 343-53.

Schams D, Berisha B, Kosmann M, Einspanier R, Amselgruber WM (1999) Possible role of growth hormone, IGFs, and IGF-binding proteins in the regulation of ovarian function in large farm animals. Domest Anim Endocrinol 17: 279-85.

Schaper S, Johnston IG, Louis AA (2011) Epistasis can lead to fragmented neutral spaces and contingency in evolution. arXiv:1108.1150v1 [q-bio.PE].

Schapper RM (1998) Antimutator mutants in bacteriophage T4 and Escherichia coli. Genetics 148: 1579–1585.

Schärer OD, Jiricny J (2001) Recent progress in the biology, chemistry and structural biology of DNA glycosylases. Bioessays 23: 270–81.

Scharloo W (1991) Canalization: genetic and developmental aspects. Annu Rev Ecol Syst 22: 65–93.

Scharnweber K, Plath M, Tobler M (2011) Feeding efficiency and food competition in coexisting sexual and asexual livebearing fishes of the genus Poecilia. Environ Biol Fish 90: 197–205.

Schartl M, Nanda I, Schlupp I, Wilde B, Epplen JT, et al. (1995a) Incorporation of subgenomic amounts of DNA as compensation for mutational load in a gynogenetic fish. Nature 373: 68–71.

Schartl M, Wilde B, Schlupp I, Parzefall J (1995b) Evolutionary origin of a parthenoform, the Amazon molly Poecilia formosa, on the basis of a molecular genealogy. Evolution 49: 827–35.

Schatz H, Behan-Pelletier VM (2008) Global diversity of oribatids (Oribatida: Acari: Arachnida). Hydrobiol 595: 323–8.

Scheiner SM (1993) Genetics and evolution of phenotypic plasticity. Annu Rev Ecol Syst 24: 35–68.

Scheiner SM (2002) Selection experiments and the study of phenotypic plasticity. J Evol Biol 15: 889-98.

Scheiner SM, Lyman RF (1989) The genetics of phenotypic plasticity. I. Heritability. J Evol Biol 2: 95–107.

Scheiner SM, Yampolsky LY (1998) The evolution of Daphnia pulex in a temporally varying environment. Genet Res 72: 25-37.

Scheiner SM, Holt RD (2012) The genetics of phenotypic plasticity. X. Variation versus uncertainty. Ecol Evol 2: 751-67.

Schenberg-Frascino A, Moustacchi E (1972) Lethal and mutagenic effects of elevated temperature on haploid yeast. Mol Gen Genet 115: 243-57.

Schetter AJ, Heegaard NH, Harris CC (2010) Inflammation and cancer: interweaving microRNA, free radical, cytokine and p53 pathways. Carcinogenesis 31: 37-49.

Scheu S, Drossel B (2007) Sexual reproduction prevails in a world of structured resources in short supply. Proc Biol Sci274: 1225–31.

Schierwater B, Hadrys H (1998) Environmental factors and mutagenesis in the hydroid Eleutheria dichotoma. Invert Reprod Dev 34: 139-48.

Schild L, Reinheckel T, Reiser M, Horn TF, Wolf G, Augustin W (2003) Nitric oxide produced in rat livermitochondria causes oxidative stress and impairment of respiration after transient hypoxia. FASEB J 17: 2194–201.

Schilling T, Schleithoff ES, Kairat A, Melino G, Stremmel W, et al. (2009) Active transcription of the human FAS/CD95/TNFRSF6 gene involves the p53 family. Biochem Biophys Res Commun 387: 399-404.

Schipper H, Brawer J, Nelson J, Felicio L, Finch C (1981) Role of the gonads in the histologic aging of the hypothalamic arcuate nucleus. Biol Reprod 25: 413-9.

Schisa JA (2012) New insights into the regulation of RNP granule assembly in oocytes. Int Rev Cell Mol Biol 295: 233-89.

Schlacher K, Goodman MF (2007) Lessons from 50 years of SOS DNA-damage-induced mutagenesis. Nat Rev Mol Cell Biol 8: 587–94.

Schlenke TA, Begun DJ (2004) Strong selective sweep associated with a transposon insertion in Drosophila simulans. Proc Natl Acad Sci USA 101: 1626-31.

Schlesinger MJ, Aliperti G, Kelley PM (1982) The response of cells to heat shock. Trends Biochem Sci 7: 222-5.

Schley D, Doncaster CP, Sluckin T (2004) Population models of sperm-dependent parthenogenesis. J Theor Biol 229: 559-72.

Schlichting CD (1986) The evolution of phenotypic plasticity in plants. Annu Rev Ecol Syst 17: 667–93.

Schlichting CD (2004) The role of phenotypic plasticity in diversification. In: DeWitt TJ, Scheiner SM, eds. Phenotypic Plasticity: Functional and Conceptual Approaches. Oxford, UK: Oxford University Press. pp 191–200.

Schlichting CD (2008) Hidden reaction norms, cryptic genetic variation, and evolvability. Ann NY Acad Sci 1133:187-203.

Schlichting CD, Pigliucci M (1998) Phenotypic evolution. A reaction norm perspective. Sunderland, MA: SinauerAssociates.

Schlosser IJ, Doeringsfeld MR, Elder JF, Arzayus LF (1998) Niche relationships of clonal and sexual fish in a heterogeneous landscape. Ecology 79: 953–68.

Schlötterer C (2000) Evolutionary dynamics of microsatellite DNA. Chromosoma 109: 365-71.

Schlupp I (2005) The evolutionary ecology of gynogenesis. Annu Rev Ecol Evol Syst 36: 399-417.

Schlupp I (2010) Mate choice and the Amazon molly: how sexuality and unisexuality can coexist. J Heredity 101: S55–S61.

Schlupp I, Plath M (2005) Male mate choice and sperm allocation in a sexual/asexual mating complex of Poecilia (Poeciliidae, Teleostei). Biol Lett 1: 166–8.

Schlupp I, Taebel-Hellwig A, Tobler M (2010) Equal fecundity in asexual and sexual mollies (Poecilia). Environ Biol Fish 88: 201–6.

Schluter D (2000a) The Ecology of Adaptive Radiations. Oxford, UK: Oxford University Press.

Schluter D (2000b) Ecological character displacement in adaptive radiation. Am Nat 157 (suppl.): S4–S16.

Schmalhausen II (1949) The Factors of Evolution. Philadelphia, PA: Blakiston.

Schmalhausen II (1960) Evolution and cybernetics. Evolution 14: 509–24.

Schmelz RM, Collado R, Myohara M (2000) A taxonomic study of Enchytraeus japonensis (Enchytraeidae, Oligochaeta): morphological and biochemical comparisons with E. bigeminus. Zoo Sci 17: 505–16.

Schmid T, Zhou J, Köhl R, Brüne B (2004) p300 relieves p53-evoked transcriptional repression of hypoxia-inducible factor-1 (HIF-1). Biochem J 380: 289-95.

Schmidt AL, Anderson LM (2006) Repetitive DNA elements as mediators of genomic change in response to environmental cues. Biol Rev Camb Philos Soc 81: 531–43.

Schmidt JA, de Avila JM, McLean DJ (2006) Effect of vascular endothelial growth factor and testis tissue culture on spermatogenesis in bovine ectopic testis tissue xenografts. Biol Reprod 75: 167–75.

Schmidt JA, de Avila JM, McLean DJ (2007) Analysis of gene expression in bovine testis tissue prior to ectopic testis tissue xenografting and during the grafting period. Biol Reprod 76: 1071–80.

Schmidt JM, Good RT, Appleton B, Sherrard J, Raymant GC, et al. (2010) Copy number variation and transposable elements feature in recent, ongoing adaptation at the Cyp6g1 locus. PLoS Genet 6: e1000998.

Schmidt-Nielsen K (1977) Animal Physiology – Adaptation and Environment. Cambridge, UK: Cambridge University Press.

Schmidt-Rose T, Pollet D, Will K, Beremann J, Wittern KP (1999) Analysis of UV-B-induced DNA damage and its repair in heat-shocked skin cells. J Photochem Photobiol B 53: 144-52.

Schmitz RJ, Schultz MD, Lewsey MG, O’Malley RC, Urich MA, et al. (2011) Transgenerational epigenetic instability is a source of novel methylation variants. Science 334: 369–373.

Schmoll T, Dietrich V, Winkel W, Epplen JT, Schurr F, Lubjuhn T (2005) Paternal genetic effects on offspring fitness are context dependent within the extra-pair mating system of a socially monogamous passerine. Evolution 59: 645–57.

Schnakenberg SL, Siegal ML, Bloch Qazi MC (2012) Oh, the places they’ll go: Female sperm storage and sperm precedence in Drosophila melanogaster. Spermatogenesis 2: 224-235.

Schneider JE, Wade GN (1990) Decreased availability of metabolic fuels induces anestrus in golden hamsters. Am J Physiol 258: R750-5.

Schneider JM, Lesmono K (2009) Courtship raises male fertilization success through post-mating sexual selection in a spider. Proc R Soc Lond B Biol Sci 276: 3105–11.

Schneider WL, Roossinck MJ (2001) Genetic diversity in RNA virus quasispecies is controlled by host-virus interactions. J Virol 75: 6566-71.

Schoeberl UE, Kurth HM, Noto T, Mochizuki K (2012) Biased transcription and selective degradation of small RNAs shape the pattern of DNA elimination in Tetrahymena. Genes Dev 26: 1729–42.

Schoener TW (2011) The newest synthesis: understanding the interplay of evolutionary and ecological dynamics. Science 331: 426-9.

Schöffl F, Prändl R, Reindl A (1998) Regulation of the heat-shock response. Plant Physiol 117: 1135–41.

Scholer HR, Ruppert S, Suzuki N, Chowdhury K, Gruss P (1990) New type of POU domain in germ line-specific protein Oct-4. Nature 344: 435–9.

Scholle F, Girard YA, Zhao Q, Higgs S, Mason PW (2004) trans-Packaged West Nile virus-like particles: infectious properties in vitro and in infected mosquito vectors. J Virol 78: 11605–14.

Schön I, Martens K(1998) Opinion: DNA repair in an ancient asexual: a new solution for an old problem? J Nat Hist 32: 943–8.

Schön I, Butlin RK, Griffiths HI, Martens K (1998) Slow molecular evolution in an ancient asexual ostracod. Proc R Soc Lond B Biol Sci 265: 235–42.

Schön I, Gandolfi A, Di Masso E, Rossi V, Griffiths HI, et al. (2000) Persistence of asexuality through mixed reproduction in Eucypris virens (Crustacea, Ostracoda). Heredity (Edinb) 84: 161-9.

Schön I, Martens K (2000) Transposable elements and asexual reproduction. Trends Ecol Evol 15: 287–8.

Schön I, Martens K, Van Doninck K, Butlin RK (2003) Evolution in the slow lane: molecular rates of evolution in sexual and asexual ostracods (Crustacea: Ostracoda). Biol J Linn Soc 79: 93–100.

Schön I, Martens K (2003) No slave to sex. Proc R Soc Lond B Biol Sci 270: 827-33.

Schön I, Arkhipova IR (2006) Two families of non-LTR retrotransposons, Syrinx and Daphne, from the darwinulid ostracod, Darwinula stevensoni. Gene 371: 296–307.

Schön I, Lamatsch D, Martens K (2008) Lessons to learn from ancient asexuals. In: Egel R, Lankenau D-H, eds. Genomic Dynamics and Stability, Vol. 2: Meiosis and Recombination. Crossing Over and Disjunction. Berlin, Germany: Springer. pp 341–376.

Schön I, Rossetti G, Martens K (2009) Darwinulid ostracods: Ancient asexual scandals or scandalous gossip? In: Schön I, Martens K, van Dijk P, eds. Lost sex. Amsterdam, The Netherlands: Springer. pp 217–40.

Schön J, Blottner S, Gabler C, Fickel J (2010) Vascular endothelial growth factor A is a putative paracrine regulator in seasonally controlled spermatogenesis: insights from a ruminant model, the roe deer. Growth Factors 28: 202-10.

Schöneck C, Braun J, Einspanier R (1996) Sperm viability is influenced in vitro by the bovine seminal protein aSFP: effects on motility, mitochondrial activity and lipid peroxidation. Theriogenology 45: 633–42.

Schönfeld P, Schlüter T, Fischer KD, Reiser G (2011) Non-esterified polyunsaturated fatty acids distinctly modulate the mitochondrial and cellular ROS production in normoxia and hypoxia. J Neurochem 118: 69-78.

Schorderet DF, Gartler SM (1992) Analysis of CpG suppression in methylated and nonmethylated species. Proc Natl Acad Sci USA 89: 957–61.

Schoustra SE, Debets AJM, Slakhorst M, Hoekstra RF (2007) Mitotic recombination accelerates adaptation in the fungus Aspergillus nidulans. PLoS Genet 3: e68.

Schoustra S, Rundle HD, Dali R, Kassen R (2010) Fitness-associated sexual reproduction in a filamentous fungus. Curr Biol 20: 1350–5.

Schrag SJ, Mooeres AO, Ndifon GT, Read AF (1994) Ecological correlates of male outcrossing ability in a simultaneous hermaphrodite snail. Am Nat 143: 636–55.

Schramke V, Allshire R (2003) Hairpin RNAs and retrotransposon LTRs effect RNAi and chromatin-based gene silencing. Science 301: 1069–74.

Schreck CB (2010) Stress and fish reproduction: the role of allostasis and hormesis. Gen Comp Endocrinol 165: 549-56.

Schreck R, Rieber P, Baeuerle PA (1991) Reactive oxygen intermediates as apparently widely used messengers in the activation of the NF-kappa B transcription factor and HIV-1. EMBO J 10: 2247-58.

Schreck R, Albermann K, Baeuerle PA (1992) Nuclear factor kappa B: an oxidative stress-responsive transcription factor of eukaryotic cells (a review). Free Radic Res Commun 17: 221-37.

Schreiber V, Ame JC, Dolle P, Schultz I, Rinaldi B, et al. (2002) Poly(ADP-ribose) polymerase-2 (PARP-2) is required for efficient base excision DNA repair in association with PARP-1 and XRCC1. J Biol Chem 277: 23028–36.

Schroedl C, McClintock DS, Budinger GR, Chandel NS (2002) Hypoxic but not anoxic stabilization of HIF-1alpha requires mitochondrial reactive oxygen species. Am J Physiol Lung Cell Mol Physiol 283: L922-31.

Schroeter SC, Dean TA, Thies K, Dixon JD (1995) Effects of shading by adults on the growth of blade-stage Macrosystis pyrifera (Phaeophyta) during and after the 1982–1984 El Niño. J Phycol 31: 697–702.

Schubert HL, Blumenthal RM, Cheng X (2003) Many paths to methyltransfer: a chronicle of convergence. Trends Biochem Sci 28: 329–35.

Schuffenecker I, Ginet N, Goldgar D, Eng C, Chambe B, et al. (1997) Prevalence and parental origin of de novo RET mutations in multiple endocrine neoplasia type 2A and familial medullary thyroid carcinoma. Le Groupe d'Etude des Tumeurs a Calcitonine. Am J Hum Genet 60: 233-7.

Schug MD, Hutter CM, Wetterstrand KA, Gaudette MS, Mackay TFC, et al. (1998) The mutation rates of di-, tri- and tetranucleotide repeats in Drosophila melanogaster. Mol Biol Evol 15: 1751–60.

Schulte RD, Makus C, Hasert B, Michiels NK, Schulenburg H (2010) Multiple reciprocal adaptations and rapid genetic change upon experimental coevolution of an animal host and its microbial parasite. Proc Natl Acad Sci USA 107: 7359–64.

Schultz AH (1938) The relative weight of the testes in primates. Anat Rec 72: 387–94.

Schultz RJ (1969) Hybridization, unisexuallity and polyploidy in the teleost Poeciliopsis (Poecilidae) and other vertebrates. Am Nat 103: 605-19.

Schultz ST, Lynch M (1997) Mutation and extinction: The role of variable mutational effects, synergistic epistasis, beneficial mutations, and degree of outcrossing. Evolution 51: 1363–71.

Schultz ST, Lynch M, Willis JH (1999) Spontaneous deleterious mutation in Arabidopsis thaliana. Proc Natl Acad Sci USA 96: 11393–8.

Schulz RW, de França LR, Lareyre JJ, Le Gac F, Chiarini-Garcia H, et al. (2010) Spermatogenesis in fish. Gen Comp Endocrinol 165: 390-411.

Schumacher B, Hofmann K, Boulton S, Gartner A (2001) The C. elegans homolog of the p53 tumor suppressor is required for DNA damage-induced apoptosis. Curr Biol 11: 1722–7.

Schumacker PT (2002) Hypoxia anoxia and O2 sensing: the search. Am J Physiol Cell Mol Physiol 283: L918–21.

Schurko AM, Neiman M, Logsdon JM Jr (2009) Signs of sex: what we know and how we know it. Trends Ecol Evol 24: 208–17.

Schuster P, Swetina J (1988) Stationary mutant distributions and evolutionary optimization. Bull Math Biol 50: 636-60.

Schützenberger M (1967) Algorithms and neo-Darwinian theory. In: Moorhead PS, Kaplan MM, eds. Mathematical challenges to the neo-Darwinian interpretation of evolution. Philadelphia, PA: Wistar Institute Press. pp 73-75.

Schwacha A, Kleckner N (1994) Identification of joint molecules that form frequently between homologs but rarely between sister chromatids during yeast meiosis. Cell 76: 51–63.

Schwacha A, Kleckner N (1997) Interhomolog bias during meiotic recombination: meiotic functions promote a highly differentiated interhomolog-only pathway. Cell 90: 1123–35.

Schwander T, Crespi BJ (2009) Twigs on the tree of life? Neutral and selective models for integrating macroevolutionary patterns with microevolutionary processes in the analysis of asexuality. Mol Ecol 18: 28-42.

Schwander T, Lee H, Crespi BJ (2011) Molecular evidence for ancient asexuality in Timema stick insects. Curr Biol 21: 1129–34.

Schwartz A, Chan DC, Brown LG, Alagappan R, Pettay D, et al. (1998) Reconstructing hominid Y evolution: X-homologous block, created by X-Y transposition, was disrupted by Yp inversion through LINELINE recombination. Hum Mol Genet 7: 1–11.

Schwartz D, Goldfinger N, Kam Z, Rotter V (1999) p53 controls low DNA damage-dependent premeiotic checkpoint and facilitates DNA repair during spermatogenesis. Cell Growth Differ 10: 665-75.

Schwartz JL, Rotmensch J, Giovanazzi S, Cohen MB, Weichselbaum RR (1988) Faster repair of DNA double-strand breaks in radioresistant human tumor cells. Int J Radiat Oncol Biol Phys 15: 907-12.

Schweizer M, Richter C (1994) Nitric oxide potently and reversibly deenergizes mitochondria at low oxygen tension. Biochem Biophys Res Commun 204: 169–75.

Scocca JJ, Poland RL, Zoon KC (1974) Specificity in deoxyribonucleic acid uptake by transformable Haemophilus influenzae. J Bacteriol 118: 369-73.

Scortegagna M, Ding K, Oktay Y, Gaur A, Thurmond F, Yan LJ, et al. (2003) Multiple organ pathology, metabolic abnormalities and impaired homeostasis of reactive oxygen species in Epas1−/− mice. Nat Genet 35: 331–40.

Scott AP, Sumpter JP (1989) Seasonal variations in testicular germ cell stages and in plasma concentrations of sex steroids in male rainbow trout (Salmo gairdneri) maturing at 2 years old. Gen Comp Endocrinol 73: 46–58.

Scott B, Eaton CJ (2008) Role of reactive oxygen species in fungal cellular differentiations. Curr Opin Microbiol 11: 488–93.

Scott DD, Norbury CJ (2013) RNA decay via 3' uridylation. Biochim Biophys Acta 1829: 654-65.

Scudo FM, Ziegler JR (1978) The Golden Age of Theoretical Ecology: 1923-1940. A collection of works by V. Volterra, V.A. Kostitzin, A.J. Lotka, and A.N. Kolmogoroff. Berlin, Germany: Springer-Verlag.

Seagroves TN, Ryan HE, Lu H, Wouters BG, Knapp M, et al. (2001) Transcription factor HIF-1 is a necessary mediator of the Pasteur effect in mammalian cells. Mol Cell Biol 21: 3436–44.

Seale JV, Wood SA, Atkinson HC, Bate E, Lightman SL, et al. (2004) Gonadectomy reverses the sexually diergic patterns of circadian and stress-induced hypothalamic-pituitary-adrenal axis activity in male and female rats. J Neuroendocrinol 16: 516-24.

Seamons TR, Bentzen P, Quinn TP (2007) DNA parentage analysis reveals inter-annual variation in selection: results from 19 consecutive brood years in steelhead trout. Evol Ecol Res 9: 409–31.

Searcy SP, Sponaugle S (2001) Selective mortality during the larval-juvenile transition in two coral reef fishes. Ecology 82: 2452-70.

Searle S, Blackwell JM (1999) Evidence for a functional-repeat polymorphism in the promoter of human NRAMP1 gene that correlates with autoimmune versus infectious disease susceptibility. J Med Gen 36: 295–9.

Sears A, Holt R, Polis G (2004) Feast and famine in food webs: the effects of pulsed productivity. In: Polis G, Power M, Huxel G, eds. Food webs at the landscape level. Chicago, IL: University of Chicago Press. pp 359-386.

Seaver LC, Imlay JA (2001) Hydrogen peroxide fluxes and compartmentalization inside growing Escherichia coli. J Bacteriol 183: 7182–9.

Seavey SR, Bawa KS (1986) Late-acting selfincompatibility in angiosperms. Bot Rev 52:195–219.

Sebastio G, Villa M, Sartorio R, Guzzetta V, Poggi V, et al. (1989) Control of lactase in human adult-type hypolactasia and in weaning rabbits and rats. Am J Hum Genet 45: 489–97.

Secombe J, Pierce SB, Eisenman RN (2004) Myc: a weapon of mass destruction. Cell 117: 153-6.

Seddon N, AmosW, Mulder RA, Tobias JA (2004) Male heterozygosity predicts territory size, song structure and reproductive success in a cooperatively breeding bird. Proc R Soc B 271: 1823–9.

Sedelnikova OA, Rogakou EP, Panuytin IG, Bonner W (2002) Quantitative detection of 125IUdr-induced DNA double-strand breaks with γ-H2AX antibody. Radiation Res 158: 486-92.

Sedelnikova OA, Redon CE, Dickey JS, Nakamura AJ, Georgakilas AG, Bonner WM (2010) Role of oxidatively induced DNA lesions in human pathogenesis. Mutat Res 704: 152–9.

Seehausen O, van Alphen JJM, Witte F (1997) Cichlid fish diversity threatened by eutrophication that curbs sexual selection. Science 277: 1808–11.

Seehausen O, Takimoto G, Roy D, Jokela J (2008) Speciation reversal and biodiversity dynamics with hybridization in changing environments. Mol Ecol 17: 30–44.

Seetharam S, Seidman MM (1992) Modulation of ultraviolet light mutational hotspots by cellular stress. J Mol Biol 228: 1031-6.

Sega GA (1974) Unscheduled DNA synthesis in the germ cells of male mice exposed in vivo to the chemical mutagen ethyl methanesulfonate. Proc Natl Acad Sci USA 71: 4955–9.

Sega GA (1979) Unscheduled DNA synthesis (DNA repair) in the germ cells of male mice–its role in the study of mammalian mutagenesis. Genetics 92(Suppl 1): s49–s58.

Sega GA, Sotomayor RE (1982) Unscheduled DNA synthesis in mammalian germ cells: its potential use in mutagenicity testing. In: de Serres FJ, Hollander A, eds. Chemical mutagens: principles and methods for their detection (Vol. 7). New York: Plenum Press. pp 421–445.

Segal E, Widom J (2009) What controls nucleosome positions? Trends Genet 25: 335-43.

Seger J, Brockmann HJ (1987) What is bet-hedging? In: Harvey P, Partridge L, eds. Oxford Surveys in Evolutionary Biology, vol. 4, New York, NY: Oxford University Press. pp 182–211.

Segrè D, Deluna A, Church GM, Kishony R (2005) Modular epistasis in yeast metabolism. Nat Genet 37: 77–83.

Ségurel L, Leffler EM, Przeworski M (2011) The case of the fickle fingers: how the PRDM9 zinc finger protein specifies meiotic recombination hotspots in humans. PLoS Biol 9: e1001211.

Seidel HS, Rockman MV, Kruglyak L (2008) Widespread genetic incompatibility in C. elegans maintained by balancing selection. Science 319: 589–94.

Seifer DB, Roapena L, Keefe DL, Zhang HP, Goodman S, et al. (1994) Increasing hypothalamic arcuate nucleus glial peroxidase activity in aging female rats is reduced by an antiestrogen and a gonadotropin releasing hormone agonist. Menopause 1: 83-90.

Seixas S, Ivanova N, Ferreira Z, Rocha J, Victor BL (2012) Loss and gain of function in SERPINB11: An example of a gene under selection on standing variation, with implications for host-pathogen interactions. PLoS ONE 7: e32518.

Seki Y, Yamaji M, Yabuta Y, Sano M, Shigeta M, et al. (2007) Cellular dynamics associated with the genome-wide epigenetic reprogramming in migrating primordial germ cells in mice. Development 134: 2627–38.

Sekita Y, Wagatsuma H, Nakamura K, Ono R, Kagami M, et al. (2008) Role of retrotransposon-derived imprinted gene, Rtl1, in the feto-maternal interface of mouse placenta. Nat Genet 40: 243–8.

Selbach M, Schwanhausser B, Thierfelder N, Fang Z, Khanin R, Rajewsky N (2008) Widespread changes in protein synthesis induced by microRNAs. Nature 455: 58-63.

Selker EU (2004) Genome defense and DNA methylation in Neurospora. Cold Spring Harb Symp Quant Biol 69: 119-24.

Sella G, Petrov DA, Przeworski M, Andolfatto P (2009) Pervasive natural selection in the Drosophila genome? PLoS Genet 5: e1000495.

Selye H (1936) A syndrome produced by diverse nocuous agents. Nature 138: 32.

Semenza GL (1999) Regulation of mammalian O2 homeostasis by hypoxia-inducible factor 1. Annu Rev Cell Dev Biol 15: 551-78.

Semenza GL (2009) Regulation of oxygen homeostasis by hypoxia-inducible factor 1. Physiology (Bethesda) 24: 97-106.

Semenza GL, Wang GL (1992) A nuclear factor induced by hypoxia via de novo protein synthesis binds to the human erythropoietin gene enhancer at a site required for transcriptional activation. Mol Cell Biol 12: 5447–54.

Semenza GL, Roth PH, Fang HM, Wang GL (1994) Transcriptional regulation of genes encoding glycolytic enzymes by hypoxia-inducible factor-1. J Biol Chem 269: 23757–63.

Semercioz A, Onur R, Ogras S, Orhan I (2003) Effects of melatonin on testicular tissue nitric oxide level and antioxidant enzyme activities in experimentally induced left varicocele. Neuroendocrinol Lett 24: 86-90.

Sémon M, Schubert M, Laudet V (2012) Programmed genome rearrangements: in lampreys, all cells are not equal. Curr Biol 22: R641-3.

Sempere LF, Cole CN, McPeek MA, Peterson KJ (2006) The phylogenetic distribution of metazoan microRNAs: insights into evolutionary complexity and constraint. J Exp Zool (Mol Dev Evol) 306B: 575–88.

Semple PD, Beastall GH, Watson WS, Hume R (1980) Serum testosterone depression associated with hypoxia in respiratory failure. Clin Sci (Lond) 58: 105–6.

Sen CK, Packer L (1996) Antioxidant and redox regulation of gene transcription. FASEB J 10: 709-20.

Sena LA, Chandel NS (2012) Physiological roles of mitochondrial reactive oxygen species. Mol Cell 48: 158–67.

Sendhoff B, Kreutz M (1999) A model for the dynamic interaction between evolution and learning. Neural Process Lett 10: 181-93.

Senftleben U, Karin M (2002) The IKK/NF-kappa B pathway. Crit Care Med 30: S18–S26.

Sengstag C (1994) The role of mitotic recombination in carcinogenesis. Crit Rev Toxicol 24: 323–53.

Sengupta MS, Boag PR (2012) Germ granules and the control of mRNA translation. IUBMB Life 64: 586-94.

Sengupta S, Harris CC (2005) p53: traffic cop at the crossroads of DNA repair and recombination. Nat Rev Mol Cell Biol 6: 44-55.

Senoo-Matsuda N, Johnston LA (2007) Soluble factors mediate competitive and cooperative interactions between cells expressing different levels of Drosophila Myc. Proc Natl Acad Sci USA 104: 18543-8.

Senti KA, Brennecke J (2010) The piRNA pathway: a fly's perspective on the guardian of the genome. Trends Genet 26: 499–509.

Seo J, Kim SC, Lee HS, Kim JK, Shon HJ, et al. (2012) Genome-wide profiles of H2AX and γ-H2AX differentiate endogenous and exogenous DNA damage hotspots in human cells. Nucleic Acids Res 40: 5965-74.

Seo TK, Kishino H (2008) Synonymous substitutions substantially improve evolutionary inference from highly diverged proteins. Syst Biol 57: 367-77.

Seo YR, Fishel ML, Amundson S, Kelley MR, Smith ML (2002) Implication of p53 in base excision DNA repair: in vivo evidence. Oncogene 21: 731-7.

Seoighe C, Gehring C, Hurst LD (2005) Gametophytic selection in Arabidopsis thaliana supports the selective model of intron length reduction. PLoS Genet 1: e13.

Seong KH, Li D, Shimizu H, Nakamura R, Ishii S (2011) Inheritance of stress-induced, ATF-2-dependent epigenetic change. Cell 145: 1049-61.

Seppälä O, Jokela J (2010) Maintenance of genetic variation in immune defense of a freshwater snail: role of environmental heterogeneity. Evolution 64: 2397–407.

Seppälä O, Karvonen A, Haataja M, Kuosa M, Jokela J (2011) Food makes you a target: disentangling genetic, physiological, and behavioral effects determining susceptibility to infection. Evolution 65: 1367-75.

Serhal PF, Craft IL (1989) Oocyte donation in 61 patients. Lancet 1: 1185-7.

Serra M, King CE (1999) Optimal rates of bisexual reproduction in cyclical parthenogens with density-dependent growth. J Evol Biol 12: 263-71.

Serra M, Snell TW (2009) Sex loss in monogonont rotifers. In: Schön I, Martens K, Van Dijk P, eds. Lost sex. Berlin: Springer Publications. pp 281-294.

Serrano AL, Andrés V (2004) Telomeres and cardiovascular disease: Does size matter? Circ Res 94: 575–84.

Serrão EA, Pearson GA, Kautsky L, Brawley SH (1996) Successful external fertilization in turbulent environments. Proc Natl Acad Sci USA 93: 5286–90.

Servant G, Pinson B, Tchalikian-Cosson A, Coulpier F, Lemoine S, et al. (2012) Tye7 regulates yeast Ty1 retrotransposon sense and antisense transcription in response to adenylic nucleotides stress. Nucleic Acids Res 40: 5271–82.

Setchell BP (1978) The Mammalian Testis. London: Elek Books, Ltd.

Setchell B (1998) The Parkes lecture: heat and the testis. J Reprod Fertil 114: 179–94.

Setchell BP (2006) The effects of heat on the testes of mammals. Anim Reprod 3: 81-91.

Setchell BP, Waites GM (1964) Blood flow and the uptake of glucose and oxygen in the testis and epididymis of the ram. J Physiol 171: 411–25.

Setchell BP, Maddocks S, Brooks DE (1994) Anatomy, vasculature, innervation, and fluids of the male reproductive tract. In: Knobil E, Neill JD, eds. The Physiology of Reproduction, Vol. I, 2nd edn. New York, NY: Raven Press. pp 1063–1175.

Setchell BP, Bergh A, Widmark A, Damber JE (1995) Effect of temperature of the testis on vasomotion and blood flow. Int J Androl 1995; 18: 120-6.

Setchell JM, Charpentier MJE, Abbott KM, Wickings EJ, Knapp LA (2010) Opposites attract: MHC-associated mate choice in a polygynous primate. J Evol Biol 23: 136-48.

Sethupathy P, Giang H, Plotkin JB, Hannenhalli S (2008) Genome-wide analysis of natural selection on human cis-elements. PLoS ONE 3: e3137.

Sette C, Dolci S, Geremia R, Rossi P (2000) The role of stem cell factor and of alternative c-kit gene products in the establishment, maintenance and function of germ cells. Int J Dev Biol 44: 599-608.

Severson WE, Schmaljohn CS, Javadian A, Jonsson CB (2003) Ribavirin causes error catastrophe during Hantaan virus replication. J Virol 77: 481–8.

Sewerynek E, Wiktorska J, Lewinski A (1999) Effects of melatonin on the oxidative stress induced by thyrotoxicosis in rats. Neuroendocrinol Lett 20: 157–61.

Sexton JP, McIntyre PJ, Angert AL, Rice KJ (2009) Evolution and ecology of species range limits. Annu Rev Ecol Evol Syst 40: 415–36.

Seydoux G, Braun RE (2006) Pathway to totipotency: lessons from germ cells. Cell 127: 891-904.

Sha XY, Xiong ZF, Liu HS, Zheng Z, Ma TH (2011) Pregnant phenotype in aquaporin 8-deficient mice. Acta Pharmacol Sin 32: 840-4.

Shabalina IG, Nedergaard J (2011) Mitochondrial ('mild') uncoupling and ROS production: physiologically relevant or not? Biochem Soc Trans 39: 1305-9.

Shackelford RE, Kaufmann WK, Paules RS (2000) Oxidative stress and cell cycle checkpoint function. Free Radic Biol Med 28: 1387–404.

Shaham S, Reddien PW, Davies B, Horvitz HR (1999) Mutational analysis of the Caenorhabditis elegans cell-death gene ced-3. Genetics 153: 1655-71.

Shahrezaei V, Swain PS (2008) The stochastic nature of biochemical networks. Curr Opin Biotechnol 19: 369-74.

Shahrzad S, Quayle L, Stone C, Plumb C, Shirasawa S, et al. (2005) Ischemia-induced K-ras mutations in human colorectal cancer cells: role of microenvironmental regulation of MSH2 expression. Cancer Res 65: 8134–41.

Shaked H, Kashkush K, Ozkan H, Feldman M, Levy AA (2001) Sequence elimination and cytosine methylation are rapid and reproducible responses of the genome to wide hybridization and allopolyploidy in wheat. Plant Cell 13: 1749–59.

Shalini S, Bansal MP (2005) Role of selenium in regulation of spermatogenesis: involvement of activator protein 1. Biofactors 23: 151–62.

Shalini S, Bansal MP (2007) Alterations in selenium status influences reproductive potential of male mice by modulation of transcription factor NFkappaB. Biometals 20: 49-59.

Shamsi MB, Venkatesh S, Tanwar M, Talwar P, Sharma RK, et al. (2009) DNA integrity and semen quality in men with low seminal antioxidant levels. Mutat Res 665: 29-36.

Shannon M, Lamerdin JE, Richardson L, McCutchen-Maloney SL, Hwang MH, et al. (1999) Characterization of the mouse Xpf DNA repair gene and differential expression during spermatogenesis. Genomics 62: 427-35.

Shapiguzov A, Vainonen JP, Wrzaczek M, Kangasjärvi J (2012) ROS-talk - how the apoplast, the chloroplast, and the nucleus get the message through. Front Plant Sci 3: 292.

Shapiro JA (1984) Observations on the formation of clones containing araB-lacZ cistron fusions. Mol Gen Genet 194:79–90.

Shapiro JA (2009) Revisiting the central dogma in the 21st century. Ann NY Acad Sci 1178: 6-28.

Shapiro JA (2010) Mobile DNA and evolution in the 21st century. Mob DNA 1: 4.

Shapiro JA (2011) Evolution: A View from the 21st Century. Upper Saddle River, NJ: FT Press Science.

Shapiro JA, Dworkin M (1997) Bacteria as multicellular organisms. Oxford, UK: Oxford University Press.

Shapiro JA, Huang W, Zhang C, Hubisz MJ, Lu J, et al. (2007) Adaptive genic evolution in the Drosophila genomes. Proc Natl Acad Sci USA 104: 2271–6.

Shapiro L, Sober E (2007) Epiphenomenalism—the do’s and the don’ts. In: Wolters G, Machamer P, eds. Studies in Causality: Historical and Contemporary. Pittsburgh, PA: University of Pittsburgh Press. pp 235–264.

Shapiro NI (1936) Is there germ cell selection in Drosophila melanogaster? Compt Rend (Dokl.) Acad Sci URSS 3: 119-22.

Sharma RK, Agarwal A (1996) Role of reactive oxygen species in male infertility. Urology 48: 835–50.

Sharma RK, Said T, Agarwal A (2004) Sperm DNA damage and its clinical relevance in assessing reproductive outcome.Asian J Androl 6: 139-48.

Sharp FR, Massa SM, Swanson RA (1999) Heat-shock protein protection. Trends Neurosci 22: 97-9.

Sharp NP, Agrawal AF (2008) Mating density and the strength of sexual selection against deleterious alleles in Drosophila melanogaster. Evolution 62: 857–67.

Sharp PM (1991) Determinants of DNA sequence divergence between Escherichia coli and Salmonella typhimurium: codon usage, map position and concerted evolution. J Mol Evol 33: 23–33.

Sharp PM, Li W-H (1987a) The Codon Adaptation Index—a measure of directional synonymous codon usage bias, and its potential applications. Nucleic Acids Res 15: 1281–95.

Sharp PM, Li W-H (1987b) The rate of synonymous substitution in enterobacterial genes is inversely related to codon usage bias. Mol Biol Evol 4: 222–30.

Sharp PM, Cowe E, Higgins DG, Shields DC, Wolfe KH, et al. (1988) Codon usage patterns in Escherichia coli, Bacillus subtilis, Saccharomyces cerevisiae, Schizosaccharomyces pombe, Drosophila melanogaster and Homo sapiens; a review of the considerable within-species diversity. Nucleic Acids Res 16: 8207–11.

Sharp V, Thurston LM, Fowkes RC, Michael AE (2007) 11β-Hydroxysteroid dehydrogenase enzymes in the testis and male reproductive tract of the boar (Sus scrofa domestica) indicate local roles for glucocorticoids in male reproductive physiology. Reproduction 134: 473-82.

Sharpe RM (1994) Regulation of spermatogenesis. In: Knobil E, Neil, JD, eds. The Physiology of Reproduction. 2nd edn. New York, NY: Raven Press. pp 1363–1434.

Sharpe RM, Maddocks S, Kerr JB (1990) Cell-cell interactions in the control of spermatogenesis as studied using Leydig cell destruction and testosterone replacement. Am J Anat 188: 3-20.

Sharpe RM, McKinnell C, Kivlin C, Fisher JS (2003) Proliferation and functional maturation of Sertoli cells, and their relevance to disorders of testis function in adulthood. Reproduction 125: 769–84.

Sharpley MS, Marciniak C, Eckel-Mahan K, McManus M, Crimi M, et al. (2012) Heteroplasmy of mouse mtDNA is genetically unstable and results in altered behavior and cognition. Cell 151: 333–343.

Shaver AC, Dombrowski PG, Sweeney JY, Treis T, Zappala RM, Sniegowski PD (2002) Fitness evolution and the rise of mutator alleles in experimental Escherichia coli populations. Genetics 162: 557-66.

Shaw FH, Geyer CJ, Shaw RG (2002) A comprehensive model of mutations affecting fitness and inferences for Arabidopsis thaliana. Evolution 56: 453–63.

Shaw FH, Baer CF (2011) Fitness-dependent mutation rates in finite populations. J Evol Biol 24: 1677–84.

Shaw RG, Byers DL, Darmo E (2000) Spontaneous mutational effects on reproductive traits of Arabidopsis thaliana. Genetics 155: 369–78.

Shaw RG, Shaw FH, Geyer C (2003) What fraction of mutations reduces fitness? A reply to Keightley and Lynch. Evolution 57: 686-9.

Shay JES, Simon MC (2012) Hypoxia-inducible factors: Crosstalk between inflammation and metabolism. Semin Cell Dev Biol 23: 389-94.

Shee C, Ponder R, Gibson JL, Rosenberg SM (2011a) What limits the efficiency of double-strand break-dependent stress-induced mutation in Escherichia coli? J Mol Microbiol Biotechnol 21:8-19.

Shee C, Gibson JL, Darrow MC, Gonzalez C, Rosenberg SM (2011b) Impact of a stress-inducible switch to mutagenic repair of DNA breaks on mutation in Escherichia coli. Proc Natl Acad Sci USA 108: 13659-64.

Sheldon BC (2000) Differential allocation: tests, mechanisms and implications. Trends Ecol Evol 15: 397–402.

Sheldon BC, Verhulst S (1996) Ecological immunology: costly parasite defences and trade-offs in evolutionary ecology. Trends Ecol Evol 11: 317–21.

Sheldon BC, Merilä J, Qvarnström A, Gustafsson L, Ellegren H (1997) Paternal genetic contribution to offspring condition predicted by size of male secondary sexual character. Proc R Soc Lond B Biol Sci 264: 297–302.

Shen JC, Rideout WM III , Jones PA (1992) High frequency mutagenesis by a DNA methyltransferase. Cell 71: 1073–80.

Shen SC, Yang LY, Lin HY, Wu CY, Su TH, Chen YC (2008) Reactive oxygen species-dependent HSP90 protein cleavage participates in arsenical As(+3)- and MMA(+3)-induced apoptosis through inhibition of telomerase activity via JNK activation. Toxicol Appl Pharmacol 229: 239-51.

Shenton D, Smirnova JB, Selley JN, Carroll K, Hubbard SJ, et al. (2006) Global translational responses to oxidative stress impact upon multiple levels of protein synthesis. J Biol Chem 281: 29011–21.

Sheoran IS, Saini HS (1996) Drought-induced male sterility in rice: changes in carbohydrate levels and enzyme activities associated with the inhibition of starch accumulation in pollen. Sex Plant Reprod 9: 161–9.

Sherbet DP, Papari-Zareei M, Khan N, Sharma KK, Brandmaier A, et al. (2007) Cofactors, redox state, and directional preferences of hydroxysteroid dehydrogenases. Mol Cell Endocrinol 265-266: 83-8.

Sherengul W, Kondo R, Matsuura ET (2006) Analysis of paternal transmission of mitochondrial DNA in Drosophila. Genes Genet Syst 81: 399–404.

ShethU, Pitt J, Dennis S, Priess JR (2010) Perinuclear P granules are the principal sites of mRNA export in adult C. elegans germ cells. Development 137: 1305–14.

Shetty G, Krishnamurthy H, Krishnamurthy HN, Bhatnagar AS, Moudgal NR (1998) Effect of long-term treatment with aromatase inhibitor on testicular function of adult male bonnet monkeys (M. radiata). Steroids 63: 414–20.

Shi Q, Gibson GE (2011) Up-regulation of the mitochondrial malate dehydrogenase by oxidative stress is mediated by miR-743a. J Neurochem 118: 440–8.

Shiang R, Thompson LM, Zhu YZ, Church DM, Fielder TJ, et al. (1994) Mutations in the transmembrane domain of FGFR3 cause the most common genetic form of dwarfism, achondroplasia. Cell 78: 335–42.

Shibahara S, Müller RM, Taguchi H (1987) Transcriptional control of rat heme oxygenase by heat shock. J Biol Chem 262: 12889-92.

Shields WM (1982) Philopatry, inbreeding, and the evolution of sex. Albany, NY : State University of New York Press.

Shields WM (1988) Sex and adaptation. In: Michod RE, Levin BR, eds. The evolution of sex: an examination of current ideas. Sunderland, MA: Sinauer. pp 253–269.

Shields-Zhou G, Och L (2011) The case for a Neoproterozoic oxygenation event: Geochemical evidence and biological consequences. GSA Today 21: 4-11.

Shigenaga MK, Gimeno CJ, Ames BN (1989) Urinary 8-hydroxy-2'-deoxyguanosine as a biological marker of in vivo oxidative DNA damage. Proc Natl Acad Sci USA 86: 9697-701.

Shigenaga MK, Hagen TM, Ames BN (1994) Oxidative damage and mitochondrial decay in aging. Proc Natl Acad Sci USA 91: 10771-8.

Shih JC, Chen K (2004) Regulation of MAO-A and MAO-B gene expression. Curr Med Chem 11: 1995–2005.

Shih JD, Hunter CP (2011) SID-1 is a dsRNA-selective dsRNA-gated channel. RNA 17: 1057-65.

Shikone T, Billig H, Hsueh AJW (1994) Experimentally-induced cryptorchidism increases apoptosis in rat testis. Biol Reprod 51: 865-72.

Shiloh Y (2003) ATM and related protein kinases: safeguarding genome integrity. Nat Rev Cancer 3: 155–68.

Shima DT, Deutsch U, D’Amore PA (1995) Hypoxic induction of vascular endothelial growth factor (VEGF) in human epithelial cells is mediated by increase in mRNA stability. FEBS Lett 370: 203–8.

Shima JS, Findlay AM (2002) Pelagic larval growth rate impacts benthic settlement and survival of a temperate reef fish. Mar Ecol Prog Ser 235: 303–9.

Shimkets LJ (1999) Intercellular signaling during fruiting-body development of Myxococcus xanthus. Annu Rev Microbiol 53: 525–49.

Shimmin LC, Chang BH, Li WH (1993) Male-driven evolution of DNA sequences. Nature 362: 745-7.

Shinde DN, Elmer DP, Calabrese P, Boulanger J, Arnheim N, Tiemann-Boege I (2013) New evidence for positive selection helps explain the paternal age effect observed in achondroplasia. Hum Mol Genet 2013 Jun 14. [Epub ahead of print]

Shindo C, Lister C, Crevillen P, Nordborg M, Dean C (2006) Variation in the epigenetic silencing of FLC contributes to natural variation in Arabidopsis vernalization response. Genes Dev 20: 3079–83.

Shinohara A, Ogawa H, Ogawa T (1992) Rad51 protein involved in repair and recombination in S. cerevisiae is a RecA-like protein. Cell 69: 457–70.

Shinohara A, Ogawa T (1995) Homologous recombination and the roles of double-strand breaks. Trends Biochem Sci 20: 387–91.

Shinohara R, Mano T, Nagasaka A, Hayashi R, Uchimura K, et al. (2000) Lipid peroxidation levels in rat cardiac muscle are affected by age and thyroid status. J Endocrinol 164: 97–102.

Shiozaki K, Russell P (1996) Conjugation, meiosis, and the osmotic stress response are regulated by Spc1 kinase through Atf1 transcription factor in fission yeast. Genes Dev 10: 2276-88.

Shiraishi K, Naito K (2005) Increased expression of Leydig cell haem oxygenase-1 preserves spermatogenesis in varicocele. Hum Reprod 20: 2608–13.

Shiraishi K, Naito K (2008) Involvement of vascular endothelial growth factor on spermatogenesis in testis with varicocele. Fertil Steril 90: 1313-6.

Shiratsuchi A, Umeda M, Ohba Y, Nakanishi Y (1997) Recognition of phosphatidylserine on the surface of apoptotic spermatogenic cells and subsequent phagocytosis by Sertoli cells of the rat. J Biol Chem 272: 2354–8.

Shiratsuchi A, Kawasaki Y, Ikemoto M, Arai H, Nakanishi Y (1999) Role of class B scavenger receptor type I in phagocytosis of apoptotic rat spermatogenic cells by Sertoli cells. J Biol Chem 274: 5901–8.

Shirayama M, Seth M, Lee HC, Gu W, Ishidate T, et al. (2012) piRNAs initiate an epigenetic memory of nonself RNA in the C. elegans germline. Cell 150: 65–77.

Shkolnik K, Tadmor A, Ben-Dor S, Nevo N, Galiani D, Dekel N (2011) Reactive oxygen species are indispensable in ovulation. Proc Natl Acad Sci USA 108: 1462-7.

Shlomi T, Benyamini T, Gottlieb E, Sharan R, Ruppin E (2011) Genome-scale metabolic modeling elucidates the role of proliferative adaptation in causing the Warburg effect. PLoS Comput Biol 7: e1002018.

Short M, Nemenoff RA, Zawada WM, Stenmark KR, Das M (2004) Hypoxia induces differentiation of pulmonary artery adventitial fibroblasts into myofibroblasts. Am J Physiol Cell Physiol 286: C416-25.

Short RV (1979) Sexual selection and its component parts, somatic and genital selection, as illustrated by man and the great apes. Adv Study Behav 9: 131–58.

Short RV (1997) The testis – the witness of the mating system, the site of mutation and the engine of desire. Acta Paediatr 86: 3–7.

Short RV, Mann T, Hay MF (1967) Male reproductive organs of the African elephant Loxodonta africana. J Reprod Fertil 13: 517–36.

Show MD, Folmer JS, Anway MD, Zirkin BR (2004) Testicular expression and distribution of the rat BCL2 modifying factor in response to reduced intratesticular testosterone. Biol Reprod 70: 1153–61.

Shpiz S, Kwon D, Uneva A, Kim M, Klenov M, et al. (2007) Characterization of Drosophila telomeric retroelement TAHRE: transcription, transpositions and RNAi-based regulation of expression. Mol Biol Evol 24: 2535–45.

Shpiz S, Kwon D, Rozovsky Y, Kalmykova A (2009) rasiRNA pathway controls antisense expression of Drosophila telomeric retrotransposons in the nucleus. Nucleic Acids Res 37: 268–78.

Shrivastav M, De Haro LP, Nickoloff JA (2008) Regulation of DNA double-strand break repair pathway choice. Cell Res 18: 134-47.

Shubin N, Tabin C, Carroll S (1997) Fossils, genes, and the evolution of animal limbs. Nature 388: 639–48.

Shubin N, Tabin C, Carroll S (2009) Deep homology and the origins of evolutionary novelty. Nature 457: 818–23.

Shufran KA, Peters DC, Webster JA (1997) Generation of clonal diversity by sexual reproduction in the greenbug, Schizapis graminum. Insect Mol Biol 6: 203-9.

Shukla LI, Chinnusamy V, Sunkar R (2008) The role of microRNAs and other endogenous small RNAs in plant stress responses. Biochim Biophys Acta 1779: 743-8.

Shvachko LP (2009) DNA hypomethylation as Achilles' heel of tumorigenesis: a working hypothesis. Cell Biol Int 33: 904-10.

Shweiki D, Itin A, Soffer D, Keshet E (1992) Vascular endothelial growth factor induced by hypoxia may mediate hypoxia-initiated angiogenesis. Nature 359: 843–5.

Shweiki D, Itin A, Neufeld G, Gitay-Goren H, Keshet E (1993) Patterns of expression of vascular endothelial growth factor (VEGF) and VEGF receptors in mice suggest a role in hormonally regulated angiogenesis. J Clin Invest 91: 2235–43.

Sia EA, Kokoska RJ, Dominska M, Greenwell P, Petes TD (1997) Microsatellite instability in yeast: dependence on repeat unit size and DNA mismatch repair genes. Mol Cell Biol 17: 2851–8.

Sia EA, Jinks-Robertson S, Petes TD (1997) Genetic control of microsatellite stability. Mutat Res 383: 61–70.

Sibly RM, Calow P (1989) A life-cycle theory of responses to stress. Biol J Linn Soc 37: 101-16.

Sibly RM, Barker D, Denham MC, Hone J, Page M (2005) On the regulation of populations of mammals, birds, fish, and insects. Science 309: 607–10.

Sibly RM, Brown JH, Kodric-Brown A, eds. (2012) Metabolic Ecology: A Scaling Approach. Chichester, UK: John Wiley & Sons.

Sickmann A, Reinders J, Wagner Y, Joppich C, Zahedi R, et al. (2003) The proteome of Saccharomyces cerevisiae mitochondria. Proc Natl Acad Sci USA 100: 13207–12.

Sidorkina O, Espey MG, Miranda KM, Wink DA, Laval J (2003) Inhibition of poly(ADP-ribose) polymerase (PARP) by nitric oxide and reactive nitrogen oxide species. Free Radic Biol Med 35: 1431-8.

Siegel HS (1980) Physiological stress in birds. BioScience 30: 529-34.

Siemons LJ, Mahler CH (1997) Hypogonadotropic hypogonadism in hemochromatosis: recovery of reproductive function after iron depletion. J Clin Endocrinol Metab 65: 585-7.

Siepel A, Haussler D (2004) Phylogenetic estimation of context-dependent substitution rates by maximum likelihood. Mol Biol Evol 21: 468–88.

Sierra S, Dávila M, Lowenstein PR, Domingo E (2000) Response of foot-and-mouth disease virus to increased mutagenesis. Influence of viral load and fitness in loss of infectivity. J Virol 74: 8316–23.

Sies H (1985) Oxidative Stress. New York, NY: Academic Press.

Sies H (1986) Biochemistry of oxidative stress. Angew Chem Int Ed Eng 25: 1058–71.

Sies H (1991) Oxidative stress: oxidants and antioxidants. London, UK: Academic Press.

Sies H (1999) Glutathione and its role in cellular functions. Free Radic Biol Med 27: 916-21.

Sigillo F, Guillou F, Fontaine I, Benahmed M, Magueresse-Battistoni B (1999) In vitro regulation of rat Sertoli cell transferrin expression by tumor necrosis factor α and retinoic acid. Mol Cell Endocrinol 148: 163–70.

Signorovitch AY, Dellaporta SL, Buss LW (2005) Molecular signatures for sex in the Placozoa. Proc Natl Acad Sci USA 102: 15518–22.

Sigurdardóttir S, Helgason A, Gulcher JR, Stefansson K, Donnelly P (2000) The mutation rate in the human mtDNA control region. Am J Hum Genet 66: 1599–609.

Sijen T, Plasterk RH (2003) Transposon silencing in the Caenorhabditis elegans germ line by natural RNAi. Nature 426: 310–4.

Sikka SC (2001) Relative impact of oxidative stress on male reproductive function. Curr Med Chem 8: 851-62.

Silander OK, Tenaillon O, Chao L (2007) Understanding the evolutionary fate of finite populations: the dynamics of mutational effects. PLoS Biol 5: e94.

Siller S (2001) Sexual selection and the maintenance of sex. Nature 411: 689–92.

Silva AJ, Ward K, White R (1993) Mosaic methylation in clonal tissue. Dev Biol 156: 391–8.

Silva JR, Figueiredo JR, van den Hurk R (2009) Involvement of growth hormone (GH) and insulin-like growth factor (IGF) system in ovarian folliculogenesis. Theriogenology 71: 1193-208.

Silverman N, Maniatis T (2001) NF-κB signaling pathways in mammalian and insect innate immunity. Genes Dev 15: 2321–42.

Silvertown J (2008) The evolutionary maintenance of sexual reproduction: evidence from the ecological distribution of asexual reproduction in clonal plants. Int J Plant Sci 169: 157–68.

Silvertown J, Franco M, Pisanty I, Mendoza A (1993) Comparative plant demography – relative importance of life-cycle components to the finite rate of increase in woody and herbaceous perennials. J Ecol 81: 465–76.

Silvertown J, McConway KJ, Hughes Z, Biss P, Macnair M, Lutman P (2002) Ecological and genetic correlates of long-term population trends in the Park Grass Experiment. Am Nat 160: 409–20.

Simic MG, Bergtold DS (1991) Dietary modulation of DNA damage in humans. Mutat Res 250: 17-24.

Simmons LW (1987) Female choice contributes to offspring fitness in the field cricket, Gryllus bimaculatus (De Geer). Behav Ecol Sociobiol 21: 313-21.

Simmons LW (2001) Sperm competition and its evolutionary consequences in the insects. Princeton, NJ: Princeton University Press.

Simmons LW (2005) The evolution of polyandry: sperm competition, sperm selection, and sperm viability. Annu Rev Ecol Evol Syst 36: 125–46.

Simmons LW, Kotiaho JS (2007) Quantitative genetic correlation between trait and preference supports a sexually selected sperm process. Proc Natl Acad Sci USA 104: 16604–8.

Simmons LW, García-González F (2008) Evolutionary reduction in testes size and competitive fertilization success in response to the experimental removal of sexual selection in dung bettles. Evolution 62: 2580–91.

Simmons MJ, Crow JF (1977) Mutations affecting fitness in Drosophila populations. Annu Rev Genet 11: 49–78.

Simmons R (1988) Offspring quality and the evolution of Cainism. Ibis 130: 339–57.

Simmons RE (1997) Why don’t all siblicidal eagles lay insurance eggs? The egg quality hypothesis. Behav Ecol 8: 544–50.

Simmons TW, Jamall IS (1988) Significance of alterations in hepatic antioxidant enzymes. Primacy of glutathione peroxidase. Biochem J 251: 913-7.

Simms EL, Rausher MD (1987) Costs and benefits of plant resistance to herbivory. Am Nat 130: 570–81.

Simms EL, Triplett J (1994) Costs and benefits of plant response to disease: resistance and tolerance. Evolution 48: 1973–85.

Simões A, Costa E (2002) Using genetic algorithms to deal with dynamic environments: A comparative study of several approaches based on promoting diversity. In: Langdon WB et al., eds. Genetic and evolutionary computation - GECCO ’02. New York, NY: Morgan Kaufmann Publishers. pp 9–13.

Simon HU, Haj-Yehia A, Levi-Schaffer F (2000) Role of reactive oxygen species (ROS) in apoptosis induction. Apoptosis 5: 415-8.

Simon JC, Carrel E, Hebert PDN, Dedryver CA, Bonhomme JC, Le Gallic JF (1996) Genetic diversity and mode of reproduction in French populations of the aphid Rhopalosiphum padi. Heredity 76: 305–13.

Simon JC, Rispe C, Sunnucks P (2002) Ecology and evolution of sex in aphids. Trends Ecol Evol 17: 34–9.

Simon JC, Delmotte F, Rispe C, Crease T (2003) Phylogenetic relationships between parthenogens and their sexual relatives: the possible routes to parthenogenesis in animals. Biol J Linn Soc 79: 151–63.

Simon JC, Pfrender ME, Tollrian R, Tagu D, Colbourne JK (2011) Genomics of environmentally induced phenotypes in 2 extremely plastic arthropods. J Hered 102: 512-25.

Simon MC, Ramirez-Bergeron D, Mack F, Hu CJ, Pan Y, Mansfield K (2003) Hypoxia, HIFs, and cardiovascular development. Cold Spring Harb Symp Quant Biol 67: 127–32.

Simon MC, Keith B (2008) The role of oxygen availability in embryonic development and stem cell function. Nat Rev Mol Cell Biol 9: 285-96.

Simonelig M (2011) Developmental functions of piRNAs and transposable elements: a Drosophila point-of-view. RNA Biol 8: 754–9.

Simons AM (2009) Fluctuating natural selection accounts for the evolution of diversification bet hedging. Proc R Soc B 276: 1987–92.

Simons AM (2011) Modes of response to environmental change and the elusive empirical evidence for bet hedging. Proc R Soc B 278: 1601-9.

Simons AM, Roff DA (1994) The effect of environmental variability on the heritabilities of traits of a field cricket. Evolution 48: 1637–49.

Simons AM, Johnston MO (2006) Environmental and genetic sources of diversification in the timing of seed germination: implications for the evolution of bet hedging. Evolution 60: 2280-92.

Simpson AJ (1997) The natural somatic mutation frequency and human carcinogenesis. Adv Cancer Res 71: 209-40.

Simpson AJ, Caballero OL, Jungbluth A, Chen YT, Old LJ (2005) Cancer/testis antigens, gametogenesis and cancer. Nat Rev Cancer 5: 615-25.

Simpson C (2012) The evolutionary history of division of labour. Proc Biol Sci 279: 116-21.

Simpson GG (1944) Tempo and mode in evolution. New York, NY: Columbia University Press.

Simpson GG (1953) The major features of evolution. New York, NY: Columbia University Press.

Simpson P (1979) Parameters of cell competition in the compartments of the wing disc of Drosophila. Dev Biol 69: 182–93.

Simpson P, Morata G (1981) Differential mitotic rates and patterns of growth in compartments in the Drosophila wing. Dev Biol 85: 299–308.

Sinervo B, Svensson E, Comendant T (2000) Density cycles and an offspring quantity and quality game driven by natural selection. Nature 406: 985–8.

Sinervo B, Svensson E (2002) Correlational selection and the evolution of genomic architecture. Heredity 89: 329-38.

Singer B, Kusmierek JT (1982) Chemical mutagenesis. Annu Rev Biochem 51: 655–91.

Singer GA, Hickey DA (2000) Nucleotide bias causes a genomewide bias in the amino acid composition of proteins. Mol Biol Evol 17: 1581–8.

Singer MF (1982) SINEs and LINEs: highly repeated short and long interspersed sequences in mammalian genomes. Cell 28: 433–4.

Singer MF, Skowronski J (1985) Making sense out of LINES: long interspersed repeat sequences in mammalian genomes. Trends Biochem Sci 10: 119-22.

Singh AH, Wolf DM, Wang P, Arkin AP (2008) Modularity of stress response evolution. Proc Natl Acad Sci USA 105: 7500-5.

Singh H, Cheng J, Deng H, Kemp R, Ishizuka T, et al. (2007) Vascular cytochrome P450 4A expression and 20-hydroxyeicosatetraenoic acid synthesis contribute to endothelial dysfunction in androgen-induced hypertension. Hypertension 50: 123-9.

Singh J, Handelsman DJ (1996) Neonatal administration of FSH increases Sertoli cell numbers and spermatogenesis in gonadotropin-deficient (hpg) mice. J Endocrinol 151: 37-48.

Singh R, Kølvraa S, Bross P, Jensen UB, Gregersen N, et al. (2006) Reduced heat shock response in human mononuclear cells during aging and its association with polymorphisms in Hsp70 genes. Cell Stress Chaperones 11: 208–15.

Singh R, Hamada AJ, Agarwal A (2011) Thyroid hormones in male reproduction and fertility. Open Reprod Sci J 3: 98-104.

Singh RP, Shiue K, Schomberg D, Zhou FC (2009) Cellular epigenetic modifications of neural stem cell differentiation. Cell Transplant 18: 1197-211.

Singh RS, Kulathinal RJ (2000) Sex gene pool evolution and speciation: a new paradigm. Genes Genet Syst 75: 119–30.

Singleton TL, Levin HL (2002) A long terminal repeat retrotransposon of fission yeast has strong preferences for specific sites of insertion. Eukaryot Cell 1: 44-55.

Sinha Hikim AP, Bartke A, Russell LD (1988) Morphometric studies on hamster testes in gonadally active and inactive states: light microscope findings. Biol Reprod 39: 1225–37.

Sinha Hikim AP, Wang C, Leung A, Swerdloff R (1995) Involvement of apoptosis in the induction of germ cell degeneration in adult rats after gonadotropin-releasing hormone antagonist treatment. Endocrinology 136: 2770–5.

Sinha Hikim AP, Rajavashisth TB, Sinha Hikim I, Lue Y, Bonavera JJ, et al. (1997) Significance of apoptosis in the temporal and stage-specific loss of germ cells in the adult rat after gonadotropin deprivation. Biol Reprod 57: 1193–201.

Sinha Hikim AP, Swerdloff RS (1999) Hormonal and genetic control of germ cell apoptosis in the testis. Rev Reprod 4: 38-47.

Sinzelle L, Izsvák Z, Ivics Z (2009) Molecular domestication of transposable elements: from detrimental parasites to useful host genes. Cell Mol Life Sci 66: 1073-93.

Siomek A (2012) NF-κB signaling pathway and free radical impact. Acta Biochim Pol 59: 323-31.

Siomi MC, Mannen T, Siomi H (2010) How does the royal family of Tudor rule the PIWI-interacting RNA pathway? Genes Dev 24: 636-46.

Siomi MC, Sato K, Pezic D, Aravin AA (2011) PIWI-interacting small RNAs: the vanguard of genome defence. Nat Rev Mol Cell Biol 12: 246–58.

Sionov RV, Cohen O, Kfir S, Zilberman Y, Yefenof E (2006) Role of mitochondrial glucocorticoid receptor in glucocorticoid-induced apoptosis. J Exp Med 203: 189-201.

Sirinathsinghji DJS, Rees L, Rivier J, Vale W (1983) Corticotropin releasing factor is a potent inhibitor of sexual receptivity in the female rat. Nature 305: 232-5.

Sistonen L, Sarge KD, Morimoto RI (1994) Human heat shock factors 1 and 2 are differentially activated and can synergistically induce hsp70 gene transcription. Mol Cell Biol 14: 2087-99.

Siva-Jothy MT, Tsubaki Y, Hooper RE (1998) Decreased immune response as a proximate cost of copulation and oviposition in a damselfly. Physiol Entomol 23: 274-7.

Siva-Jothy MT, Thompson JJW (2002) Short-term nutrient deprivation affects immune function. Physiol Entomol 27: 206–12.

Sivinski J (1984) Sperm in competition. In: Smith RL, ed. Sperm competition and the evolution of animal mating systems. NewYork, NY: Academic Press. pp 85-115.

Sjöblom T, Lähdetie J (1996) Expression of p53 in normal and gamma-irradiated rat testis suggests a role for p53 in meiotic recombination and repair. Oncogene 12: 2499-505.

Skandalis A, Ford BN, Glickman BW (1994) Strand bias in mutation involving 5-methylcytosine deamination in the human hprt gene. Mutat Res 314: 21–26.

Skilton MR, Viikari JS, Juonala M, Laitinen T, Lehtimäki T, et al. (2011) Fetal growth and preterm birth influence cardiovascular risk factors and arterial health in young adults: the cardiovascular risk in young Finns study. Arterioscler Thromb Vasc Biol 31: 2975-81.

Skinner MK (2005) Sertoli cell secreted regulatory factors. In: Skinner MK, Griswold MD, eds. Sertoli Cell Biology. San Diego, CA: Elsevier Science. pp 107–120.

Skinner MK, Guerrero-Bosagna C (2009) Environmental signals and transgenerational epigenetics. Epigenomics 1: 111–7.

Skulachev VP (1996a) Role of uncoupled and non-coupled oxidations in maintenance of safely low levels of oxygen and its one-electron reductants. Q Rev Biophys 29: 169–202.

Skulachev VP (1996b) Why are mitochondria involved in apoptosis? Permeability transition pores and apoptosis as selective mechanisms to eliminate superoxide-producing mitochondria and cell. FEBS Lett 397: 7-10.

Slack A, Thornton PC, Magner DB, Rosenberg SM, Hastings PJ (2006) On the mechanism of gene amplification induced under stress in Escherichia coli. PLoS Genet 2: e48.

Slade D, Lindner AB, Paul G, Radman M (2009) Recombination and replication in DNA repair of heavily irradiated Deinococcus radiodurans. Cell 136: 1044–55.

Slate J, Kruuk LEB, Marshall TC, Pemberton JM, Clutton-Brock TH (2000) Inbreeding depression influences lifetime breeding success in a wild population of red deer (Cervus elaphus). Proc R Soc B 267: 1657–62.

Slate J, Pemberton JM (2002) Comparing molecular measures for detecting inbreeding depression. J Evol Biol 15: 20–31.

Slatkin M (1974) Hedging one’s evolutionary bets. Nature 250: 704–5.

SlatkinM (2009) Epigenetic inheritance and the missing heritability problem. Genetics 182: 845–50.

Slatyer RA, Mautz BS, Backwell PR, Jennions MD (2012) Estimating genetic benefits of polyandry from experimental studies: a meta-analysis. Biol Rev Camb Philos Soc 87: 1-33.

Slebos RJ, Oh DS, Umbach DM, Taylor JA (2002) Mutations in tetranucleotide repeats following DNA damage depend on repeat sequence and carcinogenic agent. Cancer Res 62: 6052–60.

Sletvold N, Grindeland JM (2007) Fluctuating selection on reproductive timing in Digitalis purpurea. Oikos 116: 473–81.

Sloan DB, Panjeti VG (2010) Evolutionary feedbacks between reproductive mode and mutation rate exacerbate the paradox of sex. Evolution 64: 1129-35.

Sloane MA, Sunnucks P, Wilson AC, Hales DF (2001) Microsatellite isolation, linkage group identification and determination of recombination frequency in the peach–potato aphid, Myzus persicae (Sulzer) (Hemiptera: Aphididae). Genet Res Camb 77: 251–60.

Slobodkin LB (1961) Growth and regulation of animal populations. New York, NY: Holt, Rinehart and Winston.

Slobodkin LB (1974) Prudent predation does not require group selection. Am Nat 108: 665–78.

Sloman KA, Motherwell G, O’Connor KI, Taylor AC (2000) The effect of social stress on the standard metabolic rate (SMR) of brown trout, Salmo trutta. Fish Physiol Biochem 23: 49–53.

Slos S, Stoks R (2008) Predation risk induces stress proteins and reduces antioxidant defense. Funct Ecol 22: 637–42.

Slotkin RK, Freeling M, Lisch D (2005) Heritable transposon silencing initiated by a naturally occurring transposon inverted duplication. Nat Genet 37: 641–4.

Slotkin RK, Martienssen R (2007) Transposable elements and the epigenetic regulation of the genome. Nat Rev Genet 8: 272–85.

Slotkin RK, Vaughn M, Borges F, Tanurdzic M, Becker JD, et al. (2009) Epigenetic reprogramming and small RNA silencing of transposable elements in pollen. Cell 136: 461–72.

Slupphaug G, Kavli B, Krokan HE (2003) The interacting pathways for prevention and repair of oxidative DNA damage. Mutat Res 531: 231-51.

Sluse FE (2012) Uncoupling proteins: molecular, functional, regulatory, physiological and pathological aspects. Adv Exp Med Biol 942: 137-56.

Sluse FE, Jarmuszkiewicz W, Navet R, Douette P, Mathy G, Sluse-Goffart CM (2006) Mitochondrial UCPs: new insights into regulation and impact. Biochim Biophys Acta 1757: 480-5.

Sluzarczýk M (1995) Predator-induced diapause in Daphnia. Ecology 76: 1008–13.

Smagulova F, Gregoretti IV, Brick K, Khil P, Camerini-Otero RD, Petukhova GV (2011) Genome-wide analysis reveals novel molecular features of mouse recombination hotspots. Nature 472: 375-8.

Smalheiser NR, Torvik VI (2005) Mammalian microRNAs derived from genomic repeats. Trends Genet 21: 322–6.

Smalheiser NR, Torvik VI (2006) Alu elements within human mRNAs are probable microRNA targets. Trends Genet 22: 532–6.

Small KS, Brudno M, Hill MM, Sidow A (2007) A haplome alignment and reference sequence of the highly polymorphic Ciona savignyi genome. Genome Biol 3: R41.

Smallwood SA, Kelsey G (2012) De novo DNA methylation: a germ cell perspective. Trends Genet 28: 33-42.

Smiraglia DJ, Kulawiec M, Bistulfi GL, Gupta SG, Singh KK (2008) A novel role for mitochondria in regulating epigenetic modification in the nucleus. Cancer Biol Ther 7: 1182-90.

Smit AFA, Toth G, Riggs AD, Jurka J (1995) Ancestral, mammalian-wide subfamilies of LINE-1 repetitive sequences. J Mol Biol 246: 401–17.

Smith A, Peterson KJ (2002) Dating the time of origin of major clades: Molecular clocks and the fossil record. Annu Rev Earth Planet Sci 30: 65-88.

Smith AC, Shima JS (2011) Variation in the effects of larval history on juvenile performance of a temperate reef fish. Austral Ecol 36: 830-8.

Smith C, Reay P (1991) Cannibalism in teleost fish. Rev Fish Biol Fisheries 1: 41–64.

Smith CC (2012) Opposing effects of sperm viability and velocity on the outcome of sperm competition. Behav Ecol 23 : 820-6.

Smith CC, Fretwell SD (1974) The optimal balance between size and number of offspring. Am Nat 108: 499–506.

Smith CD, Shu S, Mungall CJ, Karpen GH (2007) The Release 5.1 annotation of Drosophila melanogaster heterochromatin. Science 316: 1586–91.

Smith CH (2012a) Natural selection: A concept in need of some evolution? Complexity 17: 8-17.

Smith CH (2012b) Alfred Russel Wallace and the elimination of the unfit. J Biosci 37: 203-5.

Smith DR, Adams AP, Kenney JL, Wang E, Weaver SC (2008) Venezuelan equine encephalitis virus in the mosquito vector Aedes taeniorhynchus: infection initiated by a small number of susceptible epithelial cells and a population bottleneck. Virology 372: 176–86.

Smith J, Tho LM, Xu N, Gillespie DA (2010) The ATM-Chk2 and ATR-Chk1 pathways in DNA damage signaling and cancer. Adv Cancer Res 108: 73-112.

Smith JJ, Antonacci F, Eichler EE, Amemiya CT (2009) Programmed loss of millions of base pairs from a vertebrate genome. Proc Natl Acad Sci USA 106: 11212–7.

Smith JJ, Baker C, Eichler EE, Amemiya CT (2012) Genetic consequences of programmed genome rearrangement. Curr Biol 22: 1524–9.

Smith JM, Smith NH, O'Rourke M, Spratt BG (1993) How clonal are bacteria? Proc Natl Acad Sci USA 90: 4384-8.

Smith KC (1978) Multiple pathways of DNA repair and their possible roles in mutagenesis. Natl Cancer Inst Monogr 107-14.

Smith KC (1992) Spontaneous mutagenesis: experimental, genetic and other factors. Mutat Res 277: 139-62.

Smith LC, Bordignon V, Couto MM, Garcia SM, Yamazaki W, et al. (2002) Mitochondrial genotype segregation and the bottleneck. Reprod Biomed Online 4: 248–255.

Smith NG, Eyre-Walker A (2002) Adaptive protein evolution in Drosophila. Nature 415: 1022–4.

Smith R, Kaune H, Parodi D, Madariaga M, Rios R, et al. (2006) Increased sperm DNA damage in patients with varicocele: relationship with seminal oxidative stress. Hum Reprod 21: 986-93.

Smith RJ, Kamiya T, Horne DJ (2006) Living males of the ‘ancient asexual’ Darwinulidae (Ostracoda: Crustacea). Proc R Soc Lond B Biol Sci 273: 1569-78.

Smith RL, ed. (1984) Sperm competition and the evolution of animal mating systems. London, UK: Academic Press.

Smith TT (1998) The modulation of sperm function by the oviductal epithelium. Biol Reprod 58: 1102–4.

Smits VA (2012) EDD induces cell cycle arrest by increasing p53 levels. Cell Cycle 11: 715-20.

Smits WK, Kuipers OP, Veening JW (2006) Phenotypic variation in bacteria: the role of feedback regulation. Nat Rev Microbiol 4: 259–71.

Snee MJ, Macdonald PM (2004) Live imaging of nuage and polar granules: evidence against a precursor-product relationship and a novel role for Oskar in stabilization of polar granule components. J Cell Sci 117: 2109–20.

Snell TW, Kubanek J, Carter W, Payne AB, Kim J, et al. (2006) A protein signal triggers sexual reproduction in Brachionus plicatilis (Rotifera). Mar Biol 149: 763–73.

Snell-Rood EC (2012) Selective processes in development: implications for the costs and benefits of phenotypic plasticity. Integr Comp Biol 52: 31-42.

Snell-Rood EC, Van Dyken JD, Cruickshank T, Wade MJ, Moczek AP (2010) Toward a population genetic framework of developmental evolution: the costs, limits, and consequences of phenotypic plasticity. Bioessays 32: 71-81.

Sneppen K, Bak P, Flyvbjerg H, Jensen MH (1995) Evolution as a self-organized critical phenomenon. Proc Natl Acad Sci USA 92: 5209-13.

Sniegowski P (2004) Evolution: bacterial mutation in stationary phase. Curr Biol 14: R245-6.

Sniegowski PD, Lenski RE (1995) Mutation and adaptation: the directed mutation controversy in evolutionary perspective. Annu Rev Ecol Syst 26: 553–78.

Sniegowski PD, Gerrish PJ, Lenski RE (1997) Evolution of high mutation rates in experimental populations of E. coli. Nature 387: 703–5.

Sniegowski PD, Gerrish PJ, Johnson T, Shaver A (2000) The evolution of mutation rates: separating causes from consequences. Bioessays 22: 1057-66.

Sniegowski PD, Murphy HA (2006) Evolvability. Curr Biol 16: R831–R834.

Sniegowski PD, Gerrish PJ (2010) Beneficial mutations and the dynamics of adaptation in asexual populations. Philos Trans R Soc Lond B Biol Sci 365: 1255-63.

Snyder LAS, Davies JK, Saunders NJ (2004) Microarray genomotyping of key experimental strains ofNeisseria gonorrhoeae reveals gene complement diversity and five new neisserial genes associated with Minimal Mobile Elements. BMC Genomics 5: 23.

Sober E (2001) The two faces of fitness. In: Singh R, Krimbas CB, Paul D, Beatty J, eds. Thinking about evolution. Cambridge, UK: Cambridge University Press. pp 309–321.

Sober E, Wilson DS (1998) Unto others: The evolution and psychology of unselfish behavior. Cambridge, MA: Harvard University Press.

Sobrino F, Palma EL, Beck E, Dávila M, de la Torre JC, et al. (1986) Fixation of mutations in the viral genome during an outbreak of foot-and-mouth disease: heterogeneity and rate variations. Gene 50: 149–59.

Söder O, Syed V, Callard GV, Toppari J, Pöllänen P, et al. (1991) Production and secretion of an interleukin-l?like factor is stage-dependent and correlates with spermatogonial DNA synthesis in the rat seminiferous epithelium. Int J Androl 14: 223-31.

Sodir NM, Swigart LB, Karnezis AN, Hanahan D, Evan GI, Soucek L (2011) Endogenous Myc maintains the tumor microenvironment. Genes Dev 25: 907–16.

Sogard SM (1997) Size-selective mortality in the juvenile stage of teleost fishes: a review. Bull Mar Sci 60: 1129-57.

Sohail A, Klapacz J, Samaranayake M, Ullah A, Bhagwat AS (2003)Human activation-induced cytidine deaminase causes transcription-dependent, strand-biased C to U deaminations. Nucleic Acids Res 31: 2990-4.

Sohal RS, Allen RG, Nations C (1986)Oxygen free radicals play a role in cellular differentiation: an hypothesis.J Free Radic Biol Med 2: 175-81.

Sohal RS, Svensson I, Brunk UT (1990) Hydrogen peroxide production by liver mitochondria in different species. Mech Ageing Dev 53: 209–15.

Sohal RS, Allen RG (1995) Relationship between metabolic rate, free radicals, differentiation and aging: a unified theory. Basic Life Sci 35: 75–104.

Sohal RS, Weindruch R (1996) Oxidative stress, caloric restriction, and aging. Science 273:59-63.

Sohara E, Ueda O, Tachibe T, Hani T, Jishage K, Rai T, et al. (2007) Morphologic and functional analysis of sperm and testes in Aquaporin 7 knockout mice. Fertil Steril 87: 671–6.

Sola-Penna M (2008) Metabolic regulation by lactate. IUBMB Life 60: 605-8.

Sol-Church K, Stabley DL, Nicholson L, Gonzalez IL, Gripp KW (2006)Paternal bias in parental origin of HRAS mutations in Costello syndrome.Hum Mutat 27:736-41.

Solé RV, Fernandez P, Kauffman SA (2003) Adaptive walks in a gene network model of morphogenesis: insights into the Cambrian explosion. Int J Dev Biol 47: 685–93.

Solé RV, Deisboeck TS (2004) An error catastrophe in cancer? J Theor Biol 228: 47–54.

Solignac M, Génermont J, Monnerot M, Mounolou J-C (1984) Genetics of mitochondria in Drosophila: mtDNA inheritance in heteroplasmic strains of D. mauritiana. Mol Gen Genet 197: 183–8.

Soll DR, Kraft B (1988) A comparison of high frequency switching in the yeast Candida albicans and the slime mold Dictyostelium discoideum. Dev Genet 9: 615-28.

Sollars V, Lu X, Xiao L, Wang X, Garfinkel MD, Ruden DM (2003) Evidence for an epigenetic mechanism by which Hsp90 acts as a capacitor for morphological evolution. Nat Genet 33: 70–4.

Solomon JM, Grossman AD (1996) Who’s competent and when: regulation of natural genetic competence in bacteria. Trends Genet 12: 150-5.

Soltis DE, Soltis PS (1999)Polyploidy: recurrent formation and genome evolution.Trends Ecol Evol 14:348-52.

Soltis DE, Soltis PS, Tate JA (2004) Advances in the study of polyploidy since plant speciation. New Phytol 161: 173–91.

Soltis PS, Soltis DE (2000) The role of genetic and genomic attributes in the success of polyploids.Proc Natl Acad Sci USA 97:7051-7.

Soltis PS, Soltis DE (2009) The role of hybridization in plant speciation. Annu Rev Plant Biol 60: 561–88.

Som C, Reyer HU (2007)Hemiclonal reproduction slows down the speed of Muller's ratchet in the hybridogenetic frog Rana esculenta.J Evol Biol 20: 650-60.

Som C, Bagheri HC, Reyer HU (2007)Mutation accumulation and fitness effects in hybridogenetic populations: a comparison to sexual and asexual systems.BMC Evol Biol 7: 80.

Somma S, Polsinelli M (1970) Quantitative autoradiographic study of competence and deoxyribonucleic acid incorporation in Bacillus subtilis. J Bacteriol 101: 851-5.

Sømme L (1969) Mannitol and glycerol in overwintering aphid eggs. Norw J Entomol 16: 107–11.

Son WY, Hwang SH, Han CT, Lee JH, Kim S, Kim YC (1999) Specific expression of heat shock protein HspA2 in human male germ cells. Mol Hum Reprod 5: 1122–6.

Son WY, Han CT, Hwang SH, Lee JH, Kim S, Kim YC (2000) Repression of hspA2 messenger RNA in human testes with abnormal spermatogenesis. Fertil Steril 73: 1138–44.

Song K, Lu P, Tang K, Osborn TC (1995) Rapid genome change in synthetic polyploids of Brassica and its implications for polyploid evolution.Proc Natl Acad Sci USA 92: 7719-23.

Song L, Lin C, Wu Z, Gong H, Zeng Y, et al. (2011) miR-18a impairs DNA damage response through downregulation of ataxia telangiectasia mutated (ATM) kinase. PLoS ONE 6: e25454.

Song R, Hennig GW, Wu Q, Jose C, Zheng H, Yan W (2011)Male germ cells express abundant endogenous siRNAs.Proc Natl Acad Sci USA 108: 13159-64.

Song Y, Scheu S, Drossel B (2012)The ecological advantage of sexual reproduction in multicellular long-lived organisms.J Evol Biol 25: 556-65.

Sonjak S, Frisvad JC, Gunde-Cimerman N (2007)Genetic variation among Penicillium crustosum isolates from arctic and other ecological niches.Microb Ecol 54:298-305.

Sönmez M, Yüce A, Türk G (2007) The protective effect of melatonin and Vitamin E on antioxidant enzyme activities and epididymal sperm characteristics of homocysteine treated male rats. Reprod Toxicol 23: 226–31.

Sonna LA, Gaffin SL, Pratt RE, Cullivan ML, Angel KC, Lilly CM (2002) Selected contribution: effect of acute heat shock on gene expression by human peripheral blood mononuclear cells. J Appl Physiol 92: 2208–20.

Sonnenblick BP (1950) The early embryology of Drosophila melanogaster. In: Demerec M, ed. Biology of Drosophila. New York, NY: John Wiley.pp 62-167.

Soppa J (2011)Ploidy and gene conversion in Archaea.Biochem Soc Trans 39: 150-4.

Soragni E, Herman D, Dent SY, Gottesfeld JM, Wells RD, Napierala M (2008) Long intronic GAA*TTC repeats induce epigenetic changes and reporter gene silencing in a molecular model of Friedreich ataxia. Nucleic Acids Res 36:6056-65.

Sørensen JG, Kristensen TN, Loeschcke V (2003) The evolutionary and ecological role of heat shock proteins. Ecol Lett 6: 1025–37.

Sorenson MD, Payne RB (2001)A single ancient origin of brood parasitism in African finches: implications for host-parasite coevolution.Evolution 55:2550-67.

Sorger PK, Pelham HR (1988)Yeast heat shock factor is an essential DNA-binding protein that exhibits temperature-dependent phosphorylation.Cell 54: 855-64.

Sorte CJB, Hofmann GE (2004) Changes in latitudes, changes in aptitudes: Nucella canaliculata (Mollusca: Gastropoda) is more stressed at its range edge. Mar Ecol Prog Ser 274: 263–8.

Sosa V, Moliné T, Somoza R, Paciucci R, Kondoh H, LLeonart ME (2013)Oxidative stress and cancer: An overview.Ageing Res Rev 12: 376–90.

Soskine M, Tawfik DS (2010) Mutational effects and the evolution of new protein functions. Nat Rev Genet 11: 572–82.

Sossou M, Flohr-Beckhaus C, Schulz I, Daboussi F, Epe B, Radicella JP (2005) APE1 overexpression in XRCC1-deficient cells complements the defective repair of oxidative single strand breaks but increases genomic instability. Nucleic Acids Res 33: 298–306.

Sotomayor RE, Sega GA, Cumming RB (1978) Unscheduled DNA synthesis in spermatogenic cells of mice treated in vivo with the indirect alkylating agents cyclophosphamide and mitomen. Mutat Res 50: 229–40.

Sotomayor RE, Sega GA, Cumming RB (1979) An autoradiographic study of unscheduled DNA synthesis in the germ cells of male mice treated with X-rays and methyl methanesulfonate. Mutat Res 62: 293–9.

Sotomayor RE, Sega GA (2000) Unscheduled DNA synthesis assay in mammalian spermatogenic cells: an update. Environ Mol Mutagen 36: 255–265.

Soucek L, Whitfield J, Martins CP, Finch AJ, Murphy DJ, et al. (2008) Modelling Myc inhibition as a cancer therapy. Nature 455: 679–83.

Soukhova-O’Hare GK, Shah ZA, Lei Z, Nozdrachev AD, Rao CV, Gozal D (2008) Erectile dysfunction in a murine model of sleep apnea. Am J Respir Crit Care Med 178: 644–50.

Soulé ME, Wilcox BA (1980) Conservation biology: an evolutionary-ecological perspective. Sunderland, MA: Sinauer Associates.

Sousa-Lopes A, Antunes F, Cyrne L, Marinho HS (2004) Decreased cellular permeability to H2O2 protects Saccharomyces cerevisiae cells in stationary phase against oxidative stress. FEBS Lett 578: 152–6.

Southwood TRE (1977) Habitat, the templet for ecological strategies. J Anim Ecol 46:337–365.

Southwood TRE (1988) Tactics, strategies and templets. Oikos 52:3–18.

Sowa Resat MB, Morgan WF (2004) Radiation-induced genomic instability: a role for secreted soluble factors in communicating the radiation response to non-irradiated cells. J Cell Biochem 92: 1013-9.

Spalding DA (1837) Instinct with original observations on young animals. Macmillan’s Magazine 27: 282–93.

Spangelo BL, Judd AM, Call GB, Zumwalt J, Gorospe WC (1995) Role of the cytokines in the hypothalamic-pituitary-adrenal and gonadal axes. Neuroimmunomodulation 2: 299–312.

Sparling PF (1966) Genetic transformation of Neisseria gonorrhoeae to streptomycin resistance. J Bact 92: 1364-71.

Speakman JR (2008) The physiological costs of reproduction in small mammals. Philos Trans R Soc Lond B Biol Sci 363: 375-98.

Speakman JR, Król E (2010) The heat dissipation limit theory and evolution of life histories in endotherms--time to dispose of the disposable soma theory? Integr Comp Biol 50: 793-807.

Spear N, Aust SD (1995) Effects of glutathione on Fenton reagent-dependent radical production and DNA oxidation. Arch Biochem Biophys 324: 111-6.

Specchia V, Piacentini L, Tritto P, Fanti L, D’Alessandro R, et al. (2010) Hsp90 prevents phenotypic variation by suppressing the mutagenic activity of transposons. Nature 463: 662–5.

Speed RM (1988) The possible role of meiotic pairing anomalies in the atresia of human fetal oocytes. Hum Genet 78: 260–6.

Spellig T, Regenfelder E, Reichmann M, Schauwecker F, Bohlmann R, et al. (1994) Control of mating and development in Ustilago maydis. Antonie Van Leeuwenhoek 65: 191-7.

Spiegelman M, Bennett D (1973) A light- and electron-microscopic study of primordial germ cells in the early mouse embryo. J Embryol Exp Morphol 30: 97–118.

Spielman D, Brook BW, Frankham R (2004) Most species are not driven to extinction before genetic factors impact them. Proc Natl Acad Sci USA 101: 15261-4.

Spinnler C, Hedström E, Li H, de Lange J, Nikulenkov F, et al. (2011) Abrogation of Wip1 expression by RITA-activated p53 potentiates apoptosis induction via activation of ATM and inhibition of HdmX. Cell Death Differ 18:1736-45.

Spiropoulos J, Turnbull DM, Chinnery PF (2002) Can mitochondrial DNA mutations cause sperm dysfunction? Mol Hum Reprod 8: 719–21.

Spitze K (1991) Chaoborus predation and lifehistory evolution in Daphnia pulex: temporal pattern of population diversity, fitness, and mean life history. Evolution 45: 82–92.

Spitze K, Burnson J, Lynch M (1991) The covariance structure of life-history characters in Daphnia pulex. Evolution 45: 1081–90.

Sporn MB, Liby KT (2012) NRF2 and cancer: the good, the bad and the importance of context. Nat Rev Cancer 12: 564–71.

Spradling AC (1993) Developmental genetics of oogenesis. In: Bate M, Martinez-Arias A, eds. The Development of Drosophila melanogaster, Vol 1. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press. pp 1-70.

Spriggs KA, Bushell M, Willis AE (2010) Translational regulation of gene expression during conditions of cell stress. Mol Cell 40: 228-37.

Springer MS, Debry RW, Douady C, Amrine HM, Madsen O, et al. (2001) Mitochondrial versus nuclear gene sequences in deep-level mammalian phylogeny reconstruction. Mol Biol Evol 18: 132–43.

Squire RJ, Campbell IH, Allen CM, Wilson CJL (2006) Did the Transgondwanan supermountain trigger the explosive radiation of animals on Earth? Earth Planet Sci Lett 250: 116–33.

Srinivasan DG, Brisson JA (2012) Aphids: a model for polyphenism and epigenetics. Genet Res Int 2012: 431531.

Srinivasan V, Kriete A, Sacan A, Jazwinski SM (2010) Comparing the yeast retrograde response and NF-κB stress responses: implications for aging. Aging Cell 9: 933-41.

Sridharan S, Brehm R, Bergmann M, Cooke PS (2007) Role of connexin 43 in Sertoli cells of testis. Ann NY Acad Sci 1120:131?43.

Srinivasan M, Laychock SG, Hill DJ, Patel MS (2003) Neonatal nutrition: metabolic programming of pancreatic islets and obesity. Exp Biol Med (Maywood) 228: 15-23.

Srivastava N, Raman MJ (2007) Homologous recombination-mediated double-strand break repair in mouse testicular extracts and comparison with different germ cell stages. Cell Biochem Funct 25: 75-86.

Stacey DA, Fellowes MDE (2002) Influence of temperature on pea aphid Acyrthosiphon pisum (Hemiptera: Aphididae) resistance to natural enemy attack. Bull Entomol Res 92: 351–7.

Stalker A, Hermo L, Antakly T (1989) Covalent affinity labeling, radioautography, and immunocytochemistry localize the glucocorticoid receptor in rat testicular Leydig cells. Am J Anat 186: 369–77.

Stamenova R, Dimitrov M, Stoycheva T, Pesheva M, Venkov P, Tsvetkov TS (2008) Transposition of Saccharomyces cerevisiae Ty1 retrotransposon is activated by improper cryopreservation. Cryobiology 56: 241–7.

Stankiewicz M, Mayer MP (2012) The universe of Hsp90. BioMol Concepts 3: 79-97.

Stankov B, Reiter RJ (1990) Melatonin receptors: current status, facts and hypothesis. Life Sci 44: 971–82.

Stanley S (1973) An ecological theory for the sudden origin of multicellular life in the Late Precambrian. Proc Natl Acad Sci USA 70: 1486–9.

Stanley SM (1975) Clades versus clones in evolution: Why we have sex. Science 190: 382–3.

Stanley S (1976) Ideas on the timing of metazoan diversification. Paleobiology 2: 209–19.

Stanley SM, Yang X (1987) Approximate evolutionary stasis for bivalve morphology over millions of years: a multivariate, multilineage study. Paleobiology 13: 113–39.

Stark A, Brennecke J, Bushati N, Russell RB, Cohen SM (2005) Animal microRNAs confer robustness to gene expression and have a significant impact on 3’UTR evolution. Cell 123: 1133–46.

Stark LR, McLetchie DN, Mishler BD (2005) Sex expression, plant size, and spatial segregation of the sexes across a stress gradient in the desert moss Syntrichia caninervis. Bryologist 108: 183–93.

Starrfelt J, Kokko H (2012) Bet-hedging--a triple trade-off between means, variances and correlations. Biol Rev Camb Philos Soc 87: 742-55.

Starz-Gaiano M, Cho NK, Forbes A, Lehmann R (2001) Spatially restricted activity of a Drosophila lipid phosphatase guides migrating germ cells. Development 128: 983–91.

Stearns SC (1983) The influence of size and phylogeny on patterns of covariation among life-history traits in the mammals. Oikos 41: 173–87.

Stearns SC ed. (1987a) The Evolution of Sex and Its Consequences. Basel, Switzerland: Birkhäuser.

Stearns SC (1987b) The selection-arena hypothesis. Experientia Suppl55: 337–49.

Stearns SC (1989) Trade-offs in life-history evolution. Funct Ecol 3: 259–68.

Stearns SC (1990) The evolutionary maintenance of sexual reproduction: the solution proposed for a longstanding problem. J Genet 69: 1–10.

Stearns SC (2002) Progress on canalization. Proc Natl Acad Sci USA 99: 10229-30.

Stebbins G (1950) Variation and evolution in plants. New York, NY: Columbia University Press.

Stebbins GL (1984) Polyploidy and the distribution of arctic-alpine flora: new evidence and a new approach. Bot Helv 94:1–13.

Stebbins GL, Ayala F (1981) Is a new evolutionary synthesis necessary? Science 213: 967-71.

Steele DF, Jinks-Robertson S (1992) An examination of adaptive reversion in Saccharomyces cerevisiae. Genetics 132: 9–21.

Steenken S (1989) Structure, acid/base properties and transformation reactions of purine radicals. Free Radic Res Commun 6: 117–20.

Stefani G, Slack FJ (2008) Small non-coding RNAs in animal development. Nat Rev Mol Cell Biol 9: 219-30.

Stefulj J, Hörtner M, Ghosh M, Schauenstein K, Rinner I, et al. (2001) Gene expression of the key enzymes of melatonin synthesis in extrapineal tissues of the rat.J Pineal Res 30: 243-7.

Stegen JC, Enquist BJ, Ferriere R (2009) Advancing the metabolic theory of diversity. Ecol Lett 12: 1001–15.

Stegen JC, Ferriere R, Enquist BJ (2012) Evolving ecological networks and the emergence of biodiversity patterns across temperature gradients. Proc Biol Sci 279: 1051-60.

Stehli FG, Douglas DG, Newell ND (1969) Generation and maintenance of gradients in taxonomic diversity. Science 164: 947–9.

Stein Z, Stein W, Susser M (1986) Attrition of trisomies as a maternal screening device. Lancet 1: 944–7.

Steinaker DF, Wilson SD (2005) Belowground litter contributions to nitrogen cycling at a northern grassland-forest boundary. Ecology 86: 2825–33.

Steinberg CEW (2012) Stress ecology: environmental stress as ecological driving force and key player in evolution. Dordrecht, The Netherlands: Springer.

Steinger T, Körner C, Schmid B (1996) Long-term persistence in a changing climate: DNA analysis suggests very old ages of clones of alpine Carex curvula. Oecologia 105: 94–9.

Steinhauer DA, Holland JJ (1987) Rapid evolution of RNA viruses. Annu Rev Microbiol 41: 409–33.

Steinhauer J, Kalderon D (2006) Microtubule polarity and axis formation in the Drosophila oocyte. Dev Dyn 235: 1455–68.

Steinmetz M, Minard K, Horvath S, McNicholas J, Srelinger J, et al. (1982) A molecular map of the immune response region from the major histocompatibility complex of the mouse. Nature 300: 35-42.

Stelkens RB, Wedekind C (2010)Environmental sex reversal, Trojan sex genes, and sex ratio adjustment: conditions and population consequences.Mol Ecol 19: 627-46.

Stelling J, Klamt S, Bettenbrock K, Schuster S, Gilles ED (2002) Metabolic network structure determines key aspects of functionality and regulation. Nature 420: 190–3.

Stelling J, Sauer U, Szallas Z, Doyle FJ, Doyle J (2004) Robustness of cellular functions. Cell 118: 675–85.

Stelzer CP (2005) Evolution of rotifer life histories. Hydrobiologia 546: 335–46.

Stelzer CP (2008) Obligate asex in a rotifer and the role of sexual signals. J Evol Biol 21: 287–93.

Stelzer CP(2011) The cost of sex and competition between cyclical and obligate parthenogenetic rotifers.Am Nat 177:E43-53.

Stelzer CP (2012) Population regulation in sexual and asexual rotifers: an eco-evolutionary feedback to population size? Funct Ecol26: 180–8.

Stelzer CP, Snell TW (2003) Induction of sexual reproduction in Brachionus plicatilis (Monogononta, Rotifera) by a density-dependent chemical cue. Limnol Oceanogr 48: 939–43.

Stelzer CP, Schmidt J, Wiedlroither A, Riss S (2010) Loss of sexual reproduction and dwarfing in a small metazoan. PLoS ONE 5: e12854.

Stenberg P, Lundmark M, Knutelski S, Saura A (2003) Evolution of clonality and polyploidy in a weevil system. Mol Biol Evol 20: 1626–32.

Stenseth NC(1985) Darwinian evolution in ecosystems: the Red Queen view. In: Greenwood PJ, Harvey PH, Slatkin M, eds. Evolution. Essays in Honour of John Maynard Smith. Cambridge, UK: Cambridge UniversityPress.pp 55-72.

Stenseth NC, Maynard Smith J (1984) Coevolution in ecosystems: Red Queen evolution or stasis? Evolution 38: 870-80.

Stenseth NC, Kirkendall R, Moran N (1985) On the evolution of pseudogamy. Evolution 39: 294-307.

Stephan H, Polzar B, Rauch F, Zanotti S, Ulke C, Mannherz HG (1996) Distribution of deoxyribonuclease I (DNAse I) and p53 in rat testis and their correlation with apoptosis. Histochem Cell Biol 106: 383–93.

Stephan W, Langley CH (1992) Evolutionary consequences of DNA mismatch inhibited repair opportunity. Genetics 132: 567–74.

Stephan W, Langley CH (1998) DNA polymorphism in lycopersicon and crossing-over per physical length. Genetics 150:1585–93.

Stephens DB (1980) Stress and its measurement in domestic animals: A review of behavioural and physiological studies under field and laboratoryconditions. Adv Vet Sci Comp Med 24: 179-210.

Stephenson AG (1981) Flower and fruit abortion: proximate causes and ultimate functions. Annu Rev Ecol Syst 12:253–79.

Stephenson AG, Winsor JA (1986) Lotus corniculatus regulates offspring quality through selective fruit abortion. Evolution 40: 453-8.

Stevens GC (1989) The latitudinal gradient in geographical range: how so many species co-exist in the tropics. Am Nat 133: 240–56.

Stevens RG, Graubard BI, Micozzi MS, Neriishi K, Blumberg BS (1994) Moderate elevation of body iron level and increased risk of cancer occurrence and death. Int J Cancer 56: 364–9.

Stevison LS (2012) Male-mediated effects on female meiotic recombination. Evolution 66: 905-11.

Stewart AJ, Parsons TL, Plotkin JB (2012)Environmental robustness and the adaptability of populations.Evolution 66: 1598-612.

Stewart JB, Freyer C, Elson JL, Wredenberg A, Cansu Z, et al. (2008a) Strong purifying selection in transmission of mammalian mitochondrial DNA. PLoS Biol 6: e10.

Stewart JB, Freyer C, Elson JL, Larsson NG (2008b)Purifying selection of mtDNA and its implications for understanding evolution and mitochondrial disease.Nat Rev Genet 9:657-62.

Stewart RJ, Preece RF, Sheppard HG (1975) Twelve generations of marginal protein deficiency. Br J Nutr 33: 233–53.

Stewart TA, Bellvé AR, Leder P (1984) Transcription and promoter usage of the c-Myc gene in normal somatic and spermatogenic cells. Science 226: 707–10.

Stewart-Savage J, Bavister BD (1988)Success of fertilization in golden hamsters is a function of the relative gamete ratio.Gamete Res 21: 1-10.

Stibor H (1992) Predator induced life-history shifts in a freshwater cladoceran. Oecologia 92: 162-5.

Stich M, Manrubia SC, Lázaro E (2010) Variable mutation rates as an adaptive strategy in replicator populations. PLoS ONE 5: e11186.

Stinchcombe JR, Dorn LA, Schmitt J (2004) Flowering time plasticity in Arabidopsis thaliana: a reanalysis of Westerman & Lawrence (1970). J Evol Biol 17: 197-207.

Stinson CH (1979) On the selective advantage of fratricide in raptors. Evolution 33: 1219-33.

Stöck M, Lampert K, Möller D, Schlupp I, Schartl M (2010) Monophyletic origin of multiple clonal lineages in an asexual fish (Poecilia formosa). Mol Ecol 19: 5204–15.

Stocco DM (2001) StAR protein and the regulation of steroid hormone biosynthesis. Annu Rev Physiol 63: 193–213.

Stocker R (1990) Induction of haem oxygenase as a defence against oxidative stress. Free Radic Res Commun 9: 101–12.

Stocker R, Yamamoto Y, McDonagh AF, Glazer AN, Ames BN (1987)Bilirubin is an antioxidant of possible physiological importance.Science 235: 1043-6.

Stockley P, Searle JB, Macdonald DW, Jones CS (1993) Female multiple mating behaviour in the common shrew as a strategy to reduce inbreeding. Proc R Soc B 254: 173–9.

Stockley P, Gage MJG, Parker GA, Møller AP (1997) Sperm competition in fishes: the evolution of testis size and ejaculate characteristics. Am Nat 149: 933–54.

Stockley P, Macdonald DW (1998) Why do female common shrews produce so many offspring? Oikos 83: 560–6.

Stoebel DM, Dorman CJ (2010) The effect of mobile element IS10 on experimental regulatory evolution in Escherichia coli. Mol Biol Evol 27: 2105–12.

Stoebel DM, Hokamp K, Last MS, Dorman CJ (2009) Compensatory evolution of gene regulation in response to stress by Escherichia coli lacking RpoS. PLoS Genet 5: e1000671.

Stoffel W, Holz B, Jenke B, Binczek E, Günter RH, et al. (2008)Delta6-desaturase (FADS2) deficiency unveils the role of omega3- and omega6-polyunsaturated fatty acids.EMBO J 27: 2281-92.

Stöhr N, Lederer M, Reinke C, Meyer S, Hatzfeld M, et al. (2006) ZBP1 regulates mRNA stability during cellular stress.J Cell Biol 175: 527-34.

Stojkovic M, Machado SA, Stojkovic P, Zakhartchenko V, Hutzler P, et al. (2001) Mitochondrial distribution and adenosine triphosphate content of bovine oocytes before and after in vitro maturation: Correlation with morphological criteria and developmental capacity after in vitro fertilization and culture. Biol Reprod 64: 904–9.

Stoletzki N, Eyre-Walker A (2007)Synonymous codon usage in Escherichia coli: selection for translational accuracy.Mol Biol Evol 24: 374-81.

Stomp M, van Dijk MA, van Overzee HM, Wortel MT, Sigon CA, et al. (2008)The timescale of phenotypic plasticity and its impact on competition in fluctuating environments.Am Nat 172: 169-85.

Stoner DS, Rinkevich B, Weissman IL (1999) Heritable germ and somatic cell lineage competitions in chimeric colonial protochordates. Proc Natl Acad Sci USA 96: 9148–53.

Storz G, Hengge-Aronis R, eds. (2000) Bacterial stress responses. Washington: ASM Press.

Storz P, Toker A (2003) NF-κB signaling – an alternate pathway for oxidative stress responses. Cell Cycle 2: A8-9.

Stouthamer R, Breeuwert JA, Luck RF, Werren JH (1993) Molecular identification of microorganisms associated with parthenogenesis. Nature 361: 66–8.

Stouthamer R, Werren JH (1993) Microbes associated with parthenogenesis in wasps of the genus Trichogramma. J Invertebr Pathol 61: 6–9.

Stouthamer R, Kazmer DJ (1994) Cytogenetics of microbe-associated parthenogenesis and its consequences for gene flow in Trichogramma wasps. Heredity 73: 317–27.

Stouthamer R, Russell JE, Vavre F, Nunney L (2010)Intragenomic conflict in populations infected by Parthenogenesis Inducing Wolbachia ends with irreversible loss of sexual reproduction.BMC Evol Biol 10: 229.

Stover RH (1986) Banana breeding: polyploidy, disease resistance and productivity. Fruits (France) 41: 175–91.

Stoycheva T, Venkov P, Tsvetkov T (2007a) Mutagenic effect of freezing on mitochondrial DNA of Saccharomyces cerevisiae. Cryobiology 54: 243–50.

Stoycheva T, Massardo DR, Pesheva M, Venkov P, Wolf K, et al. (2007b) Ty1 transposition induced by carcinogens in Saccharomyces cerevisiae yeast depends on mitochondrial function. Gene 389: 212–8.

Stoycheva T, Pesheva M, Venkov P (2010)The role of reactive oxygen species in the induction of Ty1 retrotransposition in Saccharomyces cerevisiae.Yeast 27: 259-67.

Strahl BD, Allis CD (2000) The language of covalent histone modifications. Nature 403: 41-5.

Strahl BD, Briggs SD, Brame CJ, Caldwell JA, Koh SS, et al. (2001) Methylation of histone H4 at arginine 3 occurs in vivo and is mediated by the nuclear receptor coactivator PRMT1. Curr Biol 11: 996–1000.

Strand DJ, McDonald JF (1985) Copia is transcriptionally responsive to environmental stress. Nucleic Acids Res 14: 4401-10.

Strand M, Prolla TA, Liskay RM, Petes TD (1993) Destabilization of tracts of simple repetitive DNA in yeast by mutations affecting DNA mismatch repair. Nature 365: 274–6.

Strasburg J, Kearney M, Moritz C, Templeton AR (2007) Integrating phylogeography with distribution modeling: multiple pleistocene range expansions in a parthenogenetic gecko. PLoS One 2: e760.

Strasser MJ, Mackenzie NC, Dumstrei K, Nakkrasae LI, Stebler J, Raz E (2008) Control over the morphology and segregation of Zebrafish germ cell granules during embryonic development. BMC Dev Biol 8: 58.

Strathdee AT, Bale JS, Block WC, Webb NR, Hodkinson ID, Coulson SJ (1993) Extreme adaptive life-cycle in a high arctic aphid Acyrthosiphon svalbardicum. Ecol Entomol 18: 254–8.

Strathern JN, Shafer BK, McGill CB (1995) DNA synthesis errors associated with double-strand-break repair. Genetics 140: 965-72.

Straub CS, Ives AR, Gratton C (2011)Evidence for a trade-off between host-range breadth and host-use efficiency in aphid parasitoids.Am Nat 177: 389-95.

Strauss BS (1998)Hypermutability in carcinogenesis.Genetics 148: 1619-26.

Strauss BS, Sagher D, Acharya S (1997) Role of proofreading and mismatch repair in maintaining the stability of nucleotide repeats in DNA. Nucleic Acids Res 25: 806–13.

Strauss SY, Rudgers JA, Lau JA, Irwin RE (2002) Direct and ecological costs of resistance to herbivory. Trends Ecol Evol 17: 278-85.

Strbenc M, Fazarinc G, Bavdek SV, Pogacnik A (2003) Apoptosis and proliferation during seasonal testis regression in the brown hare (Lepus europaeus L.). Anat Histol Embryol 32:48–53.

Strickberger MW (1965) Experimental control over the evolution of fitness in laboratory populations of Drosophila pseudoobscura. Genetics 51: 795-800.

Strickler K (1979) Specialization and foraging efficiency of solitary bees. Ecology 60: 998–1009.

Strome S, Wood WB (1982) Immunofluorescence visualization of germ-line-specific cytoplasmic granules in embryos, larvae, and adults of Caenorhabditis elegans. Proc Natl Acad Sci USA 79: 1558–62.

Strome S, Wood WB (1983) Generation of asymmetry and segregation of germ-line granules in early C. elegans embryos. Cell 35: 15–25.

Strome S, Lehman R (2007) Germ versus soma decisions: lessons from flies and worms. Science 316: 392-3.

Stroud CK, Nara TY, Roqueta-Rivera M, Radlowski EC, Lawrence P, et al. (2009) Disruption of FADS2 gene in mice impairs male reproduction and causes dermal and intestinal ulceration.J Lipid Res 50: 1870-80.

Struhl K (2007) Transcriptional noise and the fidelity of initiation by RNA polymerase II. Nat Struct Mol Biol 14: 103–5.

Stuart CA, Banta AM, Tallman J, Cooper HJ (1931) Available food and crowding as factors influencing the sex of Moina macrocopa 1. Physiol Zool 4: 581-93.

Stuart JA, Harper JA, Brindle KM, Brand MD (1999)Uncoupling protein 2 from carp and zebrafish, ectothermic vertebrates. Biochim Biophys Acta 1413: 50–4.

Stumpf JD, Foster PL (2005) Polyphosphate kinase regulates error-prone replication by DNA polymerase IV in Escherichia coli. Mol Microbiol 57: 751–61.

Stumpf MPH, Laidlaw Z, Jansen VAA (2002) Herpes viruses hedge their bets. Proc Natl Acad Sci USA 99: 15234–7.

Sturmey RG, Hawkhead JA, Barker EA, Leese HJ (2009)DNA damage and metabolic activity in the preimplantation embryo.Hum Reprod 24: 81-91.

Sturrock A, Cahill B, Norman K, Huecksteadt TP, Hill K, et al. (2006)Transforming growth factor-beta1 induces Nox4 NAD(P)H oxidase and reactive oxygen species-dependent proliferation in human pulmonary artery smooth muscle cells.Am J Physiol Lung Cell Mol Physiol 290: L661-L673.

Sturtevant AH (1937) Essays on evolution. I. On the effects of selection on mutation rate. Q Rev Biol 12: 464-7.

Stürzbecher HW, Donzelmann B, Henning W, Knippschild U, Buchhop S (1996)p53 is linked directly to homologous recombination processes via RAD51/RecA protein interaction.EMBO J 15:1992-2002.

Su J, Shao X, Liu H, Liu S, Wu Q, Zhang Y (2012)Genome-wide dynamic changes of DNA methylation of repetitive elements in human embryonic stem cells and fetal fibroblasts.Genomics 99: 10-7.

Su W, Qiao Y, Yi F, Guan X, Zhang D, Zhang S, Hao F, Xiao Y, Zhang H, Guo L, Yang L, Feng X, Ma T (2010)Increased female fertility in aquaporin 8-deficient mice.IUBMB Life 62: 852-7.

Suarez SS, Pacey AA (2006) Sperm transport in the female reproductive tract. Human Reprod Update 12:23–37.

Subramaniam K, Seydoux G (1999)nos-1 and nos-2, two genes related to Drosophila nanos, regulate primordial germ cell development and survival in Caenorhabditis elegans.Development 126: 4861-71.

Subramanian S (2012) The abundance of deleterious polymorphisms in humans. Genetics 190: 1579-83.

Subramanian S, Kumar S (2004) Gene expression intensity shapes evolutionary rates of the proteins encoded by the vertebrate genome. Genetics 168:373–81.

Suda T, Takahashi T, Golstein P, Nagata S (1993) Molecular cloning and expression of the Fas ligand, a novel member of the tumor necrosis factor family. Cell 75: 1169–78.

Sueldo C, Marrs RP, Berger T, Kletzky OA, O'Brien TJ (1984) Correlation of semen transferrin concentration and sperm fertilizing capacity. Am J Obstet Gynecol 150: 528-31.

Suerbaum S, Maynard Smith J, Bapumia K, Morelli G, Smith NH, et al. (1998)Free recombination within Helicobacter pylori.Proc Natl Acad Sci USA 95:12619-24.

Suh EK, Yang A, Kettenbach A, Bamberger C, Michaelis AH, et al. (2006) p63 protects the female germ line during meiotic arrest. Nature 444: 624–8.

Sullivan JL, Smith FA, Garman RH (1979) Effects of fluoroacetate on the testis of the rat. J Reprod Fert 56: 201-7.

Sultan SE (2001) Phenotypic plasticity for fitness components in Polygonum species of contrasting ecological breadth. Ecology 82: 328–43.

Sultan SE, Wilczek AM, Hann SD, Brosi BJ (1998a) Contrasting ecological breadth of co-occurring annual Polygonum species. J Ecol 86: 363–83.

Sultan SE, Wilczek AM, Bell DL, Hand G (1998b) Physiological response to complex environments in annual Polygonum species of contrasting ecological breadth. Oecologia 115: 564–78.

Sumedha, Martin OC, Wagner A (2007) New structural variation in evolutionary searches of RNA neutral networks. BioSystems 90: 475–85.

Summers K, McKeon S, Sellars J, Keusenkothen M, Morris J, et al. (2003) Parasitic exploitation as an engine of diversity. Biol Rev 78: 639-75.

Sun H, Treco D, Schultes NP, Szostak JW (1989) Double-strand breaks at an initiation site for meiotic gene conversion. Nature 338: 87–90.

Sun JG, Jurisicova A, Casper RF (1997) Detection of deoxyribonucleic acid fragmentation in human sperm: correlation with fertilization in vitro. Biol Reprod 56: 602–7.

Sun J, Yomogida K, Sakao S, Yamamoto H, Yoshida K, et al. (2009a)Rad18 is required for long-term maintenance of spermatogenesis in mouse testes.Mech Dev 126: 173-83.

Sun J, Yu EY, Yang Y, Confer LA, Sun SH, et al. (2009b) Stn1-Ten1 is an Rpa2-Rpa3-like complex at telomeres. Genes Dev 23: 2900–14.

Sun JX, Helgason A, Masson G, Ebenesersdóttir SS, Li H, et al. (2012)A direct characterization of human mutation based on microsatellites.Nat Genet 44: 1161-5.

Sun QY, Wu GM, Lai L, Park KW, Gabot R, Cheong HT, Day BN, Prather RS, Schatten H (2001) Translocation of active mitochondria during pig oocyte maturation, fertilization and early embryo development in vitro. Reproduction 122: 155–63.

Sun Y, Oberley LW (1996) Redox regulation of transcriptional activators. Free Radic Biol Med 21:335–48.

Sunanaga N, Saito Y, Kawamura K (2006) Postembryonic epigenesis of Vasa-positive germ cells from aggregated hemoblasts in the colonial ascidian, Botryllus primigenus. Dev Growth Differ 48: 87-100.

Sundar IK, Caito S, Yao H, Rahman I (2010)Oxidative stress, thiol redox signaling methods in epigenetics.Methods Enzymol 474: 213-44.

Sundaresan M, Yu ZX, Ferrans VJ, Irani K, Finkel T (1995) Requirement for generation of H2O2 for platelet-derived growth factor signal transduction. Science 270: 296–9.

Sundin GW, Weigand MR (2007) The microbiology of mutability. FEMS Microbiol Lett 277: 11–20.

Sung HM, Yeamans G, Ross CA, Yasbin RE (2003) Roles of YqjH and YqjW, homologs of the Escherichia coli UmuC/DinB or Y superfamily of DNA polymerases, in stationary-phase mutagenesis and UV-induced mutagenesis of Bacillus subtilis. J Bacteriol 185: 2153–60.

Sung JY, Hong JH, Kang HS, Choi I, Lim SD, et al. (2000) Methotrexate suppresses the interleukin-6 induced generation of reactive oxygen species in the synoviocytes of rheumatoid arthritis. Immunopharmacology 47:35-44.

Sunkar R (2010)MicroRNAs with macro-effects on plant stress responses.Semin Cell Dev Biol 21: 805-11.

Sunnucks P, England PR, Taylor AC, Hales DF (1996) Microsatellite and chromosome evolution of parthenogenetic Sitobion aphids in Australia. Genetics 144: 747–56.

Sunnucks P, De Barro PJ, Lushai G, Maclean N, Hales DF (1997) Genetic structure of an aphid studied using microsatellites: cyclic parthenogenesis, differentiated lineages and host specialization. Mol Ecol 6: 1059–73.

Sunnucks P, Chisholm D, Turak E, Hales DF (1998) Evolution of an ecological trait in parthenogenetic Sitobion aphids. Heredity 81: 638–47.

Sunyaev S, Ramensky V, Koch I, Lathe W 3rd, Kondrashov AS, Bork P (2001) Prediction of deleterious human alleles.Hum Mol Genet 10:591-7.

Suomalainen E (1950) Parthenogenesis in animals. Adv Genet 3: 193–253.

Suomalainen E, Saura A, Lokki J (1987) Cytology and evolution in parthenogenesis. Boca Raton, FL: CRC Press.

Suominen JS, Wang Y, Kaipia A, Toppari J (2004)Tumor necrosis factor-alpha (TNF-alpha) promotes cell survival during spermatogenesis, and this effect can be blocked by infliximab, a TNF-alpha antagonist.Eur J Endocrinol 151: 629-40.

Supakar PC, Jung MH, Song CS, Chatterjee B, Roy AK (1995) Nuclear factor κB functions as a negative regulator for the rat androgen receptor gene and NF-κB activity increases during the age-dependent desensitization of the liver. J Biol Chem 270: 837–42.

Surai PF, Blesbois E, Grasseau I, Chalah T, Brillard J-P, et al.(1998) Fatty acid composition, glutathione peroxidase and superoxide dismutase activity and total antioxidant activity of avian semen. Comp Biochem Physiol Part B Biochem Mol Biol 120: 527–33.

Surani MA (2001) Reprogramming of genome function through epigenetic inheritance. Nature 414: 122–8.

Surani MA, Ancelin K, Hajkova P, Lange UC, Payer B, et al. (2004)Mechanism of mouse germ cell specification: a genetic program regulating epigenetic reprogramming.Cold Spring Harb Symp Quant Biol 69: 1-9.

Surani MA, Durcova-Hills G, Hajkova P, Hayashi K, Tee WW (2008) Germ line, stem cells, and epigenetic reprogramming. Cold Spring HarbSymp Quant Biol 73: 9–15.

Suzuki S, Akiyama E (2005) Reputation and the evolution of cooperation in sizable groups. Proc R Soc Lond B 272: 1373–7.

Sutherland HG, Kearns M, Morgan HD, Headley AP, Morris C, et al. (2000) Reactivation of heritably silenced gene expression in mice. Mamm Genome 11: 347–55.

Sutherland WJ (1985) Measures of sexual selection. Oxf Surv Evol Biol 2:90–101.

Sutherland WJ (1987) Random and deterministic components of variance in mating success. In: Bradbury JW, Andersson MB, eds. Sexual selection: testing the alternatives? New York: Wiley. pp 90–101.

Sutherland WJ, Watkinson AR (1986) Somatic mutation: do plants evolve differently? Nature 320:305.

Sutovsky P (2003)Ubiquitin-dependent proteolysis in mammalian spermatogenesis, fertilization, and sperm quality control: killing three birds with one stone. Microsc Res Tech 61:88-102.

Sutovsky P, Navara CS, Schatten G (1996) Fate of the sperm mitochondria, and the incorporation, conversion, and disassembly of the sperm tail structures during bovine fertilization. Biol Reprod 55: 1195–205.

Sutovsky P, Moreno RD, Ramalho-Santos J, Dominko T, Simerly C, Schatten G (1999)Ubiquitin tag for sperm mitochondria.Nature 402: 371-2.

Sutovsky P, Moreno R, Ramalho-Santos J, Dominko T, Thompson WE, Schatten G (2001) A putative ubiquitin-dependent mechanism for the recognition and elimination of defective spermatozoa in the mammalian epididymis. J Cell Sci 114: 1665-75.

Sutovsky P, Neuber E, Schatten G (2002)Ubiquitin-dependent sperm quality control mechanism recognizes spermatozoa with DNA defects as revealed by dual ubiquitin-TUNEL assay.Mol Reprod Dev 61: 406-13.

Sutton MD, Smith BT, Godoy VG, Walker GC (2000) The SOS response: recent insights into umuDC-dependent mutagenesis and DNA damage tolerance. Annu Rev Genet 34: 479–97.

Sutton ML, Gilchrist RB, Thompson JG (2003) Effects of in-vivo and in-vitro environments on the metabolism of the cumulus-oocyte complex and its influence on oocyte developmental capacity.Hum Reprod Update 9:35-48.

Suzukawa K, Miura K, Mitsushita J, Resau J, Hirose K, et al. (2000)Nerve growth factor-induced neuronal differentiation requires generation of Rac1-regulated reactive oxygen species.J Biol Chem 275: 13175-8.

Suzuki HI, Yamagata K, Sugimoto K, Iwamoto T, Kato S, Miyazono K (2009)Modulation of microRNA processing by p53.Nature 460: 529-33.

Suzuki HI, Miyazono K (2010)Dynamics of microRNA biogenesis: crosstalk between p53 network and microRNA processing pathway.J Mol Med (Berl) 88: 1085-94.

Suzuki M, Abe K, Yoshinaga K, Obinata M, Furusawa M (1996) Specific arrest of spermatogenesis caused by apoptotic cell death in transgenic mice. Genes Cells 1: 1077- 86.

Suzuki T, Abe K, Inoue A, Aoki F (2009)Expression of c-MYC in nuclear speckles during mouse oocyte growth and preimplantation development.J Reprod Dev 55: 491-5.

Suzuki Y (2006) Natural selection on the influenza virus genome. MolBiol Evol 23: 1902-11.

Suzuki Y, Nijhout F (2006) Evolution of a polyphenism by genetic accommodation. Science 311: 650–2.

Svanbäck R, Eklöv P (2006) Genetic variation and phenotypic plasticity: causes of morphological and dietary variation in Eurasian perch. Evol Ecol Res 8: 37–49.

Svanbäck R, Bolnick DI (2007) Intraspecific competition promotes resource use diversity within a natural population. Proc R Soc B Biol Sci 274: 839–44.

Svanbäck R, Persson L (2009) Population density fluctuations change the selection gradient in Eurasian perch. Am Nat 173: 507–16.

Svanbäck R, Pineda-Krch M, Doebeli M (2009)Fluctuating population dynamics promotes the evolution of phenotypic plasticity.Am Nat 174: 176-89.

Svardal H, Rueffler C, Hermisson J (2011)Comparing environmental and genetic variance as adaptive response to fluctuating selection.Evolution 65: 2492-513.

Svechnikov KV, Sultana T, Söder O (2001) Age-dependent stimulation of Leydig cell steroidogenesis by interleukin-1 isoforms. Mol Cell Endocrinol 182: 193-201.

Sved J, Bird A (1990) The expected equilibrium of the CpG dinucleotide in vertebrate genomes under a mutation model. Proc Natl Acad Sci USA 87: 4692–6.

Sved JA, Reed TE, Bodmer WF (1967) The number of balanced polymorphisms that can be maintained in a natural population. Genetics 55: 469-81.

Svoboda P, Stein P, Anger M, Bernstein E, Hannon GJ, Schultz RM (2004) RNAi and expression of retrotransposons MuERV-L and IAP in preimplantation mouse embryos. Dev Biol 269: 276–85.

Swain DP; Morgan MJ (2001) Sex-specific temperature distribution in four populations of American plaice Hippoglossoides platessoides. Mar Ecol Prog Ser 212: 233-46.

Swain JL, Stewart TA, Leder P (1987) Parental legacy determines methylation and expression of an autosomal transgene: a molecular mechanism for parental imprinting. Cell 50: 719–27.

Swanson RJ, Serebrovska Z (2012)Intermittent hypoxia remedies male subfertility. In: Xi L, Serebrovskaya TV, eds.Intermittent hypoxia and human diseases. London, UK: Springer-Verlag. pp 221-227.

Swanson WJ, Vacquier VD (1995) Extraordinary divergence and positive Darwinian selection in a fusagenic protein coating the acrosomal process of abalone spermatozoa. Proc Natl Acad Sci USA 92: 4957–61.

Swanson WJ, Vacquier VD (1998) Concerted evolution in an egg receptor for a rapidly evolving abalone sperm protein. Science 281: 710–2.

Swanson WJ, Yang Z, Wolfner MF, Aquadro CF (2001) Positive Darwinian selection drives the evolution of several female reproductive proteins in mammals. Proc Natl Acad Sci USA 98: 2509–14.

Swanson WJ, Vacquier VD (2002) The rapid evolution of reproductive proteins. Nat Rev Genet 3: 137-44.

Swartz HM (1998) EPR studies of cells and tissues. In: Eaton GR, Eaton SS, Salikhov KM, eds. Foundations of Modern EPR. Singapore: World Scientific Publishing Co. pp 451-459.

Sweeney TE, Rozum JS, Desjardins C, Gore RW (1991) Microvascular pressure distribution in the hamster testis. Am J Physiol 260: H1581–H1589.

Sweetwyne MT, Brekken RA, Workman G, Bradshaw AD, Carbon J, et al. (2004)Functional analysis of the matricellular protein SPARC with novel monoclonal antibodies.J Histochem Cytochem 52: 723-33.

Swindell WR, Bouzat JL (2006) Associations between environmental stress, selection history and quantitative genetic variation in Drosophila melanogaster. Genetica 127: 311-20.

Swindell WR, Huebner M, Weber AP (2007) Transcriptional profiling of Arabidopsis heat shock proteins and transcription factors reveals extensive overlap between heat and non-heat stress response pathways.BMC Genomics 8:125.

Sylvester SR, Griswold MD (1994) The testicular iron shuttle: A “nurse” function of the Sertoli cells. J Androl 15: 381-5.

Symer DE, Connelly C, Szak ST, Caputo EM, Cost GJ, et al. (2002) Human L1 retrotransposition is associated with genetic instability in vivo. Cell 110: 327–38.

Symonds ME, Sebert SP, Hyatt MA, Budge H (2009) Nutritional programming of the metabolic syndrome. Nat Rev Endocrinol 5: 604–10.

Syvanen M (2012) Evolutionary implications of horizontal gene transfer. Annu Rev Genet 46: 341–58.

Szabo AR (2002) Experimental tests of intercohort competition for food and cover in the tidepool sculpin (Oligocottus maculosus Girard). Can J Zool 80: 137–44.

Szabo C, Ischiropoulos H, Radi R (2007) Peroxynitrite: biochemistry, pathophysiology and development of therapeutics. Nat Rev 6: 662–80.

Szathmáry E (1993) Do deleterious mutations act synergistically? Metabolic control theory provides a partial answer. Genetics 133: 127–32.

Szathmáry E, Kövér S (1991)A theoretical test of the DNA repair hypothesis for the maintenance of sex in eukaryotes.Genet Res 58: 157-65.

Szathmáry E, Maynard Smith J (1995) The major evolutionary transitions. Nature 374: 227–31.

Sznajder B, Sabelis MW, Egas M (2012) How adaptive learning affects evolution: reviewing theory on the Baldwin effect. Evol Biol 39: 301-10.

Szollosi GJ, Derenyi, I (2009) Congruent evolution of genetic and environmental robustness in micro-RNA. Mol Biol Evol 26: 867–74.

Szostak JW, Orr-Weaver TL, Rothstein RJ, Stahl FW (1983)The double-strand-break repair model for recombination.Cell 33:25-35.

Szyf M, Mcgowan P, Meaney MJ (2008) The social environment and the epigenome. Environ Mol Mutagen 49: 46–60.

Szymanski LA, Tabaac BJ, Schneider JE (2009) Signals that link energy to reproduction: gastric fill, bulk intake, or caloric intake? Physiol Behav 96: 540–7.

Tabara H, Grishok A, Mello CC (1998) RNAi in C. elegans: soaking in the genome sequence. Science 282: 430–1.

Tabatabaie T, Floyd RA (1994) Susceptibility of glutathione peroxidase and glutathione reductase to oxidative damage and the protective effect of spin trapping agent. Arch Biochem Biophys 314: 112-9.

Taberly G (1988) Recherches sur la parthénogenèse thélytoque de deux espèces d’acariens oribatides: Trypochthonius tectorum (Berlese) et Platynothrus peltifer (Koch). IV. Observation sur les males ataviques. Acarologia 29: 95–107.

Tachida H, Iizuka M (1992) Persistence of repeated sequences that evolve by replication slippage. Genetics 131: 471-8.

Tachon P (1989) Ferric and cupric ions requirement for DNA single-strand breakage by H2O2. Free Radic Res Commun 7: 1-10.

Taddei F, Matic I, Radman M (1995) Cyclic AMP-dependent SOS induction and mutagenesis in resting bacterial populations. Proc Natl Acad SciUSA 92: 11736–40.

Taddei F, Radman M, Maynard Smith J, Toupance B, Gouyon PH, Godelle B (1997) Role of mutator alleles in adaptive evolution. Nature 387:700-2.

Tadokoro R, Sugio M, Kutsuna J, Tochinai S, Takahashi Y (2006) Early segregation of germ and somatic lineages during gonad regeneration in the annelid Enchytraeus japonensis. Curr Biol 16: 1012–7.

Taft RJ, Pheasant M, Mattick JS (2007) The relationship between non-protein-coding DNA and eukaryotic complexity. Bioessays 29: 288–99.

Taganov KD, Boldin MP, Chang KJ, Baltimore D (2006) NF-κB-dependent induction of microRNA miR-146, an inhibitor targeted to signaling proteins of innate immune responses. Proc Natl Acad Sci USA 103: 12481–6.

Tagg N, Innes DJ, Doncaster CP (2005a) Outcomes of reciprocal invasions between genetically diverse and genetically uniform populations of Daphnia obtusa (Kurz). Oecologia 143: 527–36.

Tagg N, Doncaster CP, Innes DJ (2005b) Resource competition between genetically varied and genetically uniform populations of Daphnia pulex (Leydig): does sexual reproduction confer a short-term ecological advantage? Biol J Linn Soc 85: 111–23.

Taguchi A, Tak M, Motoishi M, Or H, Moch M, Watanabe K (2012) Analysis of localization and reorganization of germ plasm in Xenopus transgenic line with fluorescence?labeled mitochondria. Dev Growth Diff 54: 767-76.

Tahiliani M, Koh KP, Shen Y, Pastor WA, Bandukwala H, et al. (2009) Conversion of 5-methylcytosine to 5-hydroxymethylcytosine in mammalian DNA by MLL partner TET1. Science 324: 930–5.

Taipale M, Jarosz DF, Lindquist S (2010) HSP90 at the hub of protein homeostasis: emerging mechanistic insights. Nat Rev Mol Cell Biol 11: 515–28.

Tait DEN (1980) Abandonment as a reproductive tactic -  the example of grizzly bears. Am Nat 115: 779-808.

Takada S, Berezikov E, Choi YL, Yamashita Y, Mano H (2009) Potential role of miR-29b in modulation of Dnmt3a and Dnmt3b expression in primordial germ cells of female mouse embryos. RNA 15: 1507–14.

Takahashi A, Ohnishi T (2005)Does gammaH2AX foci formation depend on the presence of DNA double strand breaks?Cancer Lett 229: 171-9.

Takahashi KH (2012)Multiple capacitors for natural genetic variation in Drosophila melanogaster.Mol Ecol 22: 1356-65.

Takahashi K, Shichijo S, Noguchi M, Hirohata M, Itoh K (1995) Identification of MAGE-1 and MAGE-4 proteins in spermatogonia and primary spermatocytes of testis. Cancer Res 55: 3478–82.

Takahashi K, Yamanaka S (2006) Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell 126: 663–76.

Takahashi-Yanaga F, Taba Y, Miwa Y, Kubohara Y, Watanabe Y, et al. (2003) Dictyostelium differentiation-inducing factor-3 activates glycogen synthase kinase-3ß and degrades cyclin D1 in mammalian cells. J Biol Chem 278: 9663–70.

Takahata N (1993) Allelic genealogy and human evolution. Mol Biol Evol10:2–22.

Takami M, Preston SL, Toyloy VA, Behrman HR (1999) Antioxidants reversibly inhibit the spontaneous resumption of meiosis. Am J Physiol 276: E684–E688.

Takami M, Preston SL, Behrman HR (2000) Eicosatetraynoic and eicosatriynoic acids, lipoxygenase inhibitors, block meiosis via antioxidant action. Am J Physiol Cell Physiol 278: C646–C650.

Takasuka A, Aoki I, Mitani I (2003) Evidence of growth-selective predation on larval Japanese anchovy Engraulis japonicus in Sagami Bay. Mar Ecol Prog Ser 252: 223–38.

Takeda T, Toda T, Kominami KI, Kohnosu A, Yanagida M, Jones N (1995) Schizosaccharomyces pombe atf1+ encodes a transcription factor required for sexual development and entry into stationary phase. EMBO J 14: 6193-208.

Takemoto D, Tanaka A, Scott B (2007)NADPH oxidases in fungi: diverse roles of reactive oxygen species in fungal cellular differentiation.Fungal Genet Biol 44: 1065-76.

Takeuchi T, Neri QV, Katagiri Y, Rosenwaks Z, Palermo GD (2005) Effect of treating induced mitochondrial damage on embryonic development and epigenesis. Biol Reprod 72: 584–92.

Tal O, Kisdi E, Jablonka E (2010)Epigenetic contribution to covariance between relatives.Genetics 184: 1037-50.

Tam KKY, Russell DL, Peet DJ, Bracken CP, Rodgers RJ, et al. (2010)Hormonally regulated follicle differentiation and luteinization in the mouse is associated with hypoxia inducible factor activity.Mol Cell Endocrinol 327:47-55.

Tam PP, Snow MH (1981)Proliferation and migration of primordial germ cells during compensatory growth in mouse embryos.J Embryol Exp Morphol 64: 133-47.

Tamanini C, De Ambrogi M (2004)Angiogenesis in developing follicle and corpus luteum.Reprod Domest Anim 39:206-16.

Tamarkin L, Baird CJ, Almeida OF (1985)Melatonin: a coordinating signal for mammalian reproduction?Science 227: 714-20.

Tamassia M, Nuttinck F, May-Panloup P, Reynier P, Heyman Y, et al. (2004) In vitro embryo production efficiency in cattle and its association with oocyte adenosine triphosphate content, quantity of mitochondrial DNA, and mitochondrial DNA haplogroup. Biol Reprod 71: 697–704.

Tamir S, Rojas-Walker T, Wishnor JS, Tannenbaum SR (1996) DNA damage and genotoxicity by nitric oxide. Meth Enzymol 269: 230-42.

Tamir S, Tannenbaum SR (1996) The role of nitric oxide (NO.) in the carcinogenic process. Biochim Biophys Acta 1288: F31-F36.

Tamori Y, Bialucha CU, Tian AG, Kajita M, Huang YC, et al. (2010)Involvement of Lgl and Mahjong/VprBP in cell competition. PLoS Biol 8: e1000422.

Tamori Y, Deng WM (2011)Cell competition and its implications for development and cancer.J Genet Genomics 38: 483-95.

Tamori Y, Deng WM (2013)Tissue repair through cell competition and compensatory cellular hypertrophy in postmitotic epithelia.Dev Cell 25: 350-63.

Tamura H, Takasaki A, Miwa I, Taniguchi K, Maekawa R, et al. (2008) Oxidative stress impairs oocyte quality and melatonin protects oocytes from free radical damage and improves fertilization rate. J Pineal Res 44: 280–7.

Tan DX, Chen LD, Poeggeler B, Manchester LC, Reiter RJ (1993) Melatonin: a potent, endogenous hydroxyl radical scavenger. Endocrine J 1:57–60.

Tan DX, Reiter RJ, Manchester LC, Yan MT, El-Sawi M, et al. (2002) Chemical and physical properties and potential mechanisms: Melatonin as a broad-spectrum antioxidant and free radical scavenger. Curr Topics Med Chem 2: 181-98.

Tan DX, Manchester LC, Sainz RM, Mayo JC, Alvarez F, Reiter RJ (2003) Antioxidant strategies in protection against neurodegenerative disorders. Expert Opin Ther Patents 13: 1513–43.

Tan DX, Hardeland R, Manchester LC, Paredes SD, Korkmaz A, et al. (2010)The changing biological roles of melatonin during evolution: from an antioxidant to signals of darkness, sexual selection and fitness.Biol Rev Camb Philos Soc 85: 607-23.

Tan JC, Nocka K, Ray P, Traktman P, Besmer P (1990) The dominant W42 spotting phenotype results from a missense mutation in the c-kit receptor kinase. Science 247: 209-12.

Tan L, Shi YG (2012)Tet family proteins and 5-hydroxymethylcytosine in development and disease.Development 139: 1895-902.

Tan M, Li S, Swaroop M, Guan K, Oberley LW, Sun Y (1999)Transcriptional activation of the human glutathione peroxidase promoter by p53.J Biol Chem 274: 12061-6.

Tan NW, Li BFL (1990) Interaction of oligonucleotides containing 6-Omethylguanine with human DNA (cytosine-5-)-methyltransferase. Biochemistry 29: 9234–40.

Tan S, Sagara Y, Liu Y, Maher P, Schubert D (1998) The regulation of reactive oxygen species production during programmed cell death. J Cell Biol 141: 1423–32.

Tanaka EM (2003) Regeneration: if they can do it, why can’t we? Cell 113: 559–62.

Tanaka H, Makino Y, Okamoto K, Iida T, Yan K, Yoshikawa N (1999) Redox regulation of the glucocorticoid receptor. Antiox Redox Signal 1: 403–23.

Tanaka MM, Bergstrom CT, Levin BR (2003)The evolution of mutator genes in bacterial populations: the roles of environmental change and timing.Genetics 164: 843-54.

Tanaka N, Katoh M (1979) Unscheduled DNA synthesis in the germ cells of male mice in vivo. Jpn J Hum Genet 54: 405–14.

Tanaka T, Kurose A, Halicka HD, Traganos F, Darzynkiewicz Z (2006a) 2-Deoxy-D-glucose reduces the level of constitutive activation of ATM and phosphorylation of histone H2AX. Cell Cycle 5: 878-82.

Tanaka T, Halicka HD, Huang X, Traganos F, Darzynkiewicz Z (2006b)Constitutive histone H2AX phosphorylation and ATM activation, the reporters of DNA damage by endogenous oxidants.Cell Cycle 5: 1940-5.

Tanaka Y (1996) Sexual selection enhances population extinction in a changing environment. J Theor Biol 180: 197–206.

Tang CH, Wei W, Liu L (2012)Regulation of DNA repair by S-nitrosylation.Biochim Biophys Acta 1820: 730-5.

Tang CS, Epstein RJ (2010)Adaptive evolution hotspots at the GC-extremes of the human genome: evidence for two functionally distinct pathways of positive selection.Adv Bioinformatics 2010: 856825.

Tang F, Kaneda M, O’Carroll D, Hajkova P, Barton SC, et al. (2007) Maternal microRNAs are essential for mouse zygotic development. Genes Dev 21: 644–8.

Tang KF, Ren H (2012)The role of dicer in DNA damage repair.Int J Mol Sci 13: 16769-78.

Taniguchi Y, Nakano S (2000) Condition-specific competition: implications for the altitudinal distribution of stream fishes. Ecology 81: 2027–39.

Tanimoto K, Makino Y, Pereira T, Poellinger L (2000) Mechanism of regulation of the hypoxia-inducible factor-1alpha by the von Hippel-Lindau tumor suppressor protein. Embo J 19: 4298–309.

Tannenbaum E (2008) A comparison of three replication strategies in complex multicellular organisms: Asexual replication, sexual replication with identical gametes, and sexual replication with distinct sperm and egg gametes. Phys Re E 77: 011915.

Tannenbaum E, Sherley JL, Shakhnovich EI (2006) Semiconservative quasispecies equations for polysomic genomes: the haploid case. J Theor Biol 241: 791–805.

Tanner JT (1966) Effects of population density on growth rates of animal populations. Ecology 47: 733–45.

Tapanainen JS, Tilly JL, Vihko KK, Hsueh AJW (1993) Hormonal control of apoptotic cell death in the testis: gonadotropins and androgens as testicular cell survival factors. Mol Endocrinol 7: 643–50.

Tapia G, Cornejo P, Fernandez V, Videla LA (1999) Protein oxidation in thyroid hormone-induced liver oxidative stress: relation to lipid peroxidation. Toxicol Lett 106: 209–14.

Tapia N, Schöler HR (2010)p53 connects tumorigenesis and reprogramming to pluripotency.J Exp Med 207: 2045-8.

Tarin JJ (1996) Potential effects of age-associated oxidative stress on mammalian oocytes/embryos. Mol Hum Reprod 2: 717–24.

Tarin JJ, Vendrell FJ, Ten J, Cano A (1998) Antioxidant therapy counteracts the disturbing effects of diamide and maternal ageing on meiotic division and chromosomal segregation in mouse oocytes. Mol Hum Reprod 4: 281-8.

Tariq M, Nussbaumer U, Chen Y, Beisel C, Paro R (2009) Trithorax requires Hsp90 for maintenance of active chromatin at sites of gene expression. Proc Natl Acad Sci USA 106: 1157–62.

Tartaglia M, Cordeddu V, Chang H, Shaw A, Kalidas K, et al. (2004) Paternal germline origin and sex-ratio distortion in transmission of PTPN11 mutations in Noonan syndrome. Am J Hum Genet 75: 492-7.

Tarulli GA, Stanton PG, Lerchl A, Meachem SJ (2006) Adult Sertoli cells are not terminally differentiated in the Djungarian hamster: effect of FSH on proliferation and junction protein organization. Biol Reprod 74: 798–806.

Tarulli GA, Stanton PG, Meachem SJ (2012)Is the adult sertoli cell terminally differentiated?Biol Reprod 87: 13, 1-11.

Taruscio D, Mantovani A (2004) Factors regulating endogenous retroviral sequences in human and mouse. Cytogenet Genome Res 105: 351–62.

Tatsuta T, Langer T (2008) Quality control of mitochondria: protection against neurodegeneration and ageing. EMBO J 27: 306-14.

Tautz D (1992) Redundancies, development and the flow of information. BioEssays 14: 263–6.

Tautz D, Renz M (1984) Simple sequences are ubiquitous repetitive components of eukaryotic genomes. Nucleic Acids Res 12: 4127-38.

Tautz D, Schlötterer C (1994) Simple sequences. Curr Opin Genet Dev 4: 832-7.

Tawa R, Kimura Y, Komura J, Miyamura Y, Kurishita A, et al. (1998) Effects of X-ray irradiation on genomic DNA methylation levels in mouse tissues. J Radiat Res 39: 271–8.

Taylor CE, Condra C (1980) r- and K-selection in Drosophila pseudoobscura. Evolution 34: 1183-93.

Taylor CT, Moncada S (2010)Nitric oxide, cytochrome C oxidase, and the cellular response to hypoxia.Arterioscler Thromb Vasc Biol 30: 643-7.

Taylor EB, Rutter J (2011) Mitochondrial quality control by the ubiquitin-proteasome system. Biochem Soc Trans 39: 1509–13.

Taylor J, Tyekucheva S, Zody M, Chiaromonte F, Makova KD (2006) Strong and weak male mutation bias at different sites in the primate genomes: insights from the human-chimpanzee comparison. Mol Biol Evol 23: 565–73.

Taylor JS, Breden F (2002) The inheritance of heteroplasmy in guppies. J Fish Biol 60: 1346–50.

Taylor MV, Gusse M, Evan GI, Dathan N, Mechali M (1986) Xenopus myc proto-oncogene during development: expression as a stable maternal mRNA uncoupled from cell division. EMBO J 5: 3563-70.

Taylor PD, Wilson H, Hillier SG, Wiegand SJ, Fraser HM (2007)Effects of inhibition of vascular endothelial growth factor at time of selection on follicular angiogenesis, expansion, development and atresia in the marmoset.Mol Hum Reprod 13: 729-36.

Technau GM, Campos-Ortega JM (1986) Lineage analysis of transplanted individual cells in embryos of Drosophila melanogaster. III. Commitment and proliferative capabilities of pole cells and midgut progenitors. Roux’s Arch Dev Biol 195: 489-98.

Tedman-Aucoin K, Agrawal AF (2012) The effect of deleterious mutations and age on recombination in Drosophila melanogaster. Evolution 66:575-85.

Tegova R, Tover A, Tarassova K, Tark M, Kivisaar M (2004) Involvement of error-prone DNA polymerase IV in stationary-phase mutagenesis in Pseudomonas putida. J Bacteriol 186: 2735–44.

Telford N, Sinha YN, Finch CE (1987) Hormonal influences on the estradiol-induced and age-related increases of pituitary dopamine in C57BL/6J mice. Effects of gonadal steroids, gender, and prolactin. Neuroendocrinology 46: 481–7.

Tell G, Pellizzari L, Cimarosti D, Pucillo C, Damante G (1998a) Ref-1 controls pax-8 DNA-binding activity. Biochem Biophys Res Commun 252: 178–83.

Tell G, Scaloni A, Pellizzari L, Formisano S, Pucillo C, Damante G (1998b) Redox potential controls the structure and DNA binding activity of the paired domain. J Biol Chem 273: 25062–72.

Tell G, Zecca A, Pellizzari L, Spessotto P, Colombatti A, Kelley MR, Damante G, Pucillo C (2000) An ‘environment to nucleus’ signaling system operates in B lymphocytes: redox status modulates BSAP/Pax-5 activation through Ref-1 nuclear translocation. Nucleic Acids Res 28: 1099–105.

Tell G, Damante G, Caldwell D, Kelley MR (2005) The intracellular localization of APE1/Ref-1: More than a passive phenomenon? Antioxid Redox Signal 7: 367–84.

Tell G, Quadrifoglio F, Tiribelli C, Kelley MR (2009)The many functions of APE1/Ref-1: not only a DNA repair enzyme.Antioxid Redox Signal 11: 601-20.

Temin HM, Mizutani S (1970) RNA-dependent DNA polymerase in virions of Rous sarcoma virus. Nature 226: 1211–3.

Temin RG (1966) Homozygous viability and fertility loads in Drosphila melanogaster. Genetics 53: 27–46.

Temin RG (1978) Partial dominance of EMS-induced mutations affecting viability in Drosophila melanogaster. Genetics 89: 315-40.

Temme DH (1986) Seed size variability: A consequence of variable genetic quality among offspring? Evolution 40: 414-7.

Tempel C, Gilead A, Neeman M (2000) Hyaluronic acid as an anti-angiogenic shield in the preovulatory rat follicle. Biol Reprod 63: 134–40.

Tempel-Brami C, Neeman M (2002)Non-invasive analysis of rat ovarian angiogenesis by MRI.Mol Cell Endocrinol 187: 19–22.

Templeton AR (1996) Contingency tests of neutrality using intra/interspecific gene trees: the rejection of neutrality for the evolution of mitochondrial cytochrome oxidase II gene in the hominoid primates. Genetics 144: 1263-70.

Templeton AR, Levin DA (1979) Evolutionary consequences of seed pools. Am Nat 114: 232–49.

Templeton AR, Clark AG, Weiss KM, Nickerson DA, Boerwinkle E, Sing CF (2000) Recombinational and mutational hot spots within the human lipoprotein lipase gene. Am J Hum Genet 66: 69-83.

Ten J, Vendrell FJ, Cano A, Tarin JJ (1997) Dietary antioxidant supplementation did not affect declining sperm function with age in the mouse but did increase head abnormalities and reduced sperm production. Reprod Nutr Dev 37: 481–92.

Tenaillon MI, Hollister JD, Gaut BS (2010) A triptych of the evolution of plant transposable elements. Trends Plant Sci 15: 471–8.

Tenaillon O, Toupance B, Le Nagard H, Taddei F, Godelle B (1999)Mutators, population size, adaptive landscape and the adaptation of asexual populations of bacteria.Genetics 152:485-93.

Tenaillon O, Le Nagard H, Godelle B, Taddei F (2000)Mutators and sex in bacteria: conflict between adaptive strategies.Proc Natl Acad Sci USA 97:10465-70.

Tenaillon O, Taddei F, Radmian M, Matic I (2001)Second-order selection in bacterial evolution: selection acting on mutation and recombination rates in the course of adaptation.Res Microbiol 152:11-6.

Tenaillon O, Denamur E, Matic I (2004) Evolutionary significance of stress-induced mutagenesis in bacteria. Trends Microbiol 12:264-70.

Tenaillon O, Rodríguez-Verdugo A, Gaut RL, McDonald P, Bennett AF, et al. (2012)The molecular diversity of adaptive convergence.Science 335: 457-61.

Teng CS, Vilagrasa X (1998)Biphasic c-Myc protein expression during gossypol-induced apoptosis in rat spermatocytes.Contraception 57: 117-23.

Teng G, Papavasiliou FN (2009)Shhh! Silencing by microRNA-155.Philos Trans R Soc Lond B Biol Sci 364: 631-7.

Teodoro J, Rolo AP, Oliveira PJ, Palmeira CM (2006) Decreased ANT content in Zucker fatty rats: relevance for altered hepatic mitochondrial bioenergetics in steatosis. FEBS Lett 580: 2153–7.

Teodoro JG, Parker AE, Zhu X, Green MR (2006) p53-mediated inhibition of angiogenesis through up-regulation of a collagen prolyl hydroxylase. Science 313: 968–71.

Teotónío H, Rose MR (2000) Variation in the reversibility of evolution. Nature 408: 463–6.

Teotónío H, Rose MR (2001) Perspective: reverse evolution. Evolution 55: 653–60.

Teotónío H, Rose MR (2002) Reverse evolution of fitness in Drosophila melanogaster. J Evol Biol 15: 608–17.

Teotónio H, Chelo IM, Bradic M, Rose MR, Long AD (2009) Experimental evolution reveals natural selection on standing genetic variation. Nat Genet 41: 251–7.

Teperek-Tkacz M, Pasque V, Gentsch G, Ferguson-Smith AC (2011) Epigenetic reprogramming: is deamination key to active DNA demethylation? Reproduction 142: 621–32.

Terborgh J, Weske JS (1975) The role of competition in the distribution of Andean birds. Ecology 56: 562–76.

Terhivuo J, Saura A (2006) Dispersal and clonal diversity of North-European parthenogenetic earthworms. Biol Invasions 8: 1205–18.

Tesarik J, Guido M, Mendoza C, Greco E (1998) Human spermatogenesis in vitro: respective effects of follicle-stimulating hormone and testosterone on meiosis, spermiogenesis, and Sertoli cell apoptosis. J Clin Endocrinol Metab 83: 4467–73.

Tesarik J, Mendoza C, Anniballo R, Greco E (2000) In-vitro differentiation of germ cells from frozen testicular biopsy specimens. Hum Reprod 15: 1713–6.

Tesarik J, Martinez F, Rienzi L, Iacobelli M, Ubaldi F, et al. (2002)In-vitro effects of FSH and testosterone withdrawal on caspase activation and DNA fragmentation in different cell types of human seminiferous epithelium.Hum Reprod 17: 1811-9.

Tettelin H, Saunders NJ, Heidelberg J, Jeffries AC, Nelson KE, et al. (2000)Complete genome sequence of Neisseria meningitidis serogroup B strain MC58. Science 287: 1809-15.

te Velde AA, Pronk I, de Kort F, Stokkers PC (2008) Glutathione peroxidase 2 and aquaporin 8 as new markers for colonic inflammation in experimental colitis and inflammatory bowel diseases: an important role for H2O2? Eur J Gastroenterol Hepatol 20: 555–60.

te Velde ER, Pearson PL (2002) The variability of female reproductive ageing. Hum Reprod Update 8: 141–54.

Thai TH, Calado DP, Casola S, Ansel KM, Xiao C, et al. (2007) Regulation of the germinal center response by microRNA-155. Science 316: 604–8.

Thamotharan M, Garg M, Oak S, Rogers LM, Pan G, Sangiorgi F, et al. (2007) Transgenerational inheritance of the insulin-resistant phenotype in embryo-transferred intrauterine growth-restricted adult female rat offspring. Am J Physiol Endocrinol Metab 292: E1270–E1279.

Thannickal VJ (2009)Oxygen in the evolution of complex life and the price we pay.Am J Respir Cell Mol Biol 40: 507-10.

Thannickal VJ, Hassoun PM, White AC, Fanburg BL (1993) Enhanced rate of H2O2 release from bovine pulmonary artery endothelial cells induced by TGF-beta 1. Am J Physiol 265: L622-L626.

Thannickal VJ, Fanburg BL (2000) Reactive oxygen species in cell signaling. Am J Physiol Lung Cell Mol Physiol 279: L1005-28.

Thannickal VJ, Day RM, Klinz SG, Bastien MC, Larios JM, Fanburg BL (2000) Ras-dependent and -independent regulation of reactive oxygen species by mitogenic growth factors and TGF-beta1. FASEB J 14: 1741–8.

Thatcher JW, Shaw JM, Dickinson WJ (1998) Marginal fitness contributions of nonessential genes in yeast. Proc Natl Acad Sci U S A 95: 253-7.

Thathachar MAL, Sastry PS (2002)Varieties of learning automata: an overview.IEEE Trans Syst Man Cybern 32: 711-22.

Thattai M, van Oudenaarden A (2004) Stochastic gene expression in fluctuating environments. Genetics 167: 523–30.

Thayer KA, Ruhlen RL, Howdeshell KL, Buchanan DL, Cooke PS, et al. (2001) Altered prostate growth and daily sperm production in male mice exposed prenatally to subclinical doses of 17α-ethinyl oestradiol. Hum Reprod 16: 988–96.

The 1000 Genomes Project Consortium (2010) A map of human genome variation from population-scale sequencing. Nature 467: 1061–73.

The International Aphid Genomics Consortium (2010) Genome sequence of the pea aphid Acyrthosiphon pisum. PLoS Biol 8: e1000313.

The International HapMap Consortium (2007)A second generation human haplotype map of over 3.1 million SNPs.Nature 449: 851-61.

Theodorakis CW, Lee KL, Adams SM,Law CB (2006) Evidence of altered gene flow, mutation rate, and genetic diversity in redbreast sunfish from a pulp?mill?contaminated river. Environ Sci Technol40: 377–86.

Thibier M, Rolland A (1976) The effect of dexamethasone (DXM) on circulating testosterone (T) and luteinizing hormone (LH) in young postpubertal bulls. Theriogenology 5: 53-9.

Thieulant ML, Duval J (1985)Differential distribution of androgen and estrogen receptors in rat pituitary cell populations separated by centrifugal elutriation.Endocrinology 116: 1299-303.

Thoday JM (1953) Components of fitness. Symp Soc Exp Biol 7: 97-113.

Thoday J (1958) Natural selection and biological process. In: Barnett S, ed. A century of Darwin. London, UK: Heinemann. pp 313–33.

Thomas ALR (1997) The breath of life – did increased oxygen levels trigger the Cambrian explosion? Trends Ecol Evol 12: 44–5.

Thomas CA, Paquola AC, Muotri AR (2012)LINE-1 retrotransposition in the nervous system.Annu Rev Cell Dev Biol 28: 555-73.

Thomas F, Fisher D, Fort P, Marie J-P, Daust S, et al. (2013) Applying ecological and evolutionary theory to cancer: A long and winding road. Evol Applications 6: 1-10.

Thomas GH (1996) High male:female ratio of germ-line mutations: an alternative explanation for postulated gestational lethality in males in X-linked dominant disorders. Am J Hum Genet 58: 1364-8.

Thomas JH, Emerson RO, Shendure J (2009) Extraordinary molecular evolution in the PRDM9 fertility gene. PLoS One 4: e8505.

Thomas ML, Simmons LW (2009) Male dominance influences pheromone expression, ejaculate quality, and fertilization success in the Australian field cricket, Teleogryllus oceanicus. Behav Ecol 20: 1118–24.

Thomas P, Rahman MS, Kummer JA, Lawson S (2006) Reproductive endocrine dysfunction in Atlantic croaker exposed to hypoxia. Mar Environ Res 62: S249–S252.

Thomas P, Rahman MS, Khan IA, Kummer JA (2007) Widespread endocrine disruption and reproductive impairment in an estuarine fish population exposed to seasonal hypoxia. Proc R Soc B 274: 2693-702.

Thomas P, Rahman MS (2010) Region-wide impairment of Atlantic croaker testicular development and sperm production in the northern Gulf of Mexico hypoxic dead zone.Mar Environ Res 69 Suppl:S59-62.

Thomason P, Traynor D, Kay R (1999)Taking the plunge. Terminal differentiation in Dictyostelium.Trends Genet 15: 15-9.

Thompson AR, Chen SH (1994) Germ line origins of de novo mutations in hemophilia B families. Hum Genet 94: 299-302.

Thompson CW, Hillgarth N, Leu M, McClure HE (1997) High parasite load in house finches (Carpodacus mexicanus) is correlated with reduced expression of a sexually selected trait. Am Nat 149: 270-94.

Thompson DA, Desai MM, Murray AW (2006) Ploidy controls the success of mutators and nature of mutations during budding yeast evolution. Curr Biol 16: 1581–90.

Thompson JN (1998) Rapid evolution as an ecological process. Trends Ecol Evol 13: 329–32.

Thompson JN (1999) Specific hypotheses on the geographic mosaic of coevolution. AmNat 153 (suppl.): S1–S14.

Thompson JN (2005) The Geographic Mosaic of Coevolution. Chicago, IL: University of Chicago Press.

Thompson JN (2009) The coevolving web of life. Am Nat 173: 125–140.

Thompson JN, Cunningham BM (2002) Geographic structure and dynamics of coevolutionary selection. Nature 417: 735–8.

Thompson K, Bakker J, Bekker R (1997) The soil seed banks of North West Europe: methodology, density and longevity. Cambridge, UK: Cambridge University Press.

Thompson SL, Whitton J (2006) Patterns of recurrent evolution and geographic parthenogenesis within apomictic polyploid Easter daises (Townsendia hookeri). Mol Ecol 15: 3389–400.

Thomson JD (1989) Germination schedules of pollen grains: implications for pollen selection. Evolution 43: 220–3.

Thompson WE, Ramalho-Santos J, Sutovsky P (2003)Ubiquitination of prohibitin in mammalian sperm mitochondria: possible roles in the regulation of mitochondrial inheritance and sperm quality control.Biol Reprod 69: 254-60.

Thorne JL, Goldman N, Jones DT (1996) Combining protein evolution and secondary structure. Mol Biol Evol 13: 666–73.

Thornhill R (1983) Cryptic female choice and its implications in the scorpionfly Harpobittacus nigriceps. Am Nat 122: 765–88.

Thornhill R, Alcock J (1983) The Evolution of Insect Mating Systems. Boston, MA: Harvard University Press.

Tian B, Brasier AR (2003)Identification of a nuclear factor kappa B-dependent gene network.Recent Prog Horm Res 58: 95-130.

Tian H, Hammer RE, Matsumoto AM, Russell DW, McKnight SL (1998) The hypoxia-responsive transcription factor EPAS1 is essential for catecholamine homeostasis and protection against heart failure during embryonic development. Genes Dev 12: 3320–4.

Tiemann-Boege I, Navidi W, Grewal R, Cohn D, Eskenazi B, et al. (2002) The observed human sperm mutation frequency cannot explain the achondroplasia paternal age effect. Proc Natl Acad Sci USA 99: 14952–57.

Tijmes M, Pedraza R, Valladares L (1996)Melatonin in the rat testis: evidence for local synthesis.Steroids 61: 65–8.

Tijsterman M, May RC, Simmer F, Okihara KL, Plasterk RH (2004) Genes required for systemic RNA interference in Caenorhabditis elegans. Curr Biol 14: 111–6.

Tilbrook AJ, Turner AI, Clark IJ (2000) Effects of stress on reproduction in non-rodent mammals: the role of glucocorticoids and sex differences. Rev Reprod 5:105–113.

Tili E, Croce CM, Michaille JJ (2009)miR-155: on the crosstalk between inflammation and cancer.Int Rev Immunol 28: 264-84.

Tili E, Michaille JJ, Wernicke D, Alder H, Costinean S, et al. (2011)Mutator activity induced by microRNA-155 (miR-155) links inflammation and cancer.Proc Natl Acad Sci USA108:4908-13.

Tilly JL (2001) Commuting the death sentence: how oocytes strive to survive. Nat Rev Mol Cell Biol 2: 838–48.

Tilly JL, Kowalski KI, Johnson AL, Hsueh AJ (1991) Involvement of apoptosis in ovarian follicular atresia and postovulatory regression. Endocrinology 129: 2799–801.

Tilly JL, Tilly KI (1995) Inhibitors of oxidative stress mimic the ability of follicle stimulating hormone to suppress apoptosis in cultured rat ovarian follicles. Endocrinology 136: 242-52.

Tilly JL, Ratts VS (1996) Biological and clinical importance of ovarian cell death. Contemp Obstet Gynecol 41: 59–86.

Tilly JL, Tilly KI, Perez GI (1997) The genes of cell death and cellular susceptibility to apoptosis in the ovary: a hypothesis. Cell Death Differ 4: 180-7.

Tilman D (1982) Resource Competition and Community Structure. Princeton, NJ: Princeton University Press.

Timmermeyer N, Stelzer CP (2006) Chemical induction of mixis in the rotifer Synchaeta tremula. J Plankton Res 28: 1233–9.

Timmons L, Fire A (1998) Specific interference by ingested dsRNA. Nature 395: 854.

Tinajero JC, Fabbri A, Dufau ML (1992) Regulation of corticotropin-releasing factor secretion from Leydig cells by serotonin. Endocrinology 130: 1780-8.

Tinti F, Scali V (1996) Androgenetics and triploids from an interacting parthenogenetic hybrid and its ancestors in stick insects. Evolution 50: 1251–8.

Tippin B, Pham P, Goodman MF (2004) Error-prone replication for better or worse. Trends Microbiol 12: 288–95.

Tiravanti E, Samouilov A, Zweier JL (2004) Nitrosyl-heme complexes are formed in the ischemic heart: evidence of nitrite-derived nitric oxide formation, storage, and signaling in post-ischemic tissues. J Biol Chem 279: 11065–73.

Tischfield JA (1997) Loss of heterozygosity or: how I learned to stop worrying and love mitotic recombination. Am J Hum Genet 61: 995–9.

Tischfield SE, Keeney S (2012)Scale matters: the spatial correlation of yeast meiotic DNA breaks with histone H3 trimethylation is driven largely by independent colocalization at promoters.Cell Cycle 11: 1496-503.

Tischner R, Planchet E, Kaiser WM (2004) Mitochondrial electron transport as a source for nitric oxide in the unicellular green alga Chlorella sorokiniana. FEBS Lett 576: 151–5.

Tishkoff SA, Reed FA, Ranciaro A, Voight BF, Babbitt CC, Silverman JS, Powell K, Mortensen HM, Hirbo JB, Osman M, et al. (2007) Convergent adaptation of human lactase persistence in Africa and Europe. Nat Genet 39:31–40.

Tittensor DP, Mora C, JetzW, Lotze HK, Ricard D, Vanden Berghe E, Worm B (2010) Global patterns and predictors of marine biodiversity across taxa. Nature 466: 1098–101.

To KK, Koshiji M, Hammer S, Huang LE (2005) Genetic instability: the dark side of the hypoxic response.Cell Cycle 4:881-2.

To KK, Sedelnikova OA, Samons M, Bonner WM, Huang LE (2006) The phosphorylation status of PAS-B distinguishes HIF-1alpha from HIF-2alpha in NBS1 repression. Embo J 25: 4784–94.

Tobias JA, Seddon N (2009)Sexual selection and ecological generalism are correlated in antbirds.J Evol Biol 22: 623-36.

Tobias S (2012)Effects of glucocorticoids on vascular NADPH oxidases and endothelial NO synthase. Thesis. Mainz, Germany: Johannes Gutenberg-University Mainz

Tobler M, Schlupp I (2005) Parasites in sexual and asexual mollies (Poecilia, Poeciliidae, Teleostei): a case for the Red Queen? Biol Lett 1:166–8.

Tobler M, Schlupp I (2008) Expanding the horizon: the Red Queen and potential alternatives. Can J Zool 86: 765–73.

Tobler M, Schlupp I (2010) Differential susceptibility to food stress in neonates of sexual and asexual mollies (Poecilia, Poeciliidae). Evol Ecol 24:39–47.

Tobias (2012)

Todeschini AL, Morillon A, Springer M, Lesage P (2005) Severe adenine starvation activates Ty1 transcription and retrotransposition in Saccharomyces cerevisiae. Mol Cell Biol 25: 7459–72.

Tokuriki N, Tawfik DS (2009) Stability effects of mutations and protein evolvability. Curr Opin StructBiol 19:596–604.

Tolando R, Jovanovic A, Brigelius-Flohe R, Ursini F, Maiorino M (2000) Reactive oxygen species and proinflammatory cytokine signaling in endothelial cells: effect of selenium supplementation. Free Radic Biol Med 28: 979-86.

Tomanek L, Somero GN (1999) Evolutionary and acclimation-induced variation in the heat-shock responses of congeneric marine snails (genus Tegula) from different thermal habitats: implications for limits of thermotolerance and biogeography. J Exp Biol 202: 2925–36.

Tomita K, Barnes PJ, Adcock IM (2003)The effect of oxidative stress on histone acetylation and IL-8 release.Biochem Biophys Res Commun 301: 572-7.

Tomiuk J, Loeschcke V (1992)Evolution of parthenogenesis in the Otiorhynchus scaber complex. Heredity 68: 391–8.

Tomkins JL, Radwan J, Kotiaho JS, Tregenza T (2004) Genic capture and resolving the lek paradox. Trends Ecol Evol 19: 323–8.

Tomko RJ Jr, Bansal P, Lazo JS (2006)Airing out an antioxidant role for the tumor suppressor p53.Mol Interv 6: 23-5.

Tomkowiak E, Pienkowska JR (2010) The current knowledge of invertebrate aquaporin water channels with particular emphasis on insect AQPS. Adv Cell Biol 1-14.

Tomlinson JW, Walker EA, Bujalska IJ, Draper N, Lavery GG, et al. (2004) 11β-Hydroxysteroid dehydrogenase type 1: a tissue-specific regulator of glucocorticoid response. Endocr Rev 25: 831–66.

Tomoyasu Y, Miller SC, Tomita S, Schoppmeier M, Grossmann D, Bucher G (2008) Exploring systemic RNA interference in insects: a genome-wide survey for RNAi genes in Tribolium. Genome Biol 9: R10.

Tonegawa S (1976) Reiteration frequency of immunoglobulin light chain genes: further evidence for somatic generation of antibody diversity. Proc Natl Acad Sci USA 73: 203-7.

Tonegawa S (1983) Somatic generation of antibody diversity. Nature 302: 575–81.

Tonn WM, Magnuson JJ (1982) Patterns in the species composition and richness of fish assemblages in northern Wisconsin lakes. Ecology 1149-66.

Torgovnick A, Schiavi A, Testi R, Ventura N (2010)A role for p53 in mitochondrial stress response control of longevity in C. elegans.Exp Gerontol 45: 550-7.

Torkelson J,Harris RS, Lombardo MJ, Nagendran J, Thulin C, Rosenberg SM (1997) Genome-wide hypermutation in a subpopulation of stationary-phase cells underlies recombination-dependent adaptive mutation. EMBO J 16: 3303–11.

Torley KJ, da Silveira JC, Smith P, Anthony RV, Veeramachaneni DN, et al. (2011)Expression of miRNAs in ovine fetal gonads: potential role in gonadal differentiation.Reprod Biol Endocrinol 9: 2.

Török J, Hegyi G, Tóth L, Könczey R (2004)Unpredictable food supply modifies costs of reproduction and hampers individual optimization.Oecologia 141: 432-43.

Torres R, Velando A (2007) Male reproductive senescence: the price of immune induced oxidative damage on sexual attractiveness in the blue-footed booby. J Anim Ecol 76: 1161–8.

Tortosa P, Dubnau D (1999) Competence for transformation: a matter of taste. Curr Opin Microbiol 2: 588-92.

Toth G, Gaspari Z, Jurka J (2000) Microsatellites in different eukaryotic genomes: survey and analysis. Genome Res 10: 967-81.

Touati D, Jacques M, Tardat B, Bouchard L, Despied S (1995) Lethal oxidative damage and mutagenesis are generated by iron in delta-fur mutants of Escherichia coli: protective role of superoxide dismutase. J Bacteriol 177: 2305-14.

Townsend CR, Hildrew AG (1994) Species traits in relation to a habitat templet for river systems. Freshwater Biol 31: 265–75.

Toyokuni S (2008) Molecular mechanisms of oxidative stress-induced carcinogenesis: from epidemiology to oxygenomics. IUBMB Life 60: 441-7.

Tracey KJ, Beutler B, Lowry SF, Merryweather J, Wolpe S, et al. (1986) Shock and tissue injury induced by recombinant human cachectin. Science 234: 470-4.

Tracey KJ, Lowry SF, Fahey TJ 3rd, Albert JD, Fong Y, et al. (1987) Cachectin/tumor necrosis factor induces lethal shock and stress hormone responses in the dog. Surg Gynecol Obstet 164: 415-22.

Trachootham D, Lu W, Ogasawara MA, Nilsa RD, Huang P (2008) Redox regulation of cell survival. Antioxid Redox Signal 10: 1343-74.

Trakshel GM, Kutty K, Maines MD (1986) Purification and characterization of the major constitutive form of testicular heme oxygenase. J Biol Chem 261: 11131–7.

Tramontano F, Malanga M, Farina B, Jones R, Quesada P (2000) Heat stress reduces poly(ADPR) polymerase expression in rat testis. Mol Hum Reprod 6: 575–81.

Tran HT, Keen JD, Kricker M, Resnick MA, Gordenin DA (1997) Hypermutability of homonucleotide runs in mismatch repair and DNA polymerase proofreading yeast mutants. Mol Cell Biol 17: 2859–65.

Trappe R, Laccone F, Cobilanschi J, Meins M, Huppke P, et al. (2001) MECP2 mutations in sporadic cases of Rett syndrome are almost exclusively of paternal origin. Am J Hum Genet 68: 1093–101.

Travis J (1996) The significance of geographical variation in species interactions. Am Nat 148 (suppl.): S1–S8.

Travis J, Keen WH, Julianna J (1985) The effects of multiple factors on viability selection in Hyla gratiosa tadpoles. Evolution 39: 1087–99.

Travis JM, Travis ER (2002) Mutator dynamics in fluctuating environments. Proc Biol Sci 269: 591–7.

Travisano M, Mongold JA, Bennet AF, Lenski RE (1995) Experimental tests of the roles of adaptation, chance, and history in evolution. Science 267: 87–90.

Treangen TJ, Ambur OH, Tonjum T, Rocha EP (2008) The impact of the neisserial DNA uptake sequences on genome evolution and stability. Genome Biol 9: R60.

Treco D, Arnheim N (1986) The evolutionary conserved repetitive sequence d(TG·AC)n promotes reciprocal exchange and generates unusual recombinant tetrads during yeast meiosis. Mol Cell Biol 6: 3934-47.

Tregenza T, Wedell N (1998) Benefits of multiple mates in the cricket Gryllus bimaculatus. Evolution Int J Org Evolution 52: 1726–30.

Tregenza T, Wedell N (2000) Genetic compatibility, mate choice and patterns of parentage: invited review. Mol Ecol 9: 1013–27.

Tregenza T, Wedell N, Chapman T (2006) Introduction. Sexual conflict: a new paradigm? Philos Trans R Soc Lond B Biol Sci 361: 229-34.

Treinin M, Simchen G (1993) Mitochondrial activity is required for expression of IME1, a regulator of meiosis in yeast. Curr. Genet 23: 223-7.

Trejo R, Valadez-Salazar A, Delhumeau G (1995) Effects of quercetin on rat testis aerobic glycolysis. Can J Physiol Pharmacol 73: 1605–15.

Trelogan SA, Martin SL (1995) Tightly regulated, developmentally specific expression of the first open reading frame from LINE-1 during mouse embryogenesis. Proc Natl Acad Sci USA 92: 1520–4.

Tremblay MF, Bergeron Y, Lalonde D, Mauffette Y (2002) The potential effects of sexual reproduction and seedling recruitment on the maintenance of red maple (Acer rubrum L.) populations at the northern limit of the species range. J Biogeogr 29: 365–73.

Tremellen K (2008) Oxidative stress and male infertility—a clinical perspective. Hum Reprod Update 14: 243–58.

Tres LL, Smith EP,VanWyk JJ, Kierszenbaum AL (1986) Immunoreactive sites and accumulation of somatomedin-C in rat Sertoli-spermatogenic cell co-cultures. Exp Cell Res 162: 33–50.

Tres LL, Kierszenbaum AL (1999) Cell death patterns of the rat spermatogonial cell progeny induced by Sertoli cell geometric changes and Fas (CD95) agonist. Dev Dyn 214: 361–71.

Trifonov EN (2003) Tuning function of tandemly repeating sequences: a molecular device for fast adaptation. In: Wasser SP, ed. Evolutionary theory and processes: modern horizons, papers in honor of Eviatar Nevo. Amsterdam, The Netherlands: Kluwer Academic Publishers. pp 1–24.

Trinei M, Giorgio M, Cicalese A, Barozzi S, Ventura A, et al. (2002) A p53-p66Shc signalling pathway controls intracellular redox status, levels of oxidation-damaged DNA and oxidative stress-induced apoptosis.Oncogene 21:3872-8.

Trinh DL, Elwi AN, Kim SW (2010) Direct interaction between p53 and Tid1 proteins affects p53 mitochondrial localization and apoptosis. Oncotarget 1: 396-404.

Tripathi R, Mishra DP, Shaha C (2009) Male germ cell development: turning on the apoptotic pathways. J Reprod Immunol 83: 31-5.

Trivers RL (1974) Parent-offspring conflict. Am Zool 14: 249–64.

Trobner W, Piechocki R (1981) Competition growth between Escherichia coli mutl and mut+ in continuously growing cultures. Z Allg Mikrobiol 21: 347–9.

Tropea A, Miceli F, Minici F, Tiberi F, Orlando M, et al. (2006) Regulation of vascular endothelial growth factor synthesis and release by human luteal cells in vitro. J Clin Endocrinol Metab 91: 2303–9.

Trougakos IP, Margaritis LH (2002) Novel morphological and physiological aspects of insect eggs. In: Hilker M, Meiners T, eds. Chemoecology of Insect Eggs and Egg Deposition. Berlin, Germany: Blackwell Wissenschaftsverlag. pp 3-36.

Troy CM, Shelanski ML (1994) Down-regulation of copper/zinc superoxide dismutase causes apoptotic death in PC12 neuronal cells. Proc Natl Acad Sci USA 91: 6384-7.

True HL, Lindquist SL (2000) A yeast prion provides a mechanism for genetic variation and phenotypic diversity. Nature 407: 477–83.

True HL, Berlin I, Lindquist SL (2004) Epigenetic regulation of translation reveals hidden genetic variation to produce complex traits. Nature 431: 184–7.

Tsai NP, Wei LN (2010) RhoA/ROCK1 signaling regulates stress granule formation and apoptosis. Cell Signal 22: 668-75.

Tsai SC, Lu CC, Lin CS, Wang PS (2003) Antisteroidogenic actions of hydrogen peroxide on rat Leydig cells. J Cell Biochem 90: 1276-86.

Tsai-Turton M, Luderer U (2006) Opposing effects of glutathione depletion and follicle stimulating hormone on reactive oxygen species and apoptosis in cultured preovulatory rat follicles. Endocrinology 147:1224–36.

Tsaousis AD, Martin DP, Ladoukakis ED, Posada D, Zouros E (2005) Widespread recombination in published animal mtDNA sequences. Mol Biol Evol 22: 925-33.

Tsaur SC, Wu CI (1997) Positive selection and the molecular evolution of a gene of male reproduction, Acp26Aa of Drosophila. Mol Biol Evol 14: 544–9.

Tsaur SC, Ting CT, Wu CI (1998) Positive selection driving the evolution of a gene of male reproduction, Acp26Aa, of Drosophila. II. Divergence versus polymorphism. Mol Biol Evol 15: 1040–6.

Tsaousis AD, Martin DP, Ladoukakis ED, Posada D, Zouros E (2005) Widespread recombination in published animal mtDNA sequences. Mol Biol Evol 22: 925–33.

Tsilfidis C, MacKenzie AE, Mettler G, Barcelo J, Korneluk RG (1992) Correlation between CTG trinucleotide repeat length and frequency of severe congenital myotonic dystrophy. Nat Genet 1: 192-5.

TsimringLS, Levine H, Kessler DA (1996) RNA virus evolution via a fitness-space model. Phys Rev Lett 76: 4440–3.

Tsujimoto M, Yokota S, Vilcek J, Weissmann G (1986) Tumor necrosis factor provokes superoxide anion generation from neutrophils. Biochem Biophys Res Commun 137: 1094-100.

Tsutsui K, Ubuka T, Bentley GE, Kriegsfeld LJ (2013) Review: regulatory mechanisms of gonadotropin-inhibitory hormone (GnIH) synthesis and release in photoperiodic animals. Front Neurosci 7: 60.

Tu Y, Tornaletti S, Pfeifer GP (1996) DNA repair domains within a human gene: selective repair of sequences near the transcription initiation site. EMBO J 15: 675–83.

Tu Z, Wang L, Xu M, Zhou X, Chen T, Sun F (2006) Further understanding human disease genes by comparing with housekeeping genes and other genes. BMC Genomics 7: 31.

Tucci P (2012) Caloric restriction: is mammalian life extension linked to p53? Aging (Albany NY) 4: 525-34.

Tuchman M, Matsuda I, Munnich A, Malcolm S, Strautnieks S, Briede T (1995) Proportions of spontaneous mutations in males and females with ornithine transcarbamylase deficiency. Am J Med Genet 55: 67-70.

Tully T, Ferrière R (2008) Reproductive flexibility: genetic variation, genetic costs and long-term evolution in a Collembola. PLoS ONE 3: e3207.

Tunc O, Tremellen K (2009) Oxidative DNA damage impairs global sperm DNA methylation in infertile men. J Assist Reprod Genet 26: 537-44.

Tunnacliffe A, Lapinski J (2003) Resurrecting Van Leeuwenhoek’s rotifers: a reappraisal of the role of disaccharides in anhydrobiosis. Phil Trans R Soc B 358: 1755–71.

Tunner HG, Nopp H (1979) Heterosis in the common European water frog. Naturwissenschaften 66: 268–9.

Turak E, Sunnucks P, Hales DF (1998) Different responses to temperature in three closely-related parthenogenetic cereal aphids. Entomol Exp Appl 86: 49–58.

Turelli M (1984) Heritable genetic variation via mutation-selection balance: Lerch’s Zeta meets the abdominal bristle. Theor Popul Biol 25: 138–193.

Turgeon J, Hebert PN (1995) Genetic characterization of breeding systems, ploidy levels and species boundaries in Cypricercus (Ostracoda). Heredity 75: 561–70.

Turini ME, DuBois RN (2002) Cyclooxygenase-2: a therapeutic target. Annu Rev Med 53: 35–57.

Turpaev KT (2002)Reactive oxygen species and regulation of gene expression.Biochemistry (Mosc) 67: 281-92.

Turk PW, Laayoun A, Smith SS, Weitzman SA (1995) DNA adduct 8-hydroxyl-2'-deoxyguanosine (8-hydroxyguanine) affects function of human DNA methyltransferase. Carcinogenesis 16: 1253-5.

Turk PW, Weitzman SA (1995) Free radical DNA adduct 8-OH-deoxyguanosine affects activity of HPA II and MSP I restriction endonucleases. Free Radic Res 23: 255–8.

Türker KI, Erdogru T, Gülkesen H, Sezer C, Usta M, et al. (2004) The potential role of inducible nitric oxide synthase (iNOS) activity in the testicular dysfunction associated with varicocele: an experimental study. Int Urol Nephrol 36: 67–72.

Turner BJ, Balsano JS, Monaco PJ, Rash EM (1983) Clonal diversity and evolutionary dynamics in a diploid-triploid breeding complex of unisexual fishes (Poecilia). Evolution 37: 798–809.

Turner BM (2002) Cellular memory and the histone code. Cell 111: 285-91.

Turner BM (2009) Epigenetic responses to environmental change and their evolutionary implications. Philos Trans R Soc Lond B 364: 3403–18.

Turner N, Else PL, Hulbert AJ (2003) Docosahexaenoic acid (DHA) content of membranes determines molecular activity of the sodium pump: implications for disease states and metabolism. Naturwissenschaften 90: 521–3.

Turner N, Haga KL, Hulbert AJ, Else PL (2005a) Relationships between body size, Na+-K+-ATPase activity, and membrane lipid composition in mammal and bird kidney. Am J Physiol Regul Integr Comp Physiol 288: 301–10.

Turner N, Hulbert AJ, Else PL (2005b) Sodium pump molecular activity and membrane lipid composition in two disparate ectotherms and comparison with endotherms. J Comp Physiol B 175: 77–85.

Turner TT (2001) The study of varicocele through the use of animal models. Hum Reprod Update 7:78–84.

Turner TT, Caplis L, Miller DW (1996) Testicular microvascular blood flow: alteration after Leydig cell eradication and ischemia but not experimental varicocele. J Androl 17: 239–48.

Turner TT, Tung KS, Tomomasa H, Wilson LW (1997) Acute testicular ischemia results in germ cell-specific apoptosis in the rat. Biol Reprod 57: 1267–74.

Turner TT, Bang HJ, Lysiak JJ (2005) Experimental testicular torsion: reperfusion blood flow and subsequent testicular venous plasma testosterone concentrations. Urology 65: 390-4.

Turner TT, Lysiak JJ (2008) Oxidative stress: a common factor in testicular dysfunction. J Androl 29:488-98.

Turney PD (1996) Myths and legends of the Baldwin effect. In: Proc 13th Int Conf Mach Learning, Bari, Italy. pp 135–42.

Turney P, Whitley D, Anderson RW (1996) Evolution, learning, and instinct: 100 years of the Baldwin Effect. Evol Comput 4: 4-8.

Turney PD (1999) Increasing evolvability considered as a large scale trend in evolution. In: Marrow P, Shackleton M, Fernandez-Villacanas JL, Ray T, eds. Proceedings of the 1999 Genetic and Evolutionary Computation Conference (GECCO-99) Workshop Program. Orlando, FL. pp 43–46.

Turrens JF (2003) Mitochondrial formation of reactive oxygen species. J Physiol 552: 335-44.

Turrens JF, Alexandre A, Lehninger AL (1985) Ubisemiquinone is the electron donor for superoxide formation by complex III of heart mitochondria. Arch Biochem Biophys 237: 408–14.

Twig G, Hyde B, Shirihai OS (2008) Mitochondrial fusion, fission and autophagy as a quality control axis: the bioenergetic view. Biochim Biophys Acta 1777: 1092-7.

Twigg J, Fulton N, Gomez E, Irvine DS, Aitken RJ (1998) Analysis of the impact of intracellular reactive oxygen species generation on the structural and functional integrity of human spermatozoa: lipid peroxidation, DNA fragmentation and effectiveness of antioxidants. Hum Reprod 13: 1429-36.

Tyedmers J, Madariaga ML, Lindquist S (2008) Prion dwitching in response to environmental stress. PLoS Biol 6: e294.

Tyler DD (1975) Polarographic assay and intracellular distribution of superoxide dismutase in rat liver. Biochem J 147: 493–504.

Tyurina YY, Shvedova AA, Kawai K, Tyurin VA, Kommineni C, et al. (2000) Phospholipid signaling in apoptosis: peroxidation and externalization of phosphatidylserine.Toxicology 148: 93-101.

Tyurina YY, Kawai K, Tyurin VA, Liu SX, Kagan VE, Fabisiak JP (2004) The plasma membrane is the site of selective phosphatidylserine oxidation during apoptosis: role of cytochrome C. Antioxid Redox Signal 6: 209-25.

Udipi S, Ghugre P, Gokhale C (2012) Iron, oxidative stress and health. In: Lushchak V, Semchyshyn HM, eds. Oxidative stress: molecular mechanisms and biological effects. Rijeka, Croatia: InTech. pp 73-108.

Ueda S, Masutani H, Nakamura H, Tanaka T, Ueno M, Yodoi J (2002) Redox control of cell death. Antioxid Redox Signal 4: 405-14.

Ueno M, Masutani H, Arai RJ, Yamauchi A, Hirota K, et al. (1999) Thioredoxin-dependent redox regulation of p53-mediated p21 activation. J Biol Chem 274: 35809–15.

Uesugi R, Nishihiro J, Tsumura Y, Washitani I (2007) Restoration of genetic diversity from soil seed banks in a threatened aquatic plant, Nymphoides peltata. Conserv Genet 8: 111–21.

Uetani N, Yamaura K, Sato K (1994) Expression in situ of c-myc mRNA and c-myc protein during spermatogenesis in the adult mouse. Cell Biol Int 18: 85-7.

Uhl G, Nessler SH, Schneider J (2010) Securing paternity in spiders? A review on occurrence and effects of mating plugs and male genital mutilation. Genetica 138: 75–104.

Ujvari B, Dowton M, Madsen T (2007) Mitochondrial DNA recombination in a free-ranging Australian lizard. Biol Lett 3: 189–92.

Ulisse S, Fabbri A, Dufau ML (1989) Corticotropin-releasing factor receptors and actions in the rat Leydig cell. J Biol Chem 264: 2156-63.

Ulisse S, Fabbri A, Dufau ML (1990) A novel mechanism ofaction of corticotropin releasing factor in rat Leydig cells. J Biol Chem 265: 1964-71.

Uller T, Eklöf J, Andersson S (2005) Female egg investment in relation to male sexual traits and the potential for transgenerational effects in sexual selection. Behav Ecol Sociobiol 57: 584–90.

Umar A, Risinger JI, Glaab WE, Tindall KR, Barrett JC, Kunkel TA (1998) Functional overlap in mismatch repair by human MSH3 and MSH6. Genetics 148: 1637-46.

Underwood EM, Caulton JH, Allis CD, Mahowald AP (1980) Developmental fate of pole cells in Drosophila melanogaster. Dev Biol 77: 303-14.

Ungerer MJ, Ayres MP, Lombardero MJ (1999) Climate and the northern distribution limits of Dendroctonus frontalis Zimmermann (Coleoptera: Scolytidae). J Biogeogr 26: 1133–45.

Uniacke J, Zerges W (2008) Stress induces the assembly of RNA granules in the chloroplast of Chlamydomonas reinhardtii. J Cell Biol 182: 641–6.

Updike D, Strome S (2010) P granule assembly and function in Caenorhabditis elegans germ cells. J Androl 31: 53–60.

Upton KR, Baillie JK, Faulkner GJ (2011) Is somatic retrotransposition a parasitic or symbiotic phenomenon? Mob Genet Elements 1: 279–82.

Urios A, Herrera G, Blanco M (1995) Detection of oxidative mutagens in strains of Escherichia coli deficient in the oxyR or mutY functions: dependence on SOS mutagenesis. Mutat Res 332: 9-15.

Urner F, Sakkas D (2004) Involvement of the pentose phosphate pathway and redox regulation in fertilization in the mouse. Mol Reprod Dev 70:494–503.

Urrutia AO, Hurst LD (2003) The signature of selection mediated by expression on human genes. Genome Res 13: 2260-4.

Urushihara H (1992) Sexual development of cellular slime molds. Dev Growth Differ 34: 1-17.

Urushihara H, Muramoto T (2006) Genes involed in Dictyostelium discoideum sexual reproduction. Eur J Cell Biol 85: 961–8.

Ushijima T, Watanabe N, Okochi E, Kaneda A, Sugimura T, Miyamoto K (2003) Fidelity of the methylation pattern and its variation in the genome. Genome Res 13: 868–74.

Ushio-Fukai M, Tang Y, Fukai T, Dikalov SI, Ma Y, et al. (2002) Novel role of gp91(phox)-containing NAD(P)H oxidase in vascular endothelial growth factor-induced signaling and angiogenesis. Circ Res 91: 1160–7.

Uyenoyama MK (1984) On the evolution of parthenogenesis: A genetic representation of the ‘‘cost of meiosis’’. Evolution 38:87–102.

Vacquier VD (1998) Evolution of gamete recognition proteins. Science 281: 1995–8.

Vagin VV, Klenov MS, Kalmykova AI, Stolyarenko AD, Kotelnikov RN, Gvozdev VA (2004) The RNA interference proteins and vasa locus are involved in the silencing of retrotransposons in the female germline of Drosophila melanogaster. RNA Biol 1: 54–8.

Vagin VV, Sigova A, Li C, Seitz H, Gvozdev V, Zamore PD (2006) A distinct small RNA pathway silences selfish genetic elements in the germline. Science 313: 320–4.

Vaissière T, Sawan C, Herceg Z (2008) Epigenetic interplay between histone modifications and DNA methylation in gene silencing. Mutat Res 659:40-8.

Valdez LB, Zaobornyj T, Alvarez S, Bustamante J, Costa LE, Boveris A (2004) Heart mitochondrial nitric oxide synthase. Effects of hypoxia and aging. Mol Aspects Med 25: 49–59.

Vale W, Spiess J, Rivier C, Rivier J (1981) Characterization of a 41-residue ovine hypothalamic peptide that stimulates secretion of corticotropin and beta-endorphin. Science 213: 1394-7.

Valena S, Moczek AP (2012) Epigenetic mechanisms underlying developmental plasticity in horned beetles. Genet Res Int 2012: 576303.

Valenti S, Cuttica CM, Fazzuoli L, Giordano G, Giusti M (1999) Biphasic effect of nitric oxide on testosterone and cyclic GMP production by purified rat Leydig cells cultured in vitro. Int J Androl 22: 336-41.

Valenti S, Fazzuoli L, Giordano G, Giusti M (2001) Changes in binding of iodomelatonin to membranes of Leydig steroidogenesis after prolonged in vitro exposure to melatonin. Int J Androl 24: 80–6.

Valentine JW (1995) Why no new phyla after the Cambrian? Genome and ecospace hypotheses revisited. Palaios 10: 190–4.

Valentine JW, Erwin DH, Jablonski D (1996) Developmental evolution of metazoan bodyplans: the fossil evidence. Dev Biol 173: 373-81.

Valentine JW, Jablonski D, Erwin DH (1999) Fossils, molecules and embryos: new perspectives on the Cambrian explosion. Development 126: 851–9.

Valeri N, Gasparini P, Fabbri M, Braconi C, Veronese A, et al. (2010a) Modulation of mismatch repair and genomic stability by miR-155. Proc Natl Acad Sci USA 107: 6982–7.

Valeri N, Gasparini P, Braconi C, Paone A, Lovat F, et al. (2010b) MicroRNA-21 induces resistance to 5-fluorouracil by down-regulating human DNA MutS homolog 2 (hMSH2). Proc Natl Acad Sci USA 107: 21098-103.

Valinluck V, Tsai HH, Rogstad DK, Burdzy A, Bird A, Sowers LC (2004) Oxidative damage to methyl-CpG sequences inhibits the binding of the methyl-CpG binding domain (MBD) of methyl- CpG binding protein 2 (MeCP2). Nucleic Acids Res 32: 4100–8.

Valiunas V, Polosina YY, Miller H, Potapova IA, Valiuniene L, et al. (2005) Connexin-specific cell-to-cell transfer of short interfering RNA by gap junctions. J Physiol 568: 459–68.

Valko M, Izakovic M, Mazur M, Rhodes CJ, Telser J (2004) Role of oxygen radicals in DNA damage and cancer incidence. Mol Cell Biochem 266: 37-56.

Valko M, Leibfritz D, Moncol J, Cronin MTD, Mazur M, Tesler J (2007) Free radicals and antioxidant in normal physiological functions and human disease. Int J Biochem Cell Biol 39: 44–84.

Vallino M, Zampieri E, Murat C, Girlanda M, Picarella S, et al. (2011) Specific regions in the Sod1 locus of the ericoid mycorrhizal fungus Oidiodendron maius from metal-enriched soils show a different sequence polymorphism. FEMS Microbiol Ecol 75:321-31.

Vamosi SM (2005) On the role of enemies in divergence and diversification of prey: a review and synthesis. Can J Zool 83:894–910.

van Baalen M, Jansen VAA (2001) Dangerous liaisons: the ecology of private interest and common good. Oikos 95: 211-24.

Van Blerkom J (1998) Epigenetic influences on oocyte developmental competence: perifollicular vascularity and intrafollicular oxygen. J Assist Reprod Genet 15: 226-34.

Van Blerkom J (2000) Intrafollicular influences on human oocyte developmental competence: perifollicular vascularity, oocyte metabolism and mitochondrial function. Hum Reprod 15 Suppl 2: 173-88.

Van Blerkom J (2004) Mitochondria in human oogenesis and preimplantation embryogenesis: Engines of metabolism, ionic regulation and developmental competence. Reproduction 128: 269–80.

Van Blerkom J (2008) Mitochondria as regulatory forces in oocytes, preimplantation embryos and stem sells. Reprod Biomed Online 16: 553–69.

Van Blerkom J, Davis P, Lee J (1995) ATP content of human oocytes and developmental potential and outcome after in-vitro fertilization and embryo transfer. Hum Reprod 10: 415–24.

Van Blerkom J, Antczak M, Schrader R (1997) The developmental potential of the human oocyte is related to the dissolved oxygen content of follicular fluid: association with vascular endothelial growth factor levels and perifollicular blood flow characteristics. Hum Reprod 12: 1047-55.

Van Blerkom J, Davis P, Alexander S (2003) Inner mitochondrial membrane potential (ΔΨm), cytoplasmic ATP content and free Ca2+ levels in metaphase II mouse oocytes. Hum Reprod 18: 2429–40.

Van Breusegem F, Dat JF (2006) Reactive oxygen species in plant cell death. Plant Physiol141: 384–90.

Van Buskirk J, McCollum SA, Werner EE (1997) Natural selection for environmentally induced phenotypes in tadpoles. Evolution 51: 1983–92.

Van Buskirk J, Relyea RA (2008) Selection for phenotypic plasticity in Rana sylvatica tadpoles. Biol J Linn Soc 65: 301-28.

Van Buskirk J, Steiner UK (2009) The fitness costs of developmental canalization and plasticity. J Evol Biol 22: 852–60.

Vandekerkhove J, Martens K, Rossetti G, Mesquita-Joanes F, Namiotko T (2013) Extreme tolerance to environmental stress of sexual and parthenogenetic resting eggs of Eucypris virens (Crustacea, Ostracoda). Freshwater Biol 58: 237–47.

Vandekerckhove TTM, Willems A, Gillis M, Coomans A (2000) Occurrence of novel verrucomicrobial species, endosymbiotic and associated with parthenogenesis in Xiphinema americanum-group species (Nematoda, Longidoridae). Int J Syst Evol Microbiol 50: 2197–205.

van de Lagemaat LN, Landry JR, Mager DL, Medstrand P (2003) Transposable elements in mammals promote regulatory variation and diversification of genes with specialized functions. Trends Genet 19: 530–6.

Vandel A (1928) La parthénogénèse géographique. Contribution à l’étude biologique et cytologique de la parthénogénèse naturelle. Bull Biol France Belg 62: 164–281.

Vandel A (1940) La parthénogénèse géographique. IV. Polyploidie et distribution géographique. Bull Biol France Belg 74: 94–l00.

van Delden W, Beardmore JA (1968)Effects of small increments of genetic variability in inbred populations of Drosophila melanogaster.Mutat Res 6:117-27.

VanDemark NL, Free MJ(1970) Temperature effects. In: Johnson AD, Gomes WR, VanDemark NL, eds. The Testis, vol. III. New York, NY: Academic Press.pp 233–312.

Vandenbroucke K, Robbens S, Vandepoele K, Inzé D, Van de Peer Y, Van Breusegem F (2008)Hydrogen peroxide-induced gene expression across kingdoms: a comparative analysis.Mol Biol Evol 25:507-16.

Van de Pol M, Bruinzeel LW, Heg D, Van der Jeugd HP, Verhulst S (2006)A silver spoon for a golden future: long-term effects of natal origin on fitness prospects of oystercatchers (Haematopus ostralegus).J Anim Ecol 75: 616-26.

Vandeputte M, Dupont-Nivet M, Chavanne H, Chatain B (2007) A polygenic hypothesis for sex determination in the European sea bass—Dicentrarchus labrax. Genetics 176: 1049–57.

van der Geest LP, Elliot SL, Breeuwer JA, Beerling EA (2000) Diseases of mites. Exp Appl Acarol 24: 497-560.

Vander Heiden MG, Cantley LC, Thompson CB (2009)Understanding the Warburg effect: the metabolic requirements of cell proliferation.Science 324: 1029-33.

van derHeijen GW, Bortvin A (2009) Transient relaxation of transposon silencing at the onset of mammalian meiosis. Epigenetics 4: 76–9.

van der Laan R, Baarends WM, Wassenaar E, Roest HP, Hoeijmakers JH, Grootegoed JA (2005)Expression and possible functions of DNA lesion bypass proteins in spermatogenesis.Int J Androl 28: 1-15.

van der Laan S, Meijer OC (2008)Pharmacology of glucocorticoids: beyond receptors.Eur J Pharmacol 585: 483-91.

van der Spek PJ, Visser CE, Hanaoka F, Smit B, Hagemeijer A, Bootsma D, Hoeijmakers JHJ (1996) Cloning, comparative mapping and RNA expression of the mouse homologues of the Saccharomyces cerevisiae nucleotide excision repair gene RAD23. Genomics 31: 20-7.

Van der Veken S, Rogister J, Verheyen K, Hermy M, Nathan R (2007) Over the (range) edge: a 45-year transplant experiment with the perennial forest herb Hyacinthoides non-scripta. J Ecol 95: 343–51.

van der Veen S, van Schalkwijk S, Molenaar D, de Vos WM, Abee T, Wells-Bennik MHJ (2010) The SOS response of Listeria monocytogenes is involved in stress resistance and mutagenesis. Microbiology 156: 374–84.

Vander Werf E (1992) Lack's clutch size hypothesis: an examination of the evidence using meta-analysis. Ecology 73: 1699-705.

van der Woude MW, Bäumler AJ (2004) Phase and antigenic variation in bacteria. Clin Microbiol Rev 17: 581-611.

Van Dijck P, Colavizza D, Smet P, Thevelein JM (1995) Differential importance of trehalose in stress resistance in fermenting and nonfermenting Saccharomyces cerevisiae cells. Appl Environ Microbiol 61: 109-15.

Van Dijck P, Ma P, Versele M, Gorwa MF, Colombo S, et al. (2000)A baker's yeast mutant (fil1) with a specific, partially inactivating mutation in adenylate cyclase maintains a high stress resistance during active fermentation and growth.J Mol Microbiol Biotechnol 2: 521-30.

Van Dijk PJ (2007) Potential and realized costs of sex in dandelions, Taraxacum officinale s.l. In: Hörandl E, Grossniklaus U, van Dijk P, Sharbel T, eds. Apomixis: Evolution, Mechanisms and Perspectives. Rugell, Liechtenstein: Gantner Verlag. pp 215–33.

Van Doninck K, Schön I, De Bruyn L, Martens K (2002) A general purpose genotype in an ancient asexual. Oecologia 132: 205–12.

Van Doninck K, Schön I, Martens K, Godderis B (2003a) The life cycle of the ancient asexual ostracod Darwinula stevensoni (Brady and Robertson, 1870) (Crustacea, Ostracoda) in a temperate pond. Hydrobiologia 500: 331–40.

Van Doninck K, Schön I, Maes F, De Bruyn L, Martens K (2003b) Ecological strategies in the ancient asexual animal group Darwinulidae. Freshwater Biol 48: 1285–94.

van Doorn GS, Edelaar P, Weissing FJ (2009)On the origin of species by natural and sexual selection.Science 326: 1704-7.

Van Ex F, JacobY, Martienssen RA (2011)Multiple roles for small RNAs during plant reproduction. Curr Opin Plant Biol 14: 588–93.

Van Haaster LH, De Jong FH, Docter R, De Rooij DG (1992) The effect of hypothyroidism on Sertoli cell proliferation and differentiation and hormone levels during testicular development in the rat. Endocrinology 131: 1574-6.

Van Haaster LH, De Jong FH, Docter R, De Rooij DG (1993) High neonatal triiodothyronine levels reduce the period of Sertoli cell proliferation and accelerate tubular lumen formation in the rat testis, and increase serum inhibin levels. Endocrinology 133: 755–60.

Van Houten B (1990)Nucleotide excision repair in Escherichia coli. Microbiol Rev 54: 18-51.

van Hulten M, Pelser M, van Loon LC, Pieterse CMJ, Ton J (2006) Costs and benefits of priming for defense in Arabidopsis. Proc Natl Acad Sci USA 103: 5602–7.

van Kleunen M, Fischer M, Schmid B (2001) Effects of intraspecific competition on size variation and reproductive allocation in a clonal plant. Oikos 94: 515–24.

van KleunenM, Fischer M (2003) Effects of four generations of density-dependent selection on life history traits and their plasticity in a clonally propagated plant. J Evol Biol 16: 474–84.

van Kleunen M, Fischer M (2007) Progress in the detection of costs of phenotypic plasticity. New Phytol 176: 727–30.

Van Loon AA, Den Boer PJ, Van der Schans GP, Mackenbach P, Grootegoed JA, et al. (1991)Immunochemical detection of DNA damage induction and repair at different cellular stages of spermatogenesis of the hamster after in vitro or in vivo exposure to ionizing radiation.Exp Cell Res 193: 303-9.

van Oppen MJH, Souter M, Howells EJ, Heyward A, Berkelmans R (2011) Novel genetic diversity through somatic mutations: fuel for adaptation of reef corals? Diversity 2011:405–23.

Van Straalen NM (2003)Ecotoxicology becomes stress ecology.Environ Sci Technol 37: 324A-330A.

Van Tienderen PH (1991) Evolution of generalists and specialists in spatially heterogeneous environments. Evolution 45: 1317–31.

Van Tienderen PH (1997) Generalists, specialists, and the evolution of phenoptypic plasticity in sympatric populations of distinct species. Evolution 51: 1372–80.

van Uden P, Kenneth NS, Rocha S (2008)Regulation of hypoxia-inducible factor-1alpha by NF-kappaB.Biochem J 412: 477-84.

Van Valen L (1965) Morphological variation and width of ecological niche. Am Nat 99: 377–90.

Van Valen L (1973) A new evolutionary law. Evol Theory 1: 1–30.

Van Valen LM (1974) Two modes of evolution. Nature 252: 298-300.

Van Valen LM (1985) Why and how do mammals evolve unusually rapidly? Evol Theory 7: 127-32.

van Werven FJ, Amon A (2011) Regulation of entry into gametogenesis. Philos Trans R Soc Lond B Biol Sci 366: 3521-31.

Vaseva AV, Moll UM (2009)The mitochondrial p53 pathway.Biochim Biophys Acta 1787: 414-20.

Vaseva AV, Yallowitz AR, Marchenko ND, Xu S, Moll UM (2011) Blockade of Hsp90 by 17AAG antagonizes MDMX and synergizes with Nutlin to induce p53-mediated apoptosis in solid tumors. Cell Death Dis 2: e156.

Vashisht AA, Tuteja N (2006) Stress responsive DEAD-box helicases: a new pathway to engineer plant stress tolerance. J Photochem Photobiol B 84: 150–60.

Vasileva A, Tiedau D, Firooznia A, Muller-Reichert T, Jessberger R (2009) Tdrd6 is required for spermiogenesis, chromatoid body architecture, and regulation of miRNA expression. Curr Biol 19: 630–9.

Vasquez-Vivar J, Kalyanaraman B, Martasek P, Hogg N, Masters BS, et al. (1998) Superoxide generation by endothelial nitric oxide synthase: the influence of cofactors. Proc Natl Acad Sci USA 95: 9220–5.

Vassilieva LL, Lynch M (1999) The rate of spontaneous mutation for life-history traits in Caenorhabditis elegans. Genetics 151: 119–29.

Vassilieva LL, Hook AM, Lynch M (2000) The fitness effects of spontaneous mutations in Caenorhabditis elegans. Evolution 54: 1234–46.

Vastenhouw NL, Plasterk RH (2004) RNAi protects the Caenorhabditis elegans germline against transposition. Trends Genet 20: 314–9.

Vastenhouw NL, Brunschwig K, Okihara KL, Müller F, Tijsterman M, Plasterk RHA (2006) Long-term gene silencing by RNAi. Nature 442: 882.

Vasudevan S, Tong Y, Steitz JA (2007) Switching from repression to activation: microRNAs can up-regulate translation. Science 318: 1931-4.

Vasudevan S, Steitz JA (2007) AU-rich-element-mediated upregulation of translation by FXR1 and Argonaute 2. Cell 128: 1105-18.

Vasudevan S, Tong Y, Steitz JA (2008) Cell-cycle control of microRNA-mediated translation regulation. Cell cycle 7: 1545-9.

Vatolin S, Abdullaev Z, Pack SD, Flanagan PT, Custer M, et al. (2005) Conditional expression of the CTCF-paralogous transcriptional factor BORIS in normal cells results in demethylation and derepression of MAGE-A1 and reactivation of other cancer-testis genes. Cancer Res 65: 7751–62.

Vaughn MW, Tanurdzic M, Lippman Z, Jiang H, Carrasquillo R, et al. (2007) Epigenetic natural variation in Arabidopsis thaliana. PLoS Biol 5: e174.

Vazquez A, Bond EE, Levine AJ, Bond GL (2008) The genetics of the p53 pathway, apoptosis and cancer therapy. Nat Rev Drug Discov 7: 979-87.

Vecchione A, Croce CM (2010) Apoptomirs: small molecules have gained the license to kill. Endocr Relat Cancer 17: F37–F50.

Veening JW, Smits WK, Kuipers OP (2008a) Bistability, epigenetics, and bet-hedging in bacteria. Annu Rev Microbiol 62: 193–210.

Veening JW, Stewart EJ, Berngruber TW, Taddei F, Kuipers OP, Hamoen LW (2008b) Bethedging and epigenetic inheritance in bacterial cell development. Proc Natl Acad Sci USA 105: 4393–98.

Veevers JJ (2004) Gondwanaland from 650–500 Ma assembly through 320 Ma merger in Pangea to 185–100 Ma breakup: supercontinental tectonics via stratigraphy and radiometric dating. Earth Sci Rev 68: 1–132.

Velando A, Beamonte-Barrientos R, Torres R (2006) Pigment-based skin colour in the blue-footed booby: an honest signal of current condition used by females to adjust reproductive investment.Oecologia 149: 535-42.

Velando A, Torres R, Alonso-Alvarez C (2008)Avoiding bad genes: oxidatively damaged DNA in germ line and mate choice.Bioessays 30: 1212-9.

Velasco-Miguel S, Richardson JA, Gerlach VL, Lai WC, Gao T, et al. (2003) Constitutive and regulated expression of the mouse Dinb (Polkappa) gene encoding DNA polymerase kappa. DNA Repair (Amst) 2: 91-106.

Velentzas AD, Nezis IP, Stravopodis DJ, Papassideri IS, Margaritis LH (2007)Apoptosis and autophagy function cooperatively for the efficacious execution of programmed nurse cell death during Drosophila virilis oogenesis.Autophagy 3: 130-2.

Velkov VV (2002) New insights into the molecular mechanisms of evolution: stress increases genetic diversity. Mol Biol 36: 209–15.

Vellend M (2006) The consequences of genetic diversity in competitive communities. Ecology 87: 304-11.

Vellend M, Geber MA (2005) Connections between species diversity and genetic diversity. Ecol Lett 8: 767–81.

Venable DL (2007) Bet hedging in a guild of desert annuals. Ecology 88: 1086–90.

Venditti C, Pagel M (2010)Speciation as an active force in promoting genetic evolution.Trends Ecol Evol 25: 14-20.

Venditti P, Balestrieri M, Di Meo S, De Leo T (1997) Effect of thyroid state on lipid peroxidation, antioxidant defences, and susceptibility to oxidative stress in rat tissues. J Endocrinol 155: 151–7.

Venditti P, De Rosa R, Di Meo S (2003) Effect of thyroid state on H2O2 production by rat liver mitochondria. Mol Cell Endocrinol 205: 185-92.

Venditti P, Di Meo S (2006) Thyroid hormone-induced oxidative stress. Cell Mol Life Sci 63: 414–34.

Ventela S, Toppari J, Parvinen M (2003) Intercellular organelle traffic through cytoplasmic bridges in early spermatids of the rat: mechanisms of haploid gene product sharing. Mol Biol Cell 14: 2768–80.

Ventura N, Rea SL, Testi R (2006) Long-lived C. elegans mitochondrial mutants as a model for human mitochondrial-associated diseases. Exp Gerontol 41: 974–91.

Ventura N, Rea SL, Schiavi A, Torgovnick A, Testi R, Johnson TE (2009) p53/CEP-1 increases or decreases lifespan, depending on level of mitochondrial bioenergetic stress.Aging Cell 8: 380-93.

Vera H, Tijmes M, Valladares LE (1997) Melatonin and testicular function: characterization of binding sites for 2-[125I]-iodomelatonin in immature rat testes. Steroids 62: 226–9.

Vera Y, Diaz-Romero M, Rodrigues S, Lue Y, Wang C, et al.(2004) Mitochondria dependent pathway is involved in heat-induced male germ cells death: lessons from mutant mice. Biol Reprod 70: 1534–40.

Verbeek S, van Lohuizen M, van der Valk M, Domen J, Kraal G, Berns A (1991) Mice bearing the E mu-myc and E mu-pim-1 transgenes develop pre-B-cell leukemia prenatally. Mol Cell Biol 11: 1176–9.

Verburg R, Grava D (1998) Differences in allocation patterns in clonal and sexual offspring in a woodland pseudo-annual. Oecologia 115: 472-7.

Vercesi AE (2001) The discovery of an uncoupling mitochondrial protein in plants. Biosci Rep 21: 195–200.

Verdin E, Dequiedt F, Kasler HG (2003)Class II histone deacetylases: versatile regulators.Trends Genet 19: 286-93.

Vergara SP, Lizama C, Brouwer-Visser J, Moreno RD (2011) Expression of BCL-2 family genes in germ cells undergoing apoptosis during the first wave of spermatogenesis in the rat. Andrologia 43:242-7.

Verghese S, Bedi S, Kango-Singh M (2012)Hippo signalling controls Dronc activity to regulate organ size in Drosophila.Cell Death Differ 19: 1664-76.

Verhoeven G, Cailleau J, Van Damme J, Billiau A (1988) Interleukin-1 stimulates steroidogenesis in cultured rat Leydig cells. Mol Cell Endocrinol 57: 51-60.

Verhoeven KJF, Jansen JJ, van Dijk PJ, Biere A(2010a) Stress-induced DNA methylation changes and their heritability in asexual dandelions. New Phytol 185: 1108-18.

Verhoeven KJF, van Dijk PJ, Biere A (2010b) Changes in genomic methylation patterns during the formation of triploid asexual dandelion lineages.Mol Ecol 19: 315–24.

Verhoeven KJF, van Gurp TP (2012) Transgenerational effects of stress exposure on offspring phenotypes in apomictic dandelion. PLoS ONE 7: e38605.

Verkman AS (2005)More than just water channels: unexpected cellular roles of aquaporins.J Cell Sci 118: 3225-32.

Verkman AS (2008)Mammalian aquaporins: diverse physiological roles and potential clinical significance.Expert Rev Mol Med 10: e13.

Vermeij GJ (1983) Intimate associations and coevolution in the sea. In: Futuyma DM, Slatkin M, eds. Coevolution. Sunderland, MA: Sinauer. pp 311–327.

Vermeij GJ (1987) Evolution and Escalation: An Ecological History of Life. Princeton, NJ: Princeton University Press.

Vermeij GJ (1990) The origin of skeletons. Palaios 4: 585–89.

Vermeij GJ (1994) The evolutionary interaction among species: selection, escalation, and coevolution. Annu Rev Ecol Syst 25:219–36.

Vermeij GJ (2004) Nature: An Economic History. Princeton, NJ: Princeton University Press.

Vermes I, Haanen C, Steffens-Nakken H, Reutelingsperger C (1995) A novel assay for apoptosis. Flow cytometric detection of phosphatidylserine expression on early apoptotic cells using fluorescein labelled Annexin V. J Immunol Methods 184: 39–51.

Vermeulen L, Sprick MR, Kemper K, Stassi G, Medema JP (2008)Cancer stem cells--old concepts, new insights.Cell Death Differ 15: 947-58.

Vernon JG, Okamura B, Jones CS, Noble LR (1996) Temporal patterns of clonality and parasitism in a population of freshwater bryozoans. Proc R Soc Lond B Biol Sci 263: 1313–8.

Vesa TH, Marteau P, Korpela R (2000) Lactose intolerance. J Am Coll Nutr 19:165S–175S.

Vézina F, Salvante KG (2010) Behavioral and physiological flexibility are used by birds to manage energy and support investment in the early stages of reproduction. Curr Zool 56: 767–92.

Via S, Lande R (1985) Genotype-environment interaction and the evolution of phenotypic plasticity. Evolution 39: 505–22.

Via S, Gomulkiewicz R, De Jong G, Scheiner SM, Schlichting CD, Van Tienderen PH (1995) Adaptive phenotypic plasticity: consensus and controversy. Trends Ecol Evol 10: 212-7.

Viau V (2002)Functional cross-talk between the hypothalamic-pituitary-gonadal and -adrenal axes.J Neuroendocrinol 14: 506-13.

Viau V, Meaney MJ (1991) Individual differences in the hypothalamicpituitary-adrenal response to stress: relationship to testosterone. Soc Neurosci Abstr 17: 542.

Viau V, Meaney MJ (1992) Regulation of ACTH co-secretagogues during the estrous cycle in the rat. Soc Neurosci Abstr 18: 643.

Viau V, Meaney MJ (1996)The inhibitory effect of testosterone on hypothalamic-pituitary-adrenal responses to stress is mediated by the medial preoptic area.J Neurosci 16: 1866-76.

Vidales LE, Cardenas LC, Robleto E, Yasbin RE, Pedraza-Reyes M (2009) Defects in the error prevention oxidized guanine system potentiate stationary-phase mutagenesis in Bacillus subtilis. J Bacteriol 191: 506–13.

Vieira J, Vieira CP, Hartl DL, Lozovskaya ER (1998) Factors contributing to the hybrid dysgenesis syndrome in Drosophila virilis. Genet Res 71:109–17.

Vigliola L, Meekan MG (2002) Size at hatching and planktonic growth determine post-settlement survivorship of a coral reef fish. Oecologia 131: 89-93.

Vila-Aiub MM, Neve P, Powles SB (2009)Fitness costs associated with evolved herbicide resistance alleles in plants.New Phytol 184: 751-67.

Vilborg A, Bersani C, Wilhelm MT, Wiman KG (2011) The p53 target Wig-1: a regulator of mRNA stability and stem cell fate? Cell Death Differ 18: 1434-40.

Vilela DAR, Silva SGB, Peixoto MTD, Godinho HP, França LR (2003) Spermatogenesis in teleost: insights from the Nile tilapia (Oreochromis niloticus) model. Fish Physiol Biochem 28: 187–90.

Vilenchik MM, Knudson AG (2003) Endogenous DNA double-strand breaks: production, fidelity of repair, and induction of cancer. Proc Natl Acad Sci USA 100: 12871–6.

Vina J, ed. (1990) Glutathione: metabolism and physiological functions. Boca Raton,FL: CRC Press.

Vincent A, Crozatier M (2010)Neither too much nor too little: reactive oxygen species levels regulate Drosophila hematopoiesis. J Mol Cell Biol 2: 74–5.

Vincent JP, Kolahgar G, Gagliardi M, Piddini E (2011)Steep differences in wingless signaling trigger Myc-independent competitive cell interactions.Dev Cell 21: 366-74.

Vinces MD, Legendre M, Caldara M, Hagihara M, Verstrepen KJ (2009) Unstable tandem repeats in promoters confer transcriptional evolvability. Science 324: 1213-6.

Vinebrooke RD, Cottingham KL, Norberg J, Scheffer M, Dodson SI, et al. (2004) Impacts of multiple stressors on biodiversity and ecosystem functioning: the role of species co-tolerance. Oikos 104: 451-7.

Vineis P (2003) Cancer as an evolutionary process at the cell level: an epidemiological perspective. Carcinogenesis 24: 1-6.

Vinogradov AE (2004) Compactness of human housekeeping genes: selection for economy or genomic design? Trends Genet 20: 248–53.

Vinogradova O, Darienko T, Pavlicek T, Nevo E (2011) Cyanoprokaryotes and algae of Arubota'im salt cave (Mount Sedom, Dead Sea area, Israel). Nova Hedwigia 93: 107-24.

Violle C, Pu Z, Jiang L (2010) Experimental demonstration of the importance of competition under disturbance. Proc Natl Acad Sci USA 107: 12925-9.

Virág L (2005)Structure and function of poly(ADP-ribose) polymerase-1: role in oxidative stress-related pathologies.Curr Vasc Pharmacol 3: 209-14.

Viré E, Brenner C, Deplus R, Blanchon L, Fraga M, et al. (2006)The Polycomb group protein EZH2 directly controls DNA methylation.Nature 439: 871-4.

Vivarelli S, Wagstaff L, Piddini E (2012)Cell wars: regulation of cell survival and proliferation by cell competition.Essays Biochem 53: 69-82.

Voekel-Meiman K, Keil RL, Roeder GS (1987) Recombination-stimulating sequences in yeast ribosomal DNA correspond to sequences regulating transcription by RNA polymerase I. Cell 48: 1071–9.

Voellenkle C, van Rooij J, Guffanti A, Brini E, Fasanaro P, et al. (2012) Deep-sequencing of endothelial cells exposed to hypoxia reveals the complexity of known and novel microRNAs. RNA 18: 472–84.

Vogel F, Rathenberg R (1975) Spontaneous mutations in man. Adv Hum Genet 5: 223–318.

Vogel F, Motulsky AG (1997) Human genetics: Problems and Approaches. Berlin, Germany: Springer.

Vogel F, Motulsky AG (2010) Human genetics: problems and approaches. Berlin, Germany: Springer.

Vogel G (2011)Do jumping genes spawn diversity? Science 332: 300-1.

Vogt G, Huber M, Thiemann M, van den Boogaart G, Schmitz OJ, Schubart CD (2008) Production of different phenotypes from the same genotype in the same environment by developmental variation. J Exp Biol 211: 510–23.

Volff JN (2006)Turning junk into gold: domestication of transposable elements and the creation of new genes in eukaryotes.Bioessays 28: 913-22.

Volff JN, Schartl M (2001) Variability of genetic sex determination in poeciliid fishes. Genetica 111: 101–10.

Volinia S, Calin GA, Liu CG, Ambs S, Cimmino A, et al. (2006) A microRNA expression signature of human solid tumors defines cancer gene targets. Proc Natl Acad Sci USA 103: 2257-61.

Volkov RA, Panchuk II, Mullinaux PM, Schöffl F (2006) Heat stress-induced H2O2 is required for effective expression of heat shock genes in Arabidopsis. Plant Mol Biol 61: 733–46.

Vollmer JH, Sarup P, Kaersgaard CW, Dahlgaard J, Loeschcke V (2004) Heat and cold induced male sterility in Drosophila buzzatii: Genetic variation populations for the duration of sterility. Heredity 92: 257-62.

Volpe TA, Kidner C, Hall IM, Teng G, Grewal SIS, Martienssen RA (2002) Regulation of heterochromatic silencing and histone H3 lysine-9 methylation by RNAi. Science 297: 1833–7.

von Neumann J, Morgenstern O (1944) Theory of games and economic behavior. 1st edn. Princeton, NJ: Princeton University Press.

von Neumann J, Morgenstern O (1953) Theory of games and economic behavior. 3rd edn. Princeton, NJ: Princeton University Press.

von Boehmer H, Aifantis I, Gounari F, Azogui O, Haughn L, et al. (2003)Thymic selection revisited: how essential is it?Immunol Rev 191: 62-78.

von Schantz T, Göransson G, Andersson G, Fröberg I, Grahn M, et al. (1989) Female choice selects for a viability-based male trait in pheasants. Nature 337: 166-9.

von Schantz T, Grahn M, Göransson G (1994) Intersexual selection and reproductive success in the pheasant Phasianus colchicus. Am Nat 144: 510-27.

von Schantz TV, Bensch S, Grahn M, Hasselquist D, Wittzell H (1999) Good genes, oxidative stress and condition-dependent sexual signals. Proc R Soc Lond B Biol Sci 266:1–12.

von Sonntag C (1987a) The Chemical Basis of Radiation Biology. London, UK: Taylor and Francis.

von Sonntag C (1987b) New aspects in the free-radical chemistry of pyrimidine nucleobases. Free Radic Res Commun 2: 217–24.

von Sternberg R (2002)On the roles of repetitive DNA elements in the context of a unified genomic-epigenetic system.Ann NY Acad Sci 981: 154-88.

Vorburger C (2001)Fixation of deleterious mutations in clonal lineages: evidence from hybridogenetic frogs.Evolution 55: 2319-32.

Vorburger C (2004) Cold tolerance in obligate and cyclical parthenogens of the peach-potato aphid, Myzus persicae. Ecol Entomol 29: 498–505.

Vorburger C (2005a) Temporal dynamics of genotypic diversity reveal strong clonal selection in the aphid Myzus persicae. J Evol Biol 19: 97-107.

Vorburger C (2005b) Positive genetic correlations among major life-history traits related to ecological success in the aphid Myzus persicae. Evolution 59: 1006–15.

Vorburger C, Sunnucks P, Ward SA (2003) Explaining the coexistence of asexuals with their sexual progenitors: no evidence for general-purpose genotypes in obligate parthenogens of the peach-potato aphid, Myzus persicae. Ecol Lett 6: 1091–8.

Voronina E, Seydoux G, Sassone-Corsi P, Nagamori I (2011)RNA granules in germ cells.Cold Spring Harb Perspect Biol 3: a002774.

Vousden KH, Lu X (2002)Live or let die: the cell's response to p53.Nat Rev Cancer 2: 594-604.

Vousden KH, Lane DP (2007) p53 in health and disease. Nat Rev Mol Cell Biol 8: 275-83.

Vousden KH, Ryan KM (2009) p53 and metabolism. Nat Rev Cancer 9: 691-700.

Vrba ES (1980) Evolution, species and fossils: how does life evolve? S Afr J Sci 76: 61–84.

Vrijenhoek RC (1978) Coexistence of clones in a heterogeneous environment. Science 199: 549–52.

Vrijenhoek RC (1979) Factors affecting clonal diversity and coexistence. Am Zool 19: 787–97.

Vrijenhoek RC (1984) Ecological differentiation among clones: the frozen niche variation model. In: Wohrmann, K. Loeschcke V, eds. Population Biology and Evolution. Berlin, Germany: Springer-Verlag.pp 217–231.

Vrijenhoek RC (1989) Genetic and ecological constraints on the origins and establishment of unisexual vertebrates. In: Dawley RM, Bogart JP, eds. Evolution and Ecology of Unisexual Vertebrates. Albany, NY: The New York State Museum Bulletin. pp 24-31.

Vrijenhoek RC (1994) Genetic diversity and fitness in small populations. In: Tomiuk J, Jain SK, eds. Conservation genetics. Basel, Switzerland: Birkhauser Verlag.pp 37–53.

Vrijenhoek RC (1998) Animal clones and diversity. BioScience 48: 617–28.

Vrijenhoek RC, Dawley RM, Cole CJ, Bogart JP (1989) A list of the known unisexual vertebrates. In: Dawley RM, Bogart JP, eds. Evolution and Ecology of Unisexual Vertebrates. Albany, NY: The New York State Museum Bulletin. pp 19-23.

Vrijenhoek RC, Pfeiler E (1997) Differential survival of sexual and asexual Poeciliopsis during environmental stress. Evolution 51: 1593–1600.

Vrijenhoek RC, Parker ED Jr (2009) Geographical parthenogenesis: general purpose genotypes and frozen niche variation. In: Schön I, Martens K, Van Dijk P, eds. Lost sex. Berlin, Germany: Springer Publications. pp 99–131.

Vulic M, Kolter R (2001) Evolutionary cheating in Escherichia coli stationary phase cultures. Genetics 158: 519-26.

Vurusaner B, Poli G, Basaga H (2012) Tumor suppressor genes and ROS: complex networks of interactions. Free Radic Biol Med 52: 7-18.

Wachsman JT (1997) DNA methylation and the association between genetic and epigenetic changes: relation to carcinogenesis. Mutat Res 375:1-8.

Wada Y, Miyamoto K, Kusano T, Sano H (2004) Association between up-regulation of stress-responsive genes and hypomethylation of genomic DNA in tobacco plants. Mol Genet Genomics 271: 658–66.

Waddington CH (1942) Canalization of development and the inheritance of acquired characters. Nature 150: 563–5.

Waddington CH (1953a) Genetic assimilation of an acquired character. Evolution 7: 118–26.

Waddington CH (1953b)

Waddington CH (1956) Genetic assimilation of the bithorax phenotype. Evolution 10:1–13.

Waddington CH (1957) The strategy of the genes. London, UK: George Allen & Unwin.

Waddington CH (1961a) The nature of life. London, UK: Allen & Unwin.

Waddington CH (1961b) Genetic assimilation. Adv Genet 10: 257–90.

Wade MJ (1980) Group selection, population growth rate, and competitive ability in the flour beetles, Tribolium spp. Ecology 61: 1056-64.

Waggoner BM, PoinarG O Jr. (1993) Fossil habrotrochid rotifers in Dominican amber. Experientia 49: 354–7.

Wagener A, Blottner S, Göritz F, Streich WJ, Fickel J (2010) Circannual changes in the expression of vascular endothelial growth factor in the testis of roe deer (Capreolus capreolus). Anim Reprod Sci 117: 275–8.

Wagner A (2005a) Robustness and evolvability in living systems. Princeton, NJ: Princeton University Press.

Wagner A (2005b) Distributed robustness versus redundancy as causes of mutational robustness. Bioessays 27: 176–188.

Wagner A (2005c) Robustness, evolvability, and neutrality.FEBS Lett 579: 1772–8.

Wagner A (2008a) Robustness and evolvability: a paradox resolved. Proc Biol Sci 275:91-100.

Wagner A (2008b) Neutralism and selectionism: a network-based reconciliation.Nat Rev Genet 9: 965-74.

Wagner A (2012)The role of robustness in phenotypic adaptation and innovation.Proc Biol Sci 279: 1249-58.

Wagner GP (2003) Evolutionary genetics: The nature of hidden genetic variation unveiled. Curr Biol 13: R958–R960.

Wagner GP, Gabriel W (1990) Quantitative variation in finite pathenogenetic populations: what stops Muller’s ratchet in the absence of recombination? Evolution 44: 715–31.

Wagner GP, Altenberg L (1996) Perspective: complex adaptations and the evolution of evolvability. Evolution 50: 967–76.

Wagner GP, Booth G, Bagheri-Chaichian H (1997) A population genetic theory of canalization. Evolution 51: 329–47.

Wagner GP, Chiu CH, Hansen TF (1999) Is Hsp90 a regulator of evolvability?J Exp Zool 285: 116-8.

Wagner J, Etienne H, Janel-Bintz R, Fuchs RP (2002)Genetics of mutagenesis in E. coli: various combinations of translesion polymerases (Pol II, IV and V) deal with lesion/sequence context diversity. DNA Repair (Amst) 1: 159-67.

Wagner MS, Wajner SM, Maia AL (2008)The role of thyroid hormone in testicular development and function.J Endocrinol 199: 351-65.

Wagner MS, Wajner SM, Maia AL (2009)Is there a role for thyroid hormone on spermatogenesis?Microsc Res Tech 72: 796-808.

Wahls WP, Moore PD (1990a) Relative frequencies of homologous recombination between plasmids introduced into DNA repair-deficient and other mammalian somatic cell lines. Somat Cell Mol Genet 16: 321-9.

Wahls WP, Moore PD (1990b) Homologous recombination enhancement conferred by the Z-DNA motif d(TG)30 is abrogated by simian virus 40 T antigen binding to adjacent DNA sequences. Mol Cell Biol 10: 794-800.

Wai T, Teoli D, Shoubridge EA (2008)The mitochondrial DNA genetic bottleneck results from replication of a subpopulation of genomes.Nat Genet 40:1484-8.

Waites GMH (1970) Temperature regulation and the testis. In: Johnson AD, Gomes WR, VanDemark NL, eds. The Testis, vol. I. New York: Academic Press.pp 241–279.

Wakabayashi T, Spodnik JH (2000)Structural changes of mitochondria during free radical-induced apoptosis.Folia Morphol (Warsz) 59: 61-75.

Wakefield SL, Lane M, Schulz SJ, Hebart ML, Thompson JG, Mitchell M (2008)Maternal supply of omega-3 polyunsaturated fatty acids alter mechanisms involved in oocyte and early embryo development in the mouse.Am J Physiol Endocrinol Metab 294: E425-34.

Wakimoto BT (1998)Beyond the nucleosome: epigenetic aspects of position-effect variegation in Drosophila.Cell 93: 321-4.

Walker CW, Boom JD, Marsh AG (1992)First non-vertebrate member of the myc gene family is seasonally expressed in an invertebrate testis.Oncogene 7: 2007-12.

Walker GC (1996) The SOS response of Escherichia coli.In:Neidhardt FC, Curtiss III R, Ingraham JL, Lin ECC, Low KB, Magasanik B, Reznikoff WS, Riley M, Schaechter M, Umbarger HE, eds. Escherichia coli and Salmonella: cellular and molecular biology, 2nd edn. Washington, DC: American Society for Microbiology.pp 1400–1416.

Walker RS, Gurven M, Burger O, Hamilton MJ (2008) The trade-off between number and size of offspring in humans and other primates. Proc Biol Sci 275: 827–33.

Walker TJ (1984) Do populations self-regulate? In: Huffaker CB, Rabb RL, eds. Ecological Entomology. New York, NY: John Wiley & Sons. pp 531-558.

Walker WH (2009) Molecular mechanisms of testosterone action in spermatogenesis. Steroids 74: 602–7.

Walker WH, Cheng J (2005)FSH and testosterone signaling in Sertoli cells.Reproduction 130: 15-28.

Wall DP, Hirsh AE, Fraser HB, Kumm J, Giaever G, et al. (2005) Functional genomic analysis of the rates of protein evolution. Proc Natl Acad Sci USA 102: 5483–8.

Wallace AR (1858) On the tendency of species to form varieties; and on the perpetuation of varieties and species by natural means of selection. III. On the tendency of varieties to depart indefinitely from the original type. J Proc Linn Soc Zool 3: 53–62.

Wallace B (1958) The average effect of radiation-induced mutations on viability in Drosophila melanogaster. Evolution 12: 532-52.

Wallace DC (2007) Why do we still have a maternally inherited mitochondrial DNA? Insights from evolutionary medicine.Annu Rev Biochem 76:781-821.

Wallace DC, Fan W (2010)Energetics, epigenetics, mitochondrial genetics. Mitochondrion 10:12–31.

Wallace RL, Snell TW (2001) Phylum rotifera. In: Thorp JH, Covich AP, eds. Ecology and classification of North American freshwater invertebrates. 2nd edn. San Diego, CA: Academic Press. pp 195–254.

Wallace RL, Snell TW, Ricci C, Nogrady T (2006) Rotifera 1: Biology, Ecology and Systematics. Leiden, The Netherlands: Backhuys Publishers.

Walser JC, Ponger L, Furano AV (2008) CpG dinucleotides and the mutation rate of non-CpG DNA. Genome Res 18: 1403–14.

Walser JC, Furano AV (2010) The mutational spectrum of non-CpG DNA varies with CpG content. Genome Res 20: 875–82.

Walsh B, Blows MW (2009) Abundant genetic variation + strong selection = multivariate genetic constraints: a geometric view of adaptation. Annu Rev Ecol Evol Syst 40: 41–59.

Walsh CP, Chaillet JR, Bestor TH (1998) Transcription of IAP endogenous retroviruses is constrained by cytosine methylation. Nat Genet 20: 116–7.

Walsh TK, Brisson JA, Robertson HM, Gordon K, Jaubert-Possamai S, et al. (2010) A functional DNA methylation system in the pea aphid Acyrthosiphon pisum. Insect Mol Biol 19: 215–28.

Walter CA, Lu J, Bhakta M, Zhou ZQ, Thompson LH, McCarrey JR (1994) Testis and somatic Xrcc-1 DNA repair gene expression. Somat Cell Mol Genet 20: 451–61.

Walter CA, Trolian DA, McFarland MB, Street KA, McCarrey JR (1996) Xrcc-1 expression during male meiosis. Biol Reprod 55: 630–5.

Walter CA, Intano CW, McCarrey JR, McMahan CA, Walter RB (1998) Mutation frequency declines during spermatogenesis in young mice but increases in old mice. Proc Natl Acad Sci USA 95: 10015-9.

Walter S, Buchner J (2002) Molecular chaperones - cellular machines for protein folding. Angew Chem Int Ed 41:1098-113.

Wan D, Chang P, Yin J (2013) Causes of extra-pair paternity and its inter-specific variation in socially monogamous birds. Acta Ecol Sinica 33: 158–66.

Wandji SA, Srsen V, Nathanielsz PW, Eppig JJ, Fortune JE (1997) Initiation of growth of baboon primordial follicles in vitro. Hum Reprod 12: 1993–2001.

Wang C, Lehmann R (1991)Nanos is the localized posterior determinant in Drosophila. Cell 66: 637-47.

Wang C, Cui YG, Wang XH, Jia Y, Sinha Hikim A, et al. (2007)Transient scrotal hyperthermia and levonorgestrel enhance testosterone-induced spermatogenesis suppression in men through increased germ cell apoptosis.J Clin Endocrinol Metab 92: 3292-304.

Wang D, Nagpal ML, Calkins JH, Chang W, Sigel MM, Lin T (1991) Interleukin-1β induces interleukin-1α messenger ribonucleic acid expression in primary cultures of Leydig cells. Endocrinology 129: 2862–6.

Wang D, Kreutzer DA, Essigmann JM (1998) Mutagenicity and repair of oxidative DNA damage: insights from studies using defined lesions. Mutat Res 400: 99-115.

Wang D, Luo M, Kelley MR (2004) Human apurinic endonuclease 1 (APE1) expression and prognostic significance in osteosarcoma: enhanced sensitivity of osteosarcoma to DNA damaging agents using silencing RNA APE1 expression inhibition. Mol Cancer Ther 3: 679–686.

Wang G, Humayun MZ (1996) Induction of the Escherichia coli UVM response by oxidative stress. Mol Gen Genet 251: 573-9.

Wang G, Zhang J, Moskophidis D, Mivechi NF (2003) Targeted disruption of the heat shock transcription factor (hsf)-2 gene results in increased embryonic lethality, neuronal defects, and reduced spermatogenesis. Genesis 36: 48–61.

Wang G, Ying Z, Jin X, Tu N, Zhang Y, et al. (2004) Essential requirement for both hsf1 and hsf2 transcriptional activity in spermatogenesis and male fertility. Genesis 38: 66–80.

Wang G, Carbajal S, Vijg J, DiGiovanni J, Vasquez KM (2008) DNA structure-induced genomic instability in vivo. J Natl Cancer Inst 100: 1815–7.

Wang GL, Semenza GL (1993) General involvement of hypoxia-inducible factor 1 in transcriptional response to hypoxia. Proc Natl Acad Sci USA 90: 4304–8.

Wang GZ, Liu J, Wang W, Zhang HY, Lercher MJ (2011) A gene’s ability to buffer variation is predicted by its fitness contribution and genetic interactions. PLoS ONE 6: e17650.

Wang H, Huang ZQ, Xia L, Feng Q, Erdjument-Bromage H, et al. (2001) Methylation of histone H4 at arginine 3 facilitating transcriptional activation by nuclear hormone receptor. Science 293: 853–7.

Wang J, Emadali A, Le Bescont A, Callanan M, Rousseaux S, Khochbin S (2011) Induced malignant genome reprogramming in somatic cells by testis-specific factors. Biochim Biophys Acta 1809: 221–5.

Wang J, Mitreva M, Berriman M, Thorne A, Magrini V, et al. (2012) Silencing of germline-expressed genes by DNA elimination in somatic cells. Dev Cell 23: 1072–80.

Wang J, Tian L, Madlung A, Lee HS, Chen M, et al. (2004) Stochastic and epigenetic changes of gene expression in Arabidopsis polyploids. Genetics 167: 1961–73.

Wang KY, Shen CKJ (2004) DNA methyltransferase Dnmt1 and mismatch repair. Oncogene 23: 7898-902.

Wang Q, Dooner HK (2006) Eukaryotic transposable elements and genome evolution special feature: remarkable variation in maize genome structure inferred from haplotype diversity at the bz locus. Proc Natl Acad Sci USA 103: 17644–9.

Wang Q, Ponomareva ON, Lasarev M, Turker MS (2006) High frequency induction of mitotic recombination by ionizing radiation in Mlh1 null mouse cells. Mutat Res 594: 189–98.

Wang RY, Kuo KC, Gehrke CW, Huang LH, Ehrlich M (1982) Heat- and alkali-induced deamination of 5-methylcytosine and cytosine residues in DNA. Biochim Biophys Acta 697: 371–7.

Wang SC, Frey PA (2007)S-adenosylmethionine as an oxidant: the radical SAM superfamily.Trends Biochem Sci 32: 101-10.

Wang SH, Elgin SC (2011) Drosophila Piwi functions downstream of piRNA production mediating a chromatin-based transposon silencing mechanism in female germ line. Proc Natl Acad Sci USA 108: 21164-9.

Wang T, Marquardt C, Foker J (1976) Aerobic glycolysis during lymphocyte proliferation. Nature 261: 702–5.

Wang T, Zhang X, Li JJ (2002) The role of NF-kappaB in the regulation of cell stress responses. Int Immunopharmacol 2: 1509–20.

Wang W, Fang H, Groom L, Cheng A, Zhang W, et al. (2008) Superoxide flashes in single mitochondria. Cell 134: 279–90.

Wang WK, Lin SR, Lee CM, King CC, Chang SC (2002) Dengue type 3 virus in plasma is a population of closely related genomes: Quasispecies. J Virol 76:4662–5.

Wang X, Manganaro F, Schipper HM (1995) A cellular stress model for the sequestration of redox-active glial iron in the aging and degenerating nervous system. J Neurochem 64: 1868-77.

Wang X, Ohnishi T (1997)p53-dependent signal transduction induced by stress.J Radiat Res 38: 179-94.

Wang X, Chen C, Wang L, Chen D, Guang W, French J (2003)Conception, early pregnancy loss, and time to clinical pregnancy: a population-based prospective study. Fertil Steril 79: 577–84.

Wang X, Shen CL, Dyson MT, Eimerl S, Orly J, et al. (2005)Cyclooxygenase-2 regulation of the age-related decline in testosterone biosynthesis.Endocrinology 146: 4202-8.

Wang Y, Salmon AB, Harshman LG (2001) A cost of reproduction: oxidative stress susceptibility is associated with increased egg production in Drosophila melanogaster. Exp Gerontol 36: 1349–59.

Wang Y, Newton DC, Miller TL, Teichert AM, Phillips MJ, et al. (2002) An alternative promoter of the human neuronal nitric oxide synthase gene is expressed specifically in Leydig cells. Am J Pathol 160: 369–80.

Wang Y, Pakunlu RI, Tsao W, Pozharov V, Minko T (2004) Bimodal effect of hypoxia in cancer: role of hypoxia inducible factor in apoptosis. Mol Pharmaceut 1: 156–65.

Wang Y, Wineberg M (2006) Estimation of evolvability genetic algorithm and dynamic environments.Genet Program Evolvable Mach 7: 355–82.

Wang YM, Lin XY, Dong B, Wang YD, Liu B (2004) DNA methylation polymorphism in a set of elite rice cultivars and its possible contribution to intercultivar differential gene expression. Cell Mol Biol Lett 9: 543–56.

Wang YV, Leblanc M, Fox N, Mao JH, Tinkum KL, et al.(2011) Fine-tuning p53 activity through C-terminal modification significantly contributes to HSC homeostasis and mouse radiosensitivity.Genes Dev 25: 1426-38.

Waples RS (1998) Separating the wheat from the chaff: patterns of genetic differentiation in high gene flow species. J Hered 89: 438–450.

Warburg O, Posener K, Negelein E (1924) Über den Stoffwechsel der Tumoren. Biochem Z 152: 319-344.

Warburton D, Stein Z, Kline J (1983) In utero selection against fetuses with trisomy. Am J Hum Genet 35: 1059–64.

Ward S, Carrel JS (1979)Fertilization and sperm competition in the nematode Caenorhabditis elegans.Dev Biol 73: 304-21.

Ward WS (2011) Regulating DNA supercoiling: sperm points the way. Biol Reprod 84: 841–3.

Warner JR (1999) The economics of ribosome biosynthesis in yeast. Trends Biochem Sci 24: 437–40.

Warren DW, Pasupuleti V, Lu Y, Platler BW, Horton R (1990). Tumor necrosis factor and interleukin-1 stimulate testosterone secretion in adult male rat Leydig cells in vitro. J Androl 11: 353-60.

Wasiak S, Legendre-Guillemin V, Puertollano R, Blondeau F, Girard M, et al. (2002) Enthoprotin: a novel clathrin-associated protein identified through subcellular proteomics. J Cell Biol 158: 855–62.

Watanabe H, Hoang VT, Mättner R, Holstein TW (2009) Immortality and the base of multicellular life: lessons from cnidarian stem cells. Semin Cell Dev Biol 20: 1114–25.

Watanabe M (2006) Anhydrobiosis in invertebrates. Appl Entomol Zool 41: 15–31.

Watanabe M, Nakahara Y, Sakashita T, Kikawada T, Fujita A, et al. (2007) Physiological changes leading to anhydrobiosis improve radiation tolerance in Polypedilum vanderplanki larvae. J Insect Physiol 53: 573–9.

Watanabe T, Chuma S, Yamamoto Y, Kuramochi-Miyagawa S, Totoki Y, et al. (2011) MITOPLD is a mitochondrial protein essential for nuage formation and piRNA biogenesis in the mouse germline. Dev Cell 20: 364-75.

Waterland RA, Jirtle RL (2003)Transposable elements: targets for early nutritional effects on epigenetic gene regulation.Mol Cell Biol 23: 5293-300.

Waters LS, Minesinger BK, Wiltrout ME, D'Souza S, Woodruff RV, Walker GC (2009)Eukaryotic translesion polymerases and their roles and regulation in DNA damage tolerance.Microbiol Mol Biol Rev 73: 134-54.

Wathes DC, Abayasekara DR, Aitken RJ (2007) Polyunsaturated fatty acids in male and female reproduction. Biol Reprod 77: 190–201.

Watnick TJ, Gandolph MA, Weber H, Neumann HP, Germino GG (1998) Gene conversion is a likely cause of mutation in PKD1. Hum Mol Genet 7: 1239–43.

Watson C, Margan S, Johnston P (1998) Sex-chromosome elimination in the bandicoot Isoodon macrourus using Y-linked markers. Cytogenet Cell Genet 81: 54–9.

Watson JA, Watson CJ, McCann A, Baugh J (2010) Epigenetics the epicenter of the hypoxic response. Epigenetics 5: 293–6.

Watson JD (1965) Molecular Biology of the Gene. New York: W. A. Benjamin.

Watson MEJ, Burns JL,Smith AL (2004) Hypermutable Haemophilus influenzae with mutations in mutS are found in cystic fibrosis sputum. Microbiology 150: 2947–58.

Watts AG, Swanson LW (1989)Diurnal variations in the content of preprocorticotropin-releasing hormone messenger ribonucleic acids in the hypothalamic paraventricular nucleus of rats of both sexes as measured by in situ hybridization.Endocrinology 125: 1734-8.

Waxman D, Peck JR (1999) Sex and adaptation in a changing environment. Genetics 153: 1041-53.

Wayne R (1985) Chemistry of Atmospheres. Oxford, UK: Clarendon Press.

Waypa GB, Schumacker PT (2002) O2 sensing in hypoxic pulmonary vasoconstriction: the mitochondrial door re-opens. Respir Physiol Neurobiol 132: 81–91.

Weatherford T, Chavez D, Brasky KM, Lemon SM, Martin A, Lanford RE (2009)Lack of adaptation of chimeric GB virus B/hepatitis C virus in the marmoset model: possible effects of bottleneck.J Virol 83: 8062-75.

Weaver IC, Cervoni N, Champagne FA, D'Alessio AC, Sharma S, et al. (2004) Epigenetic programming by maternal behaviour. Nat Neurosci 7: 847–54.

Weber BH, Depew DJ (2003) Evolution and Learning: The Baldwin Effect Reconsidered. Cambridge, MA: MIT Press.

Weber CM, Eke BC, Maines MD (1994)Corticosterone regulates heme oxygenase-2 and NO synthase transcription and protein expression in rat brain.J Neurochem 63: 953-62.

Weber KE (1990) Increased selection response in larger populations. I. Selection for wing–tip height in Drosophila melanogaster at three population sizes. Genetics 125: 579–84.

Weber KE (1996) Large genetic change at small fitness cost in large populations of Drosophila melanogaster selected for wind tunnel flight: rethinking fitness surfaces. Genetics 144: 205–13.

Weber KE, Diggins LT (1990) Increased selection response in larger populations. II. Selection for ethanol vapor resistance in Drosophila melanogaster at two population sizes. Genetics 125: 585–97.

Weber K, Eisman R, Morey L, Patty A, Sparks J, et al. (1999) An analysis of polygenes affecting wing shape on chromosome 3 in Drosophila melanogaster. Genetics 153: 773-86.

Weber M (1996) Evolutionary plasticity in prokaryotes: a Panglossian view. Biol Philos 11: 67–88.

Weber M, Schübeler D (2007)Genomic patterns of DNA methylation: targets and function of an epigenetic mark.Curr Opin Cell Biol 19: 273-80.

Webster JP, Woolhouse MEJ (1999) Cost of resistance: relationship between reduced fertility and increased resistance in a Schistosoma host-parasite system. Proc R Soc Lond B Biol Sci 266: 391–6.

Webster KA, Prentice H, Bishopric NH (2001)Oxidation of zinc finger transcription factors: physiological consequences.Antioxid Redox Signal 3: 535-48.

Webster MS (2004) Density dependence via intercohort competition in a coral-reef fish. Ecology 85: 986–94.

Wedekind C (1994) Mate choice and maternal selection for specific parasite resistances before, during and after fertilization. PhilosTrans R Soc LondB 346: 303–11.

Weeda G, Ma L, van Ham RC, Bootsma D, van der Eb AJ, Hoeijmakers JH (1991) Characterization of the mouse homolog of the XPBC/ERCC-3 gene implicated in xeroderma pigmentosum and Cockayne’s syndrome. Carcinogenesis 12: 2361-8.

Weeks AR, Hoffmann AA (1998) Intense selection of mite clones in a heterogeneous environment. Evolution 52: 1325–33.

Weeks AR, Breeuwer JAJ (2001) Wolbachia-induced parthenogenesis in a genus of phytophagous mites. Proc R Soc Lond B Biol Sci 268: 2245–51.

Weeks AR, Velten R, Stouthamer R (2003) Incidence of a new sex-ratio-distorting endosymbiotic bacterium among arthropods. Proc R Soc Lond B Biol Sci 270: 1857–65.

Weeks AR, Hoffmann AA (2008) Frequency-dependent selection maintains clonal diversity in an asexual organism. Proc Natl Acad Sci 105: 17872-7.

Weeks SC (1995) Comparisons of life-history traits between clonal and sexual fish (Poeciliopsis: Poeciliidae) raised in monoculture and mixed treatments. Evol Ecol 9: 258–74.

Weeks SC (1996) A reevaluation of the Red Queen model for the maintenance of sex in a clonal-sexual fish complex (Poeciliidae: Poeciliopsis). Can J Zool 53: 1157–64.

Wegner KM, Reusch TBH, Kalbe M (2004) Multiple parasites are driving major histocompatibility complex polymorphism in the wild. J Evol Biol 16: 224–32.

Wei W, Ba Z, Gao M, Wu Y, Ma Y, et al. (2012) A role for small RNAs in DNA double-strand break repair. Cell 149: 101–12.

Weider LJ (1993a) A test of the ‘general-purpose’ genotype hypothesis: differential tolerance to thermal and salinity stress among Daphnia clones. Evolution 47: 965–9.

Weider LJ (1993b) Niche breadth and life-history variation in a hybrid Daphnia complex. Ecology 74: 935–43.

Weinbauer GF, Aslam H, Krishnamurthy H, Brinkworth MH, Einspanier A, Hodges JK (2001)Quantitative analysis of spermatogenesis and apoptosis in the common marmoset (Callithrix jacchus) reveals high rates of spermatogonial turnover and high spermatogenic efficiency.Biol Reprod 64: 120-6.

Weinberg F, Chandel NS (2009) Mitochondrial metabolism and cancer. Ann NY Acad Sci 1177: 66–73.

Weinberg MS, Villeneuve LM, Ehsani A, Amarzguioui M, Aagaard L, et al. (2006) The antisense strand of small interfering RNAs directs histone methylation and transcriptional gene silencing in human cells. RNA 12: 256–62.

Weinert T, Lydall D (1993) Cell cycle checkpoints, genetic instability and cancer. Semin Cancer Biol 4: 129–40.

Weinhouse C, Anderson OS, Jones TR, Kim J, Liberman SA, et al.An expression microarray approach for the identification of metastable epialleles in the mouse genome.Epigenetics 6: 1105-13.

Weinrauch Y, Penchev R, Dubnau E, Smith I, Dubnau D (1990) A Bacillus subtilis regulatory gene product for genetic competence and sporulation resembles sensor protein members of the bacterial two-component signal-transduction systems. Genes Dev 4: 860-72.

Weinreich DM, Rand DM (2000) Contrasting patterns of non-neutral evolution in proteins encoded in nuclear and mitochondrial genomes. Genetics 156: 385-99.

Weinreich DM, Chao L (2005) Rapid evolutionary escape by large populations from local fitness peaks is likely in nature. Evolution 59: 1175–82.

Weinreich DM, Watson RA, Chao L (2005) Perspective: Sign epistasis and genetic constraint on evolutionary trajectories. Evolution 59: 1165–74.

Weinreich DM, Delaney NF, DePristo MA, Hartl DL (2006) Darwinian evolution can follow only very few mutational paths to fitter proteins. Science 312: 111–4.

Weinzierl RP, Schmidt P, Michiels NK (1999) High fecundity and low fertility in parthenogenetic planarians. Inv Biol 118:87–94.

Weis AE, Gorman WL (1990) Measuring selection on reaction norms: an exploration of the Eurosta-Solidago system. Evolution 44: 820-31.

Weissbach R, Scadden AD (2012) Tudor-SN and ADAR1 are components of cytoplasmic stress granules. RNA 18: 462-71.

Weismann A (1889) The significance of sexual reproduction in the theory of natural selection. In: Poulton SSEB, Shipley AE, eds. Essays Upon Heredity and Kindred Biological Problems. Oxford, UK: Clarendon Press.pp 251–332.

Weismann A (1892) Das Keimplasma. Eine Theorie der Vererbung. Jena, Germany: Fischer.

Weismann A (1896) On germinal selection. Chicago IL: Open Court Publishing Co.

Weismann A (1904) The Evolution Theory. London, UK:Edward Arnold.

Weiss KM (2006)Commentary: evolution of action in cells and organisms.Int J Epidemiol 35: 1159-60.

Weissman C (2011) Germinal selection: a Weismannian solution to Lamarckian problematics. In: Gissis SB, Jablonka E, eds. Transformations of Lamarckism: from Subtle Fluids to Molecular Biology. Cambridge, MA: MIT Press. pp 57-66.

Weitzman SA, Turk PW, Milkowski DH, Kozlowski K (1994)Free radical adducts induce alterations in DNA cytosine methylation.Proc Natl Acad Sci USA 91:1261-4.

Welch JE, Brown PL, O’Brien DA, Magyar PL, Bunch DO, et al.(2000) Human glyceraldehyde 3-phosphate dehydrogenase-2 gene is expressed specifically in spermatogenic cells. J Androl 21:328–38.

Welch JJ, Waxman D (2003) Modularity and the cost of complexity. Evolution 57: 1723–34.

Welch WJ (1990) The mammalian stress response: cell physiology and biochemistry of stress proteins. In: Morimoto RJ, Tissières A, Georgopoulos C, eds. Stress proteins in biology and medicine. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press. pp 223-78.

Welch WJ, Feramisco JR (1982) Purification of the major mammalian heat shock proteins. J Biol Chem 257: 14949–59.

Welden CW, Slauson WL (1986) The intensity of competition versus its importance: an overlooked distinction and some implications. Q Rev Biol 61: 23-44.

Welke K, Schneider JM (2009) Inbreeding avoidance through cryptic female choice in the cannibalistic orb-web spider Argiope lobata. Behav Ecol 20: 1056–62.

Wellejus A, Poulsen HE, Loft S (2000) Iron-induced oxidative DNA damage in rat sperm cells in vivo and in vitro.Free Radic Res 32: 75-83.

Weller SG, Ornduff R (1977) Cryptic self-incompatibility in Amsinckia grandiflora. Evolution 31:47–51.

Wellmann S, Bettkober M, Zelmer A, Seeger K, Faigle M, Eltzschig HK, Bührer C (2008) Hypoxia upregulates the histone demethylase JMJD1A via HIF-1. Biochem Biophys Res Commun 372: 892-7.

Wells CE, Eissenstat DM (2001) Marked differences in survivorship among apple roots of different diameters. Ecology 82: 882–92.

Wells RD, Dere R, Hebert ML, Napierala M, Son LS (2005) Advances in mechanisms of genetic instability related to hereditary neurological diseases. Nucleic Acids Res 33: 3785-98.

Wells RD, Ashizawa T (2006) Genetic Instabilities and Neurological Diseases, 2nd edn. San Diego, CA: Elsevier-Academic Press.

Welsh TH Jr, Johnson BH (1981) Stress-induced alterations in secretion of corticosteroids, progesterone, luteinizing hormone and testosterone in bulls. Endocrinology 109: 185-90.

Welsh TH, Bambino TH, Hsueh AJW (1982) Mechanism of glucocorticoid-induced suppression of testicular androgen biosynthesis in vitro. Biol Reprod 27: 1138–46.

Wendel JF (2000) Genome evolution in polyploids. Plant Mol Biol 42: 225–49.

Wenger RH (2000) Mammalian oxygen sensing, signaling and gene regulation. J Exp Biol 203: 1253–63.

Wenger RH, Katschinski DM (2005) The hypoxic testis and post-meiotic expression of PAS domain proteins.Semin Cell Dev Biol 16: 547-53.

Wennstrom KL, Crews D (1995) Making males from females: the effects of aromatase inhibitors on a parthenogenetic species of whiptail lizard. Gen Comp Endocrinol 99: 316-22.

Wenseleers T, Ratnieks FLW (2004) Tragedy of the commons in Melipona bees. Proc R Soc Lond B 271: S310–S312.

Wenzel P, Mollnau H, Oelze M, Schulz E, Wickramanayake JM, et al. (2008) First evidence for a crosstalk between mitochondrial and NADPH oxidase-derived reactive oxygen species in nitroglycerin-triggered vascular dysfunction. Antioxid Redox Signal 10: 1435–47.

Werdelin L, Nilsonne Å (1999) The evolution of the scrotum and testicular descent in mammals: a phylogenetic view. J Theor Biol 196: 61–72.

Werfel J, Bar-Yam Y (2004)The evolution of reproductive restraint through social communication.Proc Natl Acad Sci USA 101: 11019-24.

Werren JH (1997) Biology of Wolbachia. Ann Rev Entomol 42: 587–609.

Werren JH (2011)Selfish genetic elements, genetic conflict, and evolutionary innovation.Proc Natl Acad Sci USA 108 Suppl 2: 10863-70.

Wessler SR (1996) Plant retrotransposons: turned on by stress. Curr Biol 6: 959–61.

Wessler SR (2006) Transposable elements and the evolution of eukaryotic genomes. Proc Natl Acad Sci USA 103: 17600–1.

West BJ, Deering W (1995) The Lure of Modern Science: Fractal Thinking, Studies of Nonlinear Phenomena in Life Sciences, vol. 3. Hackensack, NJ: World Scientific.

West BJ, Bickel DR (1998) Molecular evolution modeled as a fractal statistical process.Physica A 249: 544-52.

West SA, Lively CM, Read AF (1999) A pluralistic approach to sex and recombination. J Evol Biol 12: 1003–12.

West SA, Gemmill AW, Graham A, Viney ME, Read AF (2001) Immune stress and facultative sex in a parasitic nematode. J Evol Biol 14: 333–7.

West-Eberhard MJ (1983) Sexual selection, social competition, and speciation. Q Rev Biol 58: 155–83.

West-Eberhard MJ (2003) Developmental plasticity and evolution. Oxford, UK: Oxford University Press.

West-Eberhard MJ (2005) Phenotypic accommodation: adaptive innovation due to developmental plasticity. J Exp Zool 304B: 610–8.

West-Eberhard MJ (2007) Dancing with DNA and flirting with the ghost of Lamarck.Biol Philos 22: 439–451.

Westemeier RL, Brawn JD, Simpson SA, Esker TL, Jansen RW, et al. (1998) Tracking the long-term decline and recovery of an isolated population. Science 282: 1695-8.

West-Farrell ER, Xu M, Gomberg MA, Chow YH, Woodruff TK, Shea LD (2009)The mouse follicle microenvironment regulates antrum formation and steroid production: alterations in gene expression profiles.Biol Reprod 80: 432-9.

Wetherington JD, Schenck RA, Vrijenhoek RC (1987) A test of the spontaneous heterosis hypothesis in unisexual vertebrates. Evolution 41: 721–31.

Whangbo JS, Hunter CP (2008) Environmental RNA interference. Trends Genet 24: 297–305.

Wheelwright NT, Logan BA (2004) Previous-year reproduction reduces photosynthetic capacity and slows lifetime growth in females of a neotropical tree. Proc Natl Acad Sci USA 101: 8051–5.

Whelan S, Goldman N (2001) A general empirical model of protein evolution derived from multiple protein families using a maximum-likelihood approach. Mol Biol Evol 18: 691–9.

Wheldon TE (1977)Mutation selection and tumour progression.Med Hypotheses 3: 91-3.

Whitacre JM (2011) Genetic and environment-induced pathways to innovation: on the possibility of a universal relationship between robustness and adaptation in complex biological systems. Evol Ecol 25: 965-75.

Whitacre JM, Bender A (2010) Degeneracy: a design principle for achieving robustness and evolvability. J Theor Biol 263: 143–53.

White DJ, Wolff JN, Pierson M, Gemmell NJ (2008) Revealing the hidden complexities of mtDNA inheritance. Mol Ecol 17: 4925–42.

White MJD (1970) Heterozygosity and genetic polymorphism in parthenogenetic animals. In: Hecht MK, Steere WC, eds. Essays in evolution and genetics in honor of Theodosius Dobzhansky. New York, NY: Appleton-Century-Crofts. pp 237-262.

White MJD (1978) Modes of speciation. San Francisco, CA: WH Freeman.

Whitehead A, Crawford DL (2006) Variation within and among species in gene expression: raw material for evolution. Mol Ecol 15: 1197–211.

Whitelaw E, Martin DI (2001) Retrotransposons as epigenetic mediators of phenotypic variation in mammals. Nat Genet 27: 361-5.

Whitelaw NC, Whitelaw E (2006) How lifetimes shape epigenotype within and across generations. Hum Mol Genet 15: R131–R137.

Whitham TG, Slobodchikoff CN (1981) Evolution by individuals, plant-herbivore interactions, and mosaics of genetic variability: the adaptive significance of somatic mutations in plants. Oecologia 49: 287–92.

Whitham TG, Young WP, Martinsen GD, Gehring CA, Schweizer JA, et al. (2003) Community and ecosystem genetics: a consequence of the extended phenotype. Ecology 84: 559–73.

Whitlock MC (1996) The red queen beats the Jack-of-all-trades: the limitations on the evolution of phenotypic plasticity and niche breadth. Am Nat 148: S65–S77.

Whitlock MC (2000) Fixation of new alleles and the extinction of small populations: drift load, beneficial alleles, and sexual selection. Evolution 54: 1855–61.

Whitlock MC, Bourguet D (2000) Factors affecting the genetic load in Drosophila: synergistic epistasis and correlations among fitness components. Evolution 54: 1654–60.

Whitlock MC, Agrawal AF (2009) Purging the genome with sexual selection: reducing mutation load through selection in males. Evolution 63: 569–82.

Whittam TS, Clark AG, Stoneking M, Cann RL, Wilson AC (1986) Allelic variation in human mitochondrial genes based on patterns of restriction site polymorphism. Proc Natl Acad Sci USA 83: 9611-15.

Whittle CA, Johnston MO (2002) Male-driven evolution of mitochondrial and chloroplastidial DNA sequences in plants. Mol Biol Evol 19: 938–49.

Whittle CA, Johnston MO (2003) Male-biased transmission of deleterious mutations to the progeny in Arabidopsis thaliana. Proc Natl Acad Sci USA 100: 4055–9.

Whitton J, Sears CJ, Baack EJ, Otto SP (2008) The dynamic nature of apomixis in the angiosperms. Int J Plant Sci 169: 169–82.

Wibbels T, Crews D (1994) Putative aromatase inhibitor induces male sex determination in a female unisexual lizard and in a turtle with temperature-dependent sex determination. J Endocrinol 141: 295-9.

Wichman HA, Millstein J, Bull JJ (2005) Adaptive molecular evolution for 13,000 phage generations: a possible arms race. Genetics 170: 19–31.

Widén B, Cronberg N, Widén M (1994) Genotypic diversity, molecular markers and spatial distribution of genets in clonal plants, a literature survey. Folia Geobot Phytotaxon29:245–63.

Widlak W, Vydra N, Malusecka E, Dudaladava V, Winiarski B, et al. (2007)Heat shock transcription factor 1 down-regulates spermatocyte-specific 70 kDa heat shock protein expression prior to the induction of apoptosis in mouse testes.Genes Cells 12: 487-99.

Widmann C, Gibson S, Jarpe MB, Johnson GL (1999) Mitogen-activated protein kinase: Conservation of a three-kinase module from yeast to human. Physiol Rev 79: 143-80.

Wiemer EA (2011) Stressed tumor cell, chemosensitized cancer. Nat Med 17: 1552–4.

Wiener N (1948) Cybernetics: control and communication in the animal and in the machine. New York, NY: Wiley.

Wiens JJ (2001) Widespread loss of sexually selected traits: how the peacock lost its spots. Trends Ecol Evol 16: 517–23.

Wierdl M, Greene CN, Datta A, Jinks-Robertson S, Petes TD (1996) Destabilization of simple repetitive DNA sequences by transcription in yeast. Genetics 143: 713-21.

Wiersma P, Selman C, Speakman JR, Verhulst S (2004) Birds sacrifice oxidative protection for reproduction. Proc Biol Sci 271 Suppl 5: S360–3.

WieschausE, Szabad J (1979) The developmental and function of the female germ line in Drosophila melanogaster: A cell lineage study. Dev Biol 68: 29-46.

Wiesener MS, Maxwell PH (2003) HIF and oxygen sensing; as important to life as the air we breathe? Ann Med 35: 183–90.

Wightman B, Ha I, Ruvkun G (1993) Posttranscriptional regulation of the heterochronic gene lin-14 by lin-4 mediates temporal pattern formation in C. elegans. Cell 75: 855-62.

Wilbur HM, Rudolf VH (2006)Life-history evolution in uncertain environments: bet hedging in time.Am Nat 168: 398-411.

Wilcox AJ, Weinberg CR, O'Connor JF, Baird DD, Schlatterer JP, Canfield RE, et al. (1988) Incidence of early loss of pregnancy. N Engl J Med 319: 189–94.

Wilcox DE, Schenk AD, Feldman BM, Xu Y (2001)Oxidation of zinc-binding cysteine residues in transcription factor proteins.Antioxid Redox Signal 3: 549-64.

Wilding M, Carotenuto R, Infante V, Dale B, Marino M, Di Matteo L, Campanella C (2001)Confocal microscopy analysis of the activity of mitochondria contained within the 'mitochondrial cloud' during oogenesis in Xenopus laevis.Zygote 9: 347-52.

Wilhelm BT, Marguerat S, Watt S, Schubert F, Wood V, et al (2008) Dynamic repertoire of a eukaryotic transcriptome surveyed at single-nucleotide resolution. Nature 453: 1239–43.

Wilhelm Filho D, Marcon JL, Fraga CG, Boveris A (2000) Antioxidant defenses in vertebrates: emphasis on fish and mammals. Trends Comp Biochem Physiol 7: 33-45.

Wilhelm Filho D (2007) Reactive oxygen species, antioxidants and fish mitochondria.Frontiers Biosci 12: 1229-37.

Wilk K, Bilinski S, DoughertyMT, Kloc M (2005) Delivery of germinal granules and localized RNAs via the messenger transport organizer pathway to the vegetal cortex of Xenopus oocytes occurs through directional expansion of the mitochondrial cloud. Int J Dev Biol 49: 17–21.

Wilke CM, Adams J (1992) Fitness effects of Ty transposition in Saccharomyces cerevisiae. Genetics 131: 31-42.

Wilke CO (2004) The speed of adaptation in large asexual populations. Genetics 167: 2045–53.

Wilke CO (2005) Quasispecies theory in the context of population genetics. BMC Evol Biol 5: 44.

Wilke CO, Wang JL, Ofria C, Lenski RE, Adami C (2001) Evolution of digital organisms at high mutation rate leads to survival of the flattest. Nature 412: 331–3.

Wilke CO, Adami C (2003) Evolution of mutational robustness. Mutat Res 522: 3-11.

Wilkin DJ, Szabo JK, Cameron R, Henderson S, Bellus GA, et al. (1998) Mutations in fibroblast growth-factor receptor 3 in sporadic cases of achondroplasia occur exclusively on the paternally derived chromosome. Am J Hum Genet 63: 711–6.

Wilkins AS (1997)Canalization: a molecular genetic perspective.BioEssays 19: 257-62.

Wilkins AS (2012) Evolution: A View from the 21st Century.Genome Biol Evol 4: 423–426.

Wilkins AS, Holliday R (2009)The evolution of meiosis from mitosis.Genetics 181: 3-12.

Wilkins JF (2010) Antagonistic coevolution of two imprinted loci with pleiotropic effects. Evolution 64: 142-51.

Wilkinson GS, Presgraves DC, Crymes L (1998) Male eye span in stalk-eyed flies indicates genetic quality by meiotic drive suppression. Nature 391: 276–9.

Wilkinson GS, Taper M (1999) Evolution of genetic variation for condition-dependent traits in stalk-eyed flies. Proc R SocLondB 266: 1685-90.

Wilkinson KD (1997) Regulation of ubiquitin-dependent processes by deubiquitinating enzymes. FASEB J 11: 1245–56.

Willensdorfer M (2009) On the evolution of differentiated multicellularity. Evolution 63: 306–23.

Willi Y, Van Buskirk J, Hoffmann AA (2006) Limits to the adaptive potential of small populations. Annu Rev Ecol Evol Syst 37: 433–58.

Willi Y, Van Buskirk J, Schmid B, Fischer M(2007) Genetic isolation of fragmented populations is exacerbated by drift and selection. J Evol Biol 20:534–42.

Willi Y, Hoffmann AA (2009) Demographic factors and genetic variation influence population persistence under environmental change. J Evol Biol 22: 124–33.

Williams CT (1980) Low temperature mortality of cereal aphids. IOBC/WPRS Bull 3: 63–6.

Williams DD (1958) A histological study of the effects of subnormal temperature on the testis of the fowl. Anat Rec 130: 225-41.

Williams DG, Mack RN, Black RA (1995) Ecophysiology of introduced Pennisetum setaceum on Hawaii: the role of phenotypic plasticity. Ecology 76: 1569–80.

Williams DLH, ed. (1988) Nitrosation. Cambridge, UK: Cambridge University Press.

Williams GC (1957) Pleiotropy, natural selection, and the evolution of senescence. Evolution 11: 398–411.

Williams GC (1966a) Adaptation and Natural Selection. Princeton, NJ: Princeton University Press.

Williams GC (1966b) Natural selection, the costs of reproduction, and a refinement of Lack’s principle. Am Nat 100: 687–90.

Williams GC (1975) Sex and Evolution. Princeton, NJ: Princeton University Press.

Williams HT, Lenton TM (2007)Artificial selection of simulated microbial ecosystems.Proc Natl Acad Sci USA 104: 8918-23.

Williams JG (2006) Transcriptional regulation of Dictyostelium pattern formation. EMBO Rep 7: 694–8.

Williams JG (2010) Dictyostelium finds new roles to model. Genetics 185: 717–26.

Williams K, Christensen J, Helin K (2011)DNA methylation: TET proteins-guardians of CpG islands?EMBO Rep 13: 28-35.

Williams RS, Benjamin IJ (2000) Protective responses in the ischemic myocardium. J Clin Invest 106: 813–8.

Williams TD (1994) Intra-specific variation in egg size and egg composition in birds: effects on offspring fitness. Biol Rev 68: 35–59.

Willig MR, Kaufman DM, Stevens RD (2003) Latitudinal gradients of biodiversity: pattern, process, scale, and synthesis. Annu Rev Ecol Evol Syst 34:273–309.

Willingham AT, Gingeras TR (2006) TUF love for “junk” DNA. Cell 125: 1215–20.

Wills C (1984) The possibility of stress-triggered evolution. In: Mani GS, ed. Evolutionary dynamics of genetic diversity. Berlin, Germany: Springer.pp 299-312.

Wilsch-Brauninger M, Schwarz H, Nüsslein-Volhard C (1997) A sponge-like structure involved in the association and transport of maternal products during Drosophila oogenesis. J Cell Biol 139: 817–29.

Wilson AC, Sunnucks P, Hales DF (1999)Microevolution, low clonal diversity and genetic affinities of parthenogenetic sitobion aphids in New Zealand.Mol Ecol 8:1655-66.

Wilson ACC, Sunnucks P, Hales DF (2003) Heritable genetic variation and potential for adaptive evolution in asexual aphids (Aphidoidea). Biol J Lin Soc 79: 115–35.

Wilson AJ, Pemberton JM, Pilkington JG, Coltman DW, Mifsud DV, et al. (2006) Environmental coupling of selection and heritability limits evolution. PLoS Biol 4: e216.

Wilson CG (2011) Desiccation-tolerance in bdelloid rotifers facilitates spatiotemporal escape from multiple species of parasitic fungi. Biol J Linn Soc 104: 564–74.

Wilson CG, Sherman PW (2010) Anciently asexual bdelloid rotifers escape lethal fungal parasites by drying up and blowing away. Science 327:574-6.

Wilson DS, Turelli M (1986) Stable underdominance and the evolutionary invasion of empty niches. Am Nat 127: 835-50.

Wilson DS, Pollock GB, Dugatkin LA (1992) Can altruism evolve in purely viscous populations? Evol Ecol 6: 331–41.

Wilson DS, Wilson EO (2007) Evolution: survival of the selfless. New Sci 196: 42–6.

Wilson DT, Meekan MG (2002) Growth-related advantages for survival to the point of replenishment in the coral reef fish Stegastes partitus (Pomacentridae). Mar Ecol Progr Ser 231: 247-60.

Wilson EO (1987) Causes of ecological success—the case of the ants: the 6th Tansley lecture. JAnim Ecol 56:1–9.

Wilson EO (2012) The social conquest of earth. New York, NY: W.W. Norton & Company.

Wilson FE, Follett BK (1976) Corticosterone-induced gonadosuppression in photostimulated tree sparrows. Life Sci 17: 1451-6.

Wilson MJ, Bowles J, Koopman P (2006)The matricellular protein SPARC is internalized in Sertoli, Leydig, and germ cells during testis differentiation.Mol Reprod Dev 73: 531-9.

Wilson R, Chopra M, Bradley H, McKillop JH, Smith WE, Thomson JA (1989) Free radicals and Graves’ disease: the effects of therapy. Clin Endocrinol 30: 429–33.

Wilson TM, Rivkees SA, Deutsch WA, Kelley MR (1996) Differential expression of the apurinic/apyrimidinic endonuclease (APE/ref-1) multifunctional DNA base excision repair gene during fetal development and in adult rat brain and testis. Mutat Res 362: 237–48.

Wingfield JC, Sapolsky RM (2003) Reproduction and resistance to stress: when and how. J Neuroendocrinol 15:711–24.

Wink DA, Kasprzak KS, Maragos CM, Elespuru RK, Misra M, et al. (1991) DNA deamination ability and genotoxicity of nitric oxide and its progenitors. Science 254: 1001-3.

Wink DA, Laval J (1994) The Fpg protein, a DNA repair enzyme, is inhibited by the biomediator nitric oxide in vitro and in vivo. Carcinogenesis 15: 2125-9.

Wink DA, Nims RW, Saavedra JE, Utermahlen WE Jr, Ford PC (1994) The Fenton oxidation mechanism: reactivities of biologically relevant substrates with two oxidizing intermediates differ from those predicted for the hydroxyl radical. Proc Natl Acad Sci USA 91: 6604-8.

Winn LM, Kim PM, Nickoloff JA (2003) Oxidative stress-induced homologous recombination as a novel mechanism for phenytoin-initiated toxicity. J Pharmacol Exp Ther 306: 523–7.

Winston WM, Molodowitch C, Hunter CP (2002) Systemic RNAi in C. elegans requires the putative transmembrane protein SID-1. Science 295: 2456–9.

Winston WM, Sutherlin M, Wright AJ, Feinberg EH, Hunter CP (2007)Caenorhabditis elegans SID-2 is required for environmental RNA interference.Proc Natl Acad Sci USA 104: 10565-70.

Winther RG (2005) Evolutionary developmental biology meets levels of selection: modular integration or competition, or both? In: Rasskin-Gutman D, Callebaut W, eds. Modularity: Understanding the Development and Evolution of Natural Complex Systems. Cambridge, MA: MIT Press. pp 61–97.

Wirth R (1994) The sex pheromone system of Enterococcus faecalis: more than just a plasmid-collection mechanism. Eur J Biochem 222: 235-46.

Wise CA, Sraml M, Easteal S (1998) Departure from neutrality at the mitochondrial NADH dehydrogenase subunit 2 gene in humans, but not in chimpanzees. Genetics 148: 409-21.

Wise PM, KrajnakKM, Kashon ML (1996) Menopause: The aging of multiple pacemakers. Science 273: 67–70.

Wise PM, Kashon ML, Krajnak KM, Rosewell KL, Cai A, et al. (1997) Aging of the female reproductive system: a window into brain aging. Recent Prog Horm Res 52: 279–303.

Wishart GJ (1984) Effects of lipid peroxide formation in fowl semen on sperm motility, ATP content and fertilizing ability. J Reprod Fert 71: 113–8.

Wloch DM, Szafraniec K, Borts RH, Korona R (2001a) Direct estimate of the mutation rate and the distribution of fitness effects in the yeast Saccharomyces cerevisiae. Genetics 159: 441–52.

Wloch DM, Borts RH, Korona R (2001b) Epistatic interactions of spontaneous mutations in haploid strains of the yeast Saccharomyces cerevisiae. J Evol Biol 14: 310–6.

Woitkevitsch AA (1940) Dependence of seasonal periodicity in gonadal changes in the thyroid gland in Sturnus vulgaris L.C.R. (Doklady). Acad Sci URSS 27: 741–5.

Wojciechowski MF, Hoelzer MA, Michod RE (1989) DNA repair and the evolution of transformation in Bacillus subtilis. II. Role of inducible repair. Genetics 121: 411–22.

Wojtczak A, Poplonska K, Kwiatkowska M (2008) Phosphorylation of H2AX histone as indirect evidence for double-stranded DNA breaks related to the exchange of nuclear proteins and chromatin remodeling in Chara vulgaris spermiogenesis. Protoplasma 233: 263–7.

Wolf A, Krause-Gruszczynska M, Birkenmeier O, Ostareck-Lederer A, Hüttelmaier S, Hatzfeld M (2010) Plakophilin 1 stimulates translation by promoting eIF4A1 activity. J Cell Biol 188: 463–71.

Wolf AT, Howe RW, Hamrick JL (2000) Genetic diversity and population structure of the serpentine endemic Calystegia collina (Convolvulaceae) in northern California. Am J Bot87: 38-1146.

Wolf DM, Vazirani VV, Arkin AP (2005) Diversity in times of adversity: probabilistic strategies in microbial survival games. J Theor Biol 234: 227–53.

Wolf HG, Wöhrmann K, Tomiuk J (1987) Experimental evidence for the adaptive value of sexual reproduction. Genetica 72: 151-9.

Wolf JB, Brodie III ED, Wade MJ, eds. (2000) Epistasis and the evolutionary process. New York, NY: Oxford University Press.

Wolf MY, Wolf YI, Koonin EV (2008)Comparable contributions of structural-functional constraints and expression level to the rate of protein sequence evolution.Biol Direct 3: 40.

Wolfe KH, Sharp PM, Li W-H (1989) Mutation rates differ among regions of the mammalian genome. Nature 337: 283–85.

Wolfes H, Kogawa K, Millette CF, Cooper GM (1989)Specific expression of nuclear proto-oncogenes before entry into meiotic prophase of spermatogenesis.Science 245: 740-3.

Wolff JN, White DJ, Woodhams M, White HE, Gemmell NJ (2011) The strength and timing of the mitochondrial bottleneck in salmon suggests a conserved mechanism in vertebrates. PLoS ONE 6: e20522.

Wolff JN, Gemmell NJ (2012) Mitochondria, maternal inheritance, and asymmetric fitness: Why males die younger.Bioessays 35: 93–9.

Wolffe AP, Matzke MA (1999)Epigenetics: regulation through repression.Science 286: 481-6.

Wolffe AP, Guschin D (2000) Review: chromatin structural features and targets that regulate transcription. J Struct Biol 129: 102–22.

Wolfner MF (1997) Tokens of love: functions and regulation of Drosophila male accessory gland products. Insect Biochem Mol Biol 27: 179–92.

Wolfrum C, Shi S, Jayaprakash KN, Jayaraman M, Wang G, et al. (2007) Mechanisms and optimization of in vivo delivery of lipophilic siRNAs. Nat Biotechnol 25: 1149-57.

Wolfson A (1954) Sperm storage at lower-than-body temperature outside the body cavity in some passerine birds. Science 120: 68-71.

Wolinska J, Lively CM(2008) The cost of males in Daphnia pulex. Oikos 117: 1637–46.

Woltereck R (1909) Weitere experimentelle Untersuchungen über Artsveränderung, speziell über das Wesen quantitativer Artunterschiede bei Daphniden. Verh Deutsch Zool Ges 19: 110–72.

Wong AH, Gottesman II, Petronis A (2005)Phenotypic differences in genetically identical organisms: the epigenetic perspective.Hum Mol Genet 14 Spec No 1: R11-8.

Wong W (2012) Peroxide promotes survival. Sci Signal 5: ec252.

Wood RD, Mitchell M, Sgouros J, Lindahl T (2001) Human DNA repair genes. Science 291: 1284-9.

Wood TE, Burke JM, Rieseberg LH (2005) Parallel genotypic adaptation: when evolution repeats itself. Genetica 123: 157–70.

Woodford N, Ellington MJ (2007)The emergence of antibiotic resistance by mutation.Clin Microbiol Infect 13: 5-18.

Woodruff RC, Thompson JN (2002) Mutation and premating isolation. Genetica 116: 371–82.

Woodruff RC, Zhang M (2009) Adaptation from leaps in the dark. J Hered 100: 7–10.

Woods RJ, Barrick JE, Cooper TF, Shrestha U, KauthMR, Lenski RE (2011) Second-order selection for evolvability in a large Escherichia coli population. Science331:1433–6.

Woodson JD, Chory J (2008) Coordination of gene expression between organellar and nuclear genomes. Nat Rev Genet 9: 383–95.

Woolfit M,Bromham L (2005) Population size and molecular evolution on islands. Proc Biol Sci 272:2277–82.

Woolhouse ME, Webster JP, Domingo E, Charlesworth B, Levin BR (2002)Biological and biomedical implications of the co-evolution of pathogens and their hosts.Nat Genet 32: 569-77.

Woolveridge I, de Boer-Brower M, Taylor MF, Teerds KJ, Wu FC, Morris ID (1999) Apoptosis in the rat spermatogenic epithelium following androgen withdrawal: changes in apoptosis-related genes. Biol Reprod 60: 461–70.

Working PK, Butterworth BE (1984) An assay to detect chemically induced DNA repair in rat spermatocytes. Environ Mutagen 6: 273–86.

Wossidlo M, Nakamura T, Lepikhov K, Marques CJ, Zakhartchenko V, et al. (2011)5-Hydroxymethylcytosine in the mammalian zygote is linked with epigenetic reprogramming.Nat Commun 2: 241.

Wouters BG, van den Beucken T, Magagnin MG, Koritzinsky M, Fels D, Koumenis C (2005) Control of the hypoxic response through regulation of mRNA translation. Semin Cell Dev Biol 16: 487–501.

Wouters-Tyrou D, Martinage A, Chevaillier P, Sautiere P (1998) Nuclear basic proteins in spermiogenesis. Biochimie 80: 117-28.

Wren JD, Forgacs E, Fondon JW 3rd, Pertsemlidis A, Cheng SY, et al. (2000) Repeat polymorphisms within gene regions: phenotypic and evolutionary implications. Am J Hum Genet 67: 345–56.

Wright A, Reiley WW, Chang M, Jin W, Lee AJ, et al. (2007)Regulation of early wave of germ cell apoptosis and spermatogenesis by deubiquitinating enzyme CYLD.Dev Cell 13: 705-16.

Wright AF, Murphy MP, Turnbull DM (2009) Do organellar genomes function as long-term redox damage sensors? Trends Genet 25: 253-61.

Wright BE (2000) A biochemical mechanism for nonrandom mutations and evolution. J Bacteriol 182: 2993–3001.

Wright BE (2004)Stress-directed adaptive mutations and evolution.Mol Microbiol 52: 643-50.

Wright BE, Longacre A, Reimers JM (1999) Hypermutation in derepressed operons of Escherichia coli K12. Proc Natl Acad Sci USA 96: 5089–94.

Wright EG (2010) Manifestations and mechanisms of non-targeted effects of ionizing radiation. Mutat Res 687: 28-33.

Wright GL, Maroulakou IG, Eldridge J, Liby TL, Sridharan V, et al. (2008)VEGF stimulation of mitochondrial biogenesis: requirement of AKT3 kinase.FASEB J 22: 3264-75.

Wright JB, Brown SJ, Cole MD (2010) Upregulation of c-MYC in cis through a large chromatin loop linked to a cancer risk-associated single-nucleotide polymorphism in colorectal cancer cells. Mol Cell Biol 30: 1411–20.

Wright JW, Lowe CH (1968) Weeds, polyploids, parthenogenesis, and the geographical and ecological distribution of all-female species of Cnemidophorus. Copeia 1968: 128–38.

Wright S (1931) Evolution in Mendelian populations. Genetics 16: 97-159.

Wright SD, Gray RD, Gardner RC (2003) Energy and the rate of evolution: inferences from plant rDNA substitution rates in the western Pacific. Evolution 57: 2893–8.

Wright SD, Keeling J, Gillman L (2006)The road from Santa Rosalia: a faster tempo of evolution in tropical climates. Proc Natl Acad Sci USA 103:7718-22.

Wright SD, Gillman LN, Ross HA, Keeling DJ (2010) Energy and the tempo of evolution in amphibians. Global Ecol Biogeogr 19: 733-40.

Wright SD, Ross HA, Keeling DJ, McBride P,Gillman LN (2011) Thermal energy and the rate of genetic evolution in marine fishes. Evol Ecol25: 525-30.

Wright SI, Lauga B, Charlesworth D (2002) Rates and patterns of molecular evolution in inbred and outbred Arabidopsis. Mol Biol Evol 19: 1407–20.

Wright SI, Yau CB, Looseley M, Meyers BC (2004) Effects of gene expression on molecular evolution in Arabidopsis thaliana and Arabidopsis lyrata. MolBiol Evol 21: 1719–26.

Wright SL, Crawford CB, Anderson JL (1988) Allocation of reproductive effort in Mus domesticus: Response of offspring sex ratio and quality to social density and food availability. Behav Ecol Sociobiol 23: 357-65.

Wu BJ, Else PL, Storlien LH, Hulbert AJ (2001) Molecular activity of Na/K-ATPase from different sources is related to the packing of membrane lipids. J Exp Biol 204: 4271–80.

Wu BJ, Hulbert AJ, Storlien LH, Else PL (2004) Membrane lipids and sodium pumps of cattle and crocodiles: an experimental test of the membrane pacemaker theory of metabolism. Am J Physiol Regul Integr Comp Physiol 287: R633–R641.

Wu C, Shen Y, Tang T (2009) Evolution under canalization and the dual roles of microRNAs - A hypothesis. Genome Res 19: 734–43.

Wu H, Sun YE (2006)Epigenetic regulation of stem cell differentiation.Pediatr Res 59: 21R-5R.

Wu H, Zhang Y (2011)Mechanisms and functions of Tet protein-mediated 5-methylcytosine oxidation.Genes Dev 25: 2436-52.

Wu Q, Song R, Ortogero N, Zheng H, Evanoff R, et al. (2012) The RNase III enzyme DROSHA is essential for microRNA production and spermatogenesis. J Biol Chem 287: 25173–90.

Wu RSS, Zhou BS, Randall DJ, Woo NYS,Lam PKS (2003) Aquatic hypoxia is an endocrine disruptor and impairs fish reproduction. Environ SciTechnol 37: 1137–41.

Wu SY, McLeod M (1995) The sak1+ gene of Schizosaccharomyces pombe encodes an RFX family DNA-binding protein that positively regulates cyclic-AMP-dependent protein kinase-mediated exit from the mitotic cell cycle. Mol Cell Biol 15: 1479-88.

Wu X, Bishopric NH, Discher DJ, Murphy BJ, Webster KA (1996)Physical and functional sensitivity of zinc finger transcription factors to redox change.Mol Cell Biol 16: 1035-46.

Wu X, Tanwar PS, Raftery LA (2008) Drosophila follicle cells: morphogenesis in an eggshell. Semin Cell Dev Biol 19: 271–82.

Wulff C, Wiegand SJ, Saunders PT, Scobie GA, Fraser HM (2001) Angiogenesis during follicular development in the primate and its inhibition by treatment with truncated Flt-1-Fc (vascular endothelial growth factor Trap(A40)). Endocrinology 142: 3244–54.

Wulff C, Wilson H, Wiegand SJ, Rudge JS, Fraser HM (2002) Prevention of thecal angiogenesis, antral follicular growth, and ovulation in the primate by treatment with vascular endothelial growth factor Trap R1R2. Endocrinology 143: 2797–807.

Wurdak ES, Gilbert JJ, Jagels R (1978) Fine structure of the resting eggs of the rotifers Brachionus calyciflorus and Asplanchna sieboldi. Trans Am Microsc Soc 97: 49–72.

Wyckoff GJ, Wang W, Wu CI (2000) Rapid evolution of male reproductive genes in the descent of man. Nature 403: 304-9.

Wyles JS, Kunkel JG, Wilson AC (1983) Birds, behavior, and anatomical evolution. ProcNatl Acad Sci USA 80: 4394-7.

Wylie C (1999)Germ cells.Cell 96: 165-74.

Wynne-Edwards VC (1962) Animal dispersion in relation to social behavior. New York, NY: Hafner Publishing Co.

Xanthoudakis S, Curran T (1992) Identification and characerization of Ref-1, a nuclear protein that facilitates AP-1 DNA binding activity. EMBO J 11: 653–65.

Xanthoudakis S, Miao G, Wang F, Pan YC, Curran T (1992) Redox activation of Fos-Jun DNA binding activity is mediated by a DNA repair enzyme. EMBO J 11: 3323–35.

Xanthoudakis S, Miao GG, Curran T (1994) The redox and DNA-repair activities of Ref-1 are encoded by nonoverlapping domains. Proc Natl Acad Sci USA 91: 23–7.

Xi Z, Gavotte L, Xie Y, Dobson SL (2008) Genome-wide analysis of the interaction between the endosymbiotic bacterium Wolbachia and its Drosophila host. BMC Genomics 9: 1.

Xia X, Kung AL (2009)Preferential binding of HIF-1 to transcriptionally active loci determines cell-type specific response to hypoxia. Genome Biol 10: R113.

Xia X, Lemieux ME, Li W, Carroll JS, Brown M, Liu XS, Kung AL (2009) Integrative analysis of HIF binding and transactivation reveals its role in maintaining histone methylation homeostasis. Proc Natl Acad Sci USA 106:4260-5.

Xiao A, Li H, Shechter D, Ahn SH, Fabrizio LA, et al. (2009) WSTF regulates the H2A.X DNA damage response via a novel tyrosine kinase activity. Nature 457: 57–62.

Xiao S, Laflamme M (2009) On the eve of animal radiation: phylogeny, ecology and evolution of the Ediacara biota. Trends Ecol Evol 24: 31–40.

Xiao X, Zuo X, Davis AA, McMillan DR, Curry BB, Richardson JA, Benjamin IJ (1999) HSF1 is required for extra-embryonic development, postnatal growth and protection during inflammatory responses in mice. EMBO J 18: 5943–52.

Xie CH, Naito A, Mizumachi T, Evans TT, Douglas MG, et al. (2007) Mitochondrial regulation of cancer associated nuclear DNA methylation. Biochem Biophys Res Commun 364: 656-61.

Xie M, Roy R (2012)Increased levels of hydrogen peroxide induce a HIF-1-dependent modification of lipid metabolism in AMPK compromised C. elegans dauer larvae.Cell Metab 16: 322-35.

Xie S, Wang Q, Wu H, Cogswell J, Lu L, Jhanwar-Uniyal M, Dai W. (2001)Reactive oxygen species-induced phosphorylation of p53 on serine 20 is mediated in part by polo-like kinase-3.J Biol Chem 276: 36194-9.

Xie X, Lu J, Kulbokas EJ, Golub TR, Mootha V, et al. (2005) Systematic discovery of regulatory motifs in human promoters and 3' UTRs by comparison of several mammals. Nature 434: 338-45.

Xie Z, Johansen LK, Gustafson AM, Kasschau KD, Lellis AD, et al.(2004) Genetic and functional diversification of small RNA pathways in plants. PLoS Biol 2: 1–11.

Xing Y, Lee C (2005) Evidence of functional selection pressure for alternative splicing events that accelerate evolution of protein subsequences. Proc Natl AcadSci USA 102: 13526–31.

Xiong Y, Hales DB (1994) Immune-endocrine interactions in the mouse testis: cytokine-mediated inhibition of Leydig cell steroidogenesis. Endocrine J 2: 223–8.

Xiong Y, Fang JH, Yun JP, Yang J, Zhang Y, et al. (2010) Effects of microRNA-29 on apoptosis, tumorigenicity, and prognosis of hepatocellular carcinoma. Hepatology 51: 836–45.

Xiong Z, Gaeta RT, Pires JC (2011) Homoeologous shuffling and chromosome compensation maintain genome balance in resynthesized allopolyploid Brassica napus. Proc Natl Acad Sci USA 108: 7908-13.

Xu G, Goodridge AG (1998) A CT repeat in the promoter of the chicken malic enzyme gene is essential for function at an alternative transcription start site. Arch Biochem Biophys 358: 83-91.

Xu G, Intano GW, McCarrey JR, Walter RB, McMahan CA, Walter CA (2008) Recovery of a low mutant frequency after ionizing radiation-induced mutagenesis during spermatogenesis. Mutat Res 654: 150–7.

Xu G, Vogel KS, McMahan CA, Herbert DC, Walter CA (2010) BAX and tumor suppressor TRP53 are important in regulating mutagenesis in spermatogenic cells in mice. Biol Reprod 83: 979-87.

Xu GL, Bestor TH, Bourc’his D, Hsieh CL, Tommerup N, et al. (1999) Chromosome instability and immunodeficiency syndrome caused by mutations in a DNA methyltransferase gene. Nature 402: 187-91.

Xu H, Gui J, Hong Y (2005)Differential expression of vasa RNA and protein during spermatogenesis and oogenesis in the gibel carp (Carassius auratus gibelio), a bisexually and gynogenetically reproducing vertebrate.Dev Dyn 233: 872-82.

Xu J, Wang Q (2010) Mechanisms of last male precedence in a moth: sperm displacement at ejaculation and storage sites. Behav Ecol 21: 714–21.

Xu JP(2004a) The prevalence and evolution of sex in microorganisms. Genome 47: 775–80.

Xu JP (2004b) Genotype-environment interactions of spontaneous mutations for vegetative fitness in the human pathogenic fungus Cryptococcus neoformans. Genetics 168: 1177–88.

Xu KP, Yadav BR, Rorie RW, Plante L, Betteridge KJ, King WA (1992) Development and viability of bovine embryos derived from oocytes matured and fertilized in vitro and co-cultured with bovine oviductal epithelial cells. J Reprod Fertil 94: 33-43.

Xu W, Han Z (2008) Cloning and phylogenetic analysis of sid-1-like genes from aphids. J Insect Sci 8: 1–6.

Xu Y, Moore DH, Broshears J, Liu L, Wilson TM, Kelley MR (1997) The apurinic/apyrimidinic endonuclease (APE/ref-1) DNA repair enzyme is elevated in premalignant and malignant cervical cancer. Anticancer Res 17: 3713–9.

Xu Y, Wu F, Tan L, Kong L, Xiong L, et al. (2011)Genome-wide regulation of 5hmC, 5mC, and gene expression by Tet1 hydroxylase in mouse embryonic stem cells.Mol Cell 42: 451-64.

Xue Y, Wang Q, Long Q, Ng BL, Swerdlow H, et al. (2009) Human Y chromosome base-substitution mutation rate measured by direct sequencing in a deep-rooting pedigree. Curr Biol 19: 1453–7.

Yacobi K, Tsafriri A, Gross A (2007) Luteinizing hormone-induced caspase activation in rat preovulatory follicles is coupled to mitochondrial steroidogenesis. Endocrinology 148: 1717–26.

Yaeram J, Setchell BP, Maddocks S (2006) Effect of heat stress on the fertility of male mice in vivo and in vitro. Reprod Fertil Dev 18: 647–53.

Yamada A, Masutani C, Iwai S, Hanaoka F (2000) Complementation of defective translesion synthesis and UV light sensitivity in xeroderma pigmentosum variant cells by human and mouse DNA polymerase eta. Nucleic Acids Res 28: 2473–80.

Yamada Y, Davis KD, Coffman CR (2008)Programmed cell death of primordial germ cells in Drosophila is regulated by p53 and the Outsiders monocarboxylate transporter.Development 135: 207-16.

Yamamoto H, Adachi Y, Taniguchi H, Kunimoto H, Nosho K, Suzuki H, Shinomura Y (2012)Interrelationship between microsatellite instability and microRNA in gastrointestinal cancer.World J Gastroenterol 18: 2745-55.

Yamamoto N, Yoneda K, Asai K, Sobue K, Tada T, Fujita Y, et al. (2001) Alterations in the expression of the AQP family in cultured rat astrocytes during hypoxia and reoxygenation. Brain Res Mol Brain Res 90: 26–38.

Yamamoto T, Hikino T, Nakayama Y, Abé S (1999)Newt RAD51: cloning of cDNA and analysis of gene expression during spermatogenesis.Dev Growth Differ 41:401-6.

Yamamoto T, Kuramoto H, Kadowaki M (2007) Downregulation in aquaporin 4 and aquaporin 8 expression of the colon associated with the induction of allergic diarrhea in a mouse model of food allergy. Life Sci 81:115–20.

Yamamoto Y, Sasaki T, Tokoro M (1999) Adaptability of Darwinian and Lamarckian populations toward an unknown new world. In: Floreano D, Nicoud J, Mondada F,eds. Advances in Artificial Life. Berlin, Germany: Springer.pp 39-48.

Yamanaka S, Blau HM (2010) Nuclear reprogramming to a pluripotent state by three approaches. Nature 465: 704–12.

Yamasaki K, Nakasa T, Miyaki S, Yamasaki T, Yasunaga Y, Ochi M (2012)Angiogenic microRNA-210 is present in cells surrounding osteonecrosis.J Orthop Res 30: 1263-70.

Yamasaki S, Anderson P (2008)Reprogramming mRNA translation during stress.Curr Opin Cell Biol 20: 222-6.

Yamazaki K, Boyse EA, Mike V, Thaler HT, Mathieson BJ, et al. (1976) Control of mating preferences in mice by genes in the major histocompatibility complex. J Exp Med 144: 1324-35.

Yamazaki K, Beauchamp GK (2007)Genetic basis for MHC-dependent mate choice.Adv Genet 59: 129-45.

Yamazaki Y, Mann MR, Lee SS, Marh J, McCarrey JR, et al. (2003) Reprogramming of primordial germ cells begins before migration into the genital ridge, making these cells inadequate donors for reproductive cloning. Proc Natl Acad Sci USA 100: 12207–12.

Yan G, Severson DW, Christensen BM (1997) Costs and benefits of mosquito refractoriness to malaria parasites: implications for genetic variability of mosquitoes and genetic control of malaria. Evolution 51: 441–50.

Yan L, Fu D, Li C, Blechl A, Tranquilli G, et al. (2006) The wheat and barley vernalization gene VRN3 is an orthologue of FT. Proc Natl Acad Sci USA 103: 19581–6.

Yan Y, Wei CL, Zhang WR, Cheng HP, Liu J (2006) Cross-talk between calcium and reactive oxygen species signaling.Acta Pharmacol Sin 27: 821-6.

Yan Y, Zhang Y, Yang K, Sun Z, Fu Y, et al. (2011) Small RNAs from MITE-derived stem-loop precursors regulate abscisic acid signaling and abiotic stress responses in rice. Plant J 65: 820–8.

Yang AS, Jones PA, Shibata A (1996a) The mutational burden of 5-methylcytosine. In: Riggs AD, Martienssen RA, Russo VEA; eds. Epigenetic mechanisms of gene regulation. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press.pp 77–94.

Yang AS, Gonzalgo ML, Zingg JM, Millar RP, Buckley JD, Jones PA (1996b) The rate of CpG mutation in Alu repetitive elements within the p53 tumor suppressor gene in the primate germline. J Mol Biol 258: 240–50.

Yang B, Song Y, Zhao D, Verkman AS (2005) Phenotype analysis of aquaporin-8 null mice. Am J Physiol Cell Physiol 288: C1161–70.

Yang B, Zhao D, Verkman AS (2006) Evidence against functionally significant aquaporin expression in mitochondria. J Biol Chem 281: 16202–6.

Yang D, Elner SG, Bian ZM, Till GO, Petty HR, Elner VM (2007) Pro-inflammatory cytokines increase reactive oxygen species through mitochondria and NADPH oxidase in cultured RPE cells. Exp Eye Res 85: 462–72.

Yang H, Magilnick N, Ou X, Lu SC (2005)Tumour necrosis factor alpha induces co-ordinated activation of rat GSH synthetic enzymes via nuclear factor kappaB and activator protein-1.Biochem J 391: 399-408.

Yang HP, Tanikawa AY, Van Voorhies WA, Silva JC, Kondrashov AS (2001) Whole-genome effects of ethyl methanesulfonate-induced mutation on nine quantitative traits in outbred Drosophila melanogaster. Genetics 157: 1257-65.

Yang J, Su AI, Li WH (2005) Gene expression evolves faster in narrowly than in broadly expressed mammalian genes. Mol Biol Evol 22: 2113-8.

Yang J, Ledaki I, Turley H, Gatter KC, Montero JC, et al. (2009)Role of hypoxia-inducible factors in epigenetic regulation via histone demethylases.Ann NY Acad Sci 1177:185-97.

Yang M, Gao F, Liu H, Yu WH, Sun SQ (2009) Temporal changes in expression of aquaporin-3, -4, -5 and -8 in rat brains after permanent focal cerebral ischemia. Brain Res 1290: 121-32.

Yang MY, Fortune JE (2008) The capacity of primordial follicles in fetal bovine ovaries to initiate growth in vitro develops during mid-gestation and is associated with meiotic arrest of oocytes. Biol Reprod 78: 1153–61.

Yang S, Misner BJ, Chiu RJ, Meyskens FL Jr (2007) Redox effector factor-1, combined with reactive oxygen species, plays an important role in the transformation of JB6 cells. Carcinogenesis 28: 2382-90.

Yang SL, Hsu C, Hsu HK, Liu KM, Peng MT (1993) Effects of long-term estradiol exposure on the hypothalamic neuron number. Acta Endocrinol 129: 543–7.

Yang W, Woodgate R (2007) What a difference a decade makes: insights into translesion DNA synthesis. Proc Natl Acad Sci USA 104: 15591–8.

Yang Y, Sterling J, Storici F, Resnick MA, Gordenin DA (2008) Hypermutability of damaged single-strand DNA formed at double-strand breaks and uncapped telomeres in yeast Saccharomyces cerevisiae. PLoS Genet 4: 1000264.

Yannopoulos G, Stamatis N, Monastirioti M, Hatzopoulos P, Louis C (1987) Hobo is responsible for the induction of hybrid dysgenesis by strains of Drosophila melanogaster bearing the male recombination factor 23.5mrf. Cell 49: 487–95.

Yao KS, Xanthoudakis S, Curran T, O'Dwyer PJ (1994) Activation of AP-1 and of a nuclear redox factor, Ref-1, in the response of HT29 colon cancer cells to hypoxia. Mol Cell Biol 14: 5997–6003.

Yao MC, Chao JL (2005) RNA-guided DNA deletion in Tetrahymena: An RNAi-based mechanism for programmed genome rearrangements. Annu Rev Genet 39: 537–59.

Yarmolinsky MB (1995) Programmed cell death in bacterial populations. Science 267: 836-7.

Yasbin RE, Cheo D, Bayles KW (1991) The SOB system of Bacillus subtilis: a global regulon involved in DNA repair and differentiation. Res Microbiol 142: 885-92.

Yasbin RE, Cheo DL, Bayles KW (1992) Inducible DNA repair and differentiation in Bacillus subtilis: interactions between global regulons. Mol Microbiol 6: 1263-70.

Yasui Y (1997) A “good-sperm” model can explain the evolution of costly multiple mating by females. Am Nat 149:573–84.

Yazawa H, Sasagawa I, Nakada T (1999) Effect of immobilization stress on testicular germ cell apoptosis in rats. Hum Reprod 14: 1806–10.

Yazawa H, Sasagawa I, Nakada T (2000) Apoptosis of testicular germ cells induced by exogenous glucocorticoid in rats. Hum Reprod 15: 1917–20.

Yazawa H, Sasagawa I, Suzuki Y, Nakada T (2001)Glucocorticoid hormone can suppress apoptosis of rat testicular germ cells induced by testicular ischemia.Fertil Steril 75: 980-5.

Yazawa T, Yamamoto K, Kikuyama S, Abé S-I (1999) Elevation of plasm prolactin concentrations by low temperature is the cause of spermatogonial cell death in the newt, Cynops pyrrhogaster. Gen Comp Endocrinol 103: 302–11.

Yazawa T, Yamamoto T, Abé S-I (2000) Prolactin induces apoptosis in the penultimate spermatogonial stage of the testes in japanese red-bellied newt (Cynops pyrrhogaster). Endocrinology 141: 2027–32.

Yazawa T, Nakayama Y, Fujimoto K, Matsuda Y, Abe K, et al.(2003) Abnormal spermatogenesis at low temperatures in the japanese redbellied newt, Cynops pyrrhogaster: possible biological significance of the cessation of spermatocytogenesis. Mol Reprod Dev 66: 60–6.

Ye L, Li X, Kong X, Wang W, Bi Y, et al.(2005) Hypomethylation in the promoter region of POMC gene correlates with ectopic overexpression in thymic carcinoids. J Endocrinol 185: 337–43.

Yeh PJ, Price TD (2004) Adaptive phenotypic plasticity and the successful colonization of a novel environment. Am Nat 164: 531–42.

Yeiser B, Pepper ED, Goodman MF, Finkel SE (2002) SOS-induced DNA polymerases enhance long-term survival and evolutionary fitness. Proc Natl Acad Sci USA 99: 8737–41.

Yeung C-H (2010) Aquaporins in spermatozoa and testicular germ cells: identification and potential role. Asian J Androl 12: 490-9.

Yeung CH, Callies C, Rojek A, Nielsen S, Cooper TG (2009) Aquaporin isoforms involved in physiological volume regulation of murine spermatozoa. Biol Reprod 80: 350–7.

Yeyati PL, Bancewicz RM, Maule J, van Heyningen V (2007) Hsp90 selectively modulates phenotype in vertebrate development. PLoS Genet 3: e43.

Yi L, Lu C, Hu W, Sun Y, Levine AJ (2012)Multiple roles of p53-related pathways in somatic cell reprogramming and stem cell differentiation.Cancer Res 72: 5635-45.

Yim MB, Chock PB, Stadtman ER (1990) Copper, zinc superoxide dismutase catalyzes hydroxyl radical production from hydrogen peroxide. Proc Natl Acad Sci USA 87: 5006–10.

Yim MB, Chock PB, Stadtman ER (1993) Enzyme function of copper, zinc superoxide dismutase as a free radical generator. J Biol Chem 268: 4099–105.

Yin H, Lin H (2007) An epigenetic activation role of Piwi and a Piwi-associated piRNA in Drosophila melanogaster. Nature 450: 304–8.

Yin Y, Stahl BC, DeWolf WC, Morgentaler A (1998) p53-mediated germ cell quality control in spermatogenesis. Dev Biol 204: 165-71.

Yin Y, Stahl BC, DeWolf WC, Morgentaler A (2002)P53 and Fas are sequential mechanisms of testicular germ cell apoptosis.J Androl 23: 64-70.

Yin YZ, Hawkins KL, Dewolf WC, Morgentaler A (1997) Heat stress causes testicular germ cell apoptosis in adult mice. J Androl 18: 159-65.

Yoder JA, Walsh CP, Bestor TH (1997) Cytosine methylation and the ecology of intragenomic parasites. Trends Genet 13: 335–40.

Yokota S (2008) Historical survey on chromatoid body research. Acta Histochem Cytochem 41: 65–82.

Yomo T, Sato K, Ito Y, Kaneko K (2006)Responses of fluctuating biological systems. Lect Notes Comput Sci 3853: 107-12.

Yonei S, Yokota R, Sato Y (1987) The distinct role of catalase and DNA repair systems in protection against hydrogen peroxide in Escherichia coli. Biochem Biophys Res Commun 143: 638-44.

Yoo BH (1980a) Long-term selection for a quantitative character in large replicate populations of Drosophila melanogaster. I. Response to selection. Genet Res 35: 1–17.

Yoo BH (1980b) Long-term selection for a quantitative character in large replicate populations of Drosophila melanogaster. II. Lethals and visible mutants with large effects. Genet Res 35: 19–31.

Yoo BH (1980c) Long-term selection for a quantitative character in large replicate populations of Drosophila melanogaster. III. The nature of residual genetic variability. Theor Appl Genet 57: 25–32.

Yoon DJ, Sklar C, David R (1988) Presence of immunoreactive corticotropin-releasing factor in the testis. Endocrinology 122: 759-61.

Yoon SR, Qin J, Glaser RL, Jabs EW, Wexler NS, et al. (2009) The ups and downs of mutation frequencies during aging can account for the Apert syndrome paternal age effect. PLoS Genet 5: e1000558.

Yoon SR, Choi SK, Eboreime J, Gelb BD, Calabrese P, Arnheim N (2013)Age-dependent germline mosaicism of the most common Noonan syndrome mutation shows the signature of germline selection.Am J Hum Genet 92: 917–26.

Yorimitsu T, Klionsky DJ (2007)Eating the endoplasmic reticulum: quality control by autophagy.Trends Cell Biol 17: 279-85.

Yoshida M, Ishigaki K, Nagai T, Chikyu M, Pursel VG (1993) Glutathione concentration during maturation and after fertilization in pig oocytes: relevance to the ability of oocytes to form male pronucleus. Biol Reprod 49: 89–94.

Yoshida T, Jones LE, Ellner SP, Fussman GF, Hairston NG Jr (2003) Rapid evolution drives ecological dynamics in a predator-prey system. Nature 424: 303–6.

Yoshida-Noro C, Tochinai S (2010)Stem cell system in asexual and sexual reproduction of Enchytraeus japonensis (Oligochaeta, Annelida). Dev Growth Differ 52: 43-55.

Yoshimura J, Clark CW (1991) Individual adaptations in stochastic environments. Evol Ecol 5: 173–92.

Yoshimura J, Shields WM (1992) Components of uncertainty in clutch-size optimization. Bull Math Biol 54: 445–64.

Yoshimura J, Jansen VAA (1996)Evolution and population dynamics in stochastic environments.Res Popul Ecol 38: 165-82.

Yoshimura J, Tanaka Y, Togashi T, Iwata S, Tainaka K (2009) Mathematical equivalence of geometric mean fitness with probabilistic optimization under environmental uncertainty. Ecol Model 220: 2611–7.

Yoshinaga K, Nishikawa S, Ogawa M, Hayashi S, Kunisada T, et al. (1991)Role of c-kit in mouse spermatogenesis: identification of spermatogonia as a specific site of c-kit expression and function.Development 113: 689-99.

Yoshinaga T, Hagiwara A, Tsukamoto K (2000) Effect of periodical starvation on the life history of Brachionus plicatilis O.F. Müller (Rotifera): a possible strategy for population stability. J Exp Mar Biol Ecol 253: 253–60.

Yoshinaga T, Hagiwara A, Tsukamoto K (2001) Effect of periodical starvation on the survival of offspring in the rotifer Brachionus plicatilis. Fish Sci 67: 373–4.

Yoshinaga T, Kaneko G, Kinoshita S, Tsukamoto K, Watabe S (2003)The molecular mechanisms of life history alterations in a rotifer: a novel approach in population dynamics.Comp Biochem Physiol B Biochem Mol Biol 136: 715-22.

Youn CK, Kim SH, Lee do Y, Song SH, Chang IY, et al. (2005) Cadmium down-regulates human OGG1 through suppression of Sp1 activity. J Biol Chem 280: 25185–95.

Young HE (2004)Existence of reserve quiescent stem cells in adults, from amphibians to humans.Curr Top Microbiol Immunol 280: 71-109.

Young JA, Yourth CP, Agrawal AF (2009) The effects of pathogens on selection against deleterious mutations in Drosophila melanogaster. J Evol Biol. 22: 2125–9.

Young JC, Moarefi I, Hartl FU (2001) Hsp90: a specialized but essential protein folding tool. J Cell Biol 154: 267-74.

Young JZ (1951) Doubt and certainty in science. Oxford, UK: Oxford University Press.

Young KA, Nelson RJ (2000) Short photoperiods reduce vascular endothelial growth factor in the testes of Peromyscus leucopus. Am J Physiol 279: R1132–R1137.

Young LE (2001) Imprinting of genes and the Barker hypothesis. Twin Res 4: 307–17.

Young SL, Diolaiti D, Conacci-Sorrell M, Ruiz-Trillo I, Eisenman RN, King N (2011)Premetazoan ancestry of the Myc-Max network.Mol Biol Evol 28: 2961-71.

Youngson NA, Whitelaw E (2008)Transgenerational epigenetic effects.Annu Rev Genomics Hum Genet 9:233-57.

Younts TJ, Hughes FM Jr (2009) Emerging role of water channels in regulating cellular volume during oxygen deprivation and cell death. In: Haddad GG, S.P. Yu SP, eds. Brain Hypoxia and Ischemia. Humana Press. pp 79-96.

Yu N, Zhao Z, Fu YX, Sambuughin N, Ramsay M, et al. (2001) Global patterns of human DNA sequence variation in a 10-kb region on chromosome 1. Mol Biol Evol 18: 214–22.

Yu B, Lin H, Yang L, Chen K, Luo H, et al. (2012)Genetic variation in the Nrf2 promoter associates with defective spermatogenesis in humans.J Mol Med (Berl) 90: 1333-42.

Yu BP (1996) Aging and oxidative stress: modulation by dietary restriction. Free Radic Biol Med 21: 651-68.

Yu MC, Lamming DW, Eskin JA, Sinclair DA, Silver PA (2006) The role of protein arginine methylation in the formation of silent chromatin. Genes Dev 20: 3249–54.

Yu NW, Hsu CY, Ku HH, Chang LT, Liu HW (1993) Gonadal differentiation and secretions of estradiol and testosterone of the ovaries of Rana catesbeiana tadpoles treated with 4-hydroxyandrostene-dione. J Exp Zool 265: 252–7.

Yu X, Harris SL, Levine AJ (2006)The regulation of exosome secretion: a novel function of the p53 protein.Cancer Res 66: 4795-801.

Yu Y, Dumollard R, Rossbach A, Lai FA, Swann K (2010) Redistribution of mitochondria leads to bursts of ATP production during spontaneous mouse oocyte maturation. J Cell Physiol 224: 672-80.

Yu Z, Guo R, Ge Y, Ma J, Guan J, et al.(2003) Gene expression profiles in different stages of mouse spermatogenic cells during spermatogenesis. Biol Reprod 69:37–47.

Yuan J, Narayanan L, Rockwell S, Glazer PM (2000) Diminished DNA repair and elevated mutagenesis in mammalian cells exposed to hypoxia and low pH. Cancer Res 60: 4372–6.

Yuan LW, Keil RL (1990) Distance-independence of mitotic intrachromosomal recombination in Saccharomyces cerevisiae. Genetics 124: 263-73.

Yue L, Karr TL, Nathan DF, Swift H, Srinivasan S, Lindquist S (1999) Genetic analysis of viable Hsp90 alleles reveals a critical role in Drosophila spermatogenesis. Genetics 151:1065-79.

Yueh A, Schneider RJ (2000) Translation by ribosome shunting on adenovirus and hsp70 mRNAs facilitated by complementarity to 18S rRNA. Genes Dev 14: 414–21.

Yukilevich R, Lachance J, Aoki F, True JR (2008) Long-term adaptation of epistatic genetic networks. Evolution 62: 2215–35.

Yund PO (2000) How severe is sperm limitation in natural populations of marine free-spawners? Trends Ecol Evol 15: 10–13.

Yund PO, McCartney MA (1994) Male reproductive success in sessile invertebrates: competition for fertilizations. Ecology 75: 2168-84.

Yurtçu M, Abasiyanik A, Avunduk MC, Muhtaroglu S (2008) Effects of melatonin on spermatogenesis and testicular ischemia-reperfusion injury after unilateral testicular torsion-detorsion. J Pediatr Surg 43: 1873-8.

Yuste E, Sánchez-Palomino S, Casado C, Domingo E, López-Galíndez C (1999) Drastic fitness loss in human immunodeficiency virus type 1 upon serial bottleneck events. J Virol 73: 2745-51.

Zachos J, Pagani M, Sloan L, Thomas E, Billups K (2001) Trend, rhythms, and aberrations in global climates 65 Ma to present. Science 292: 686–93.

Zackova M, Skobisová E, Urbánková E, Jezek P (2003)Activating omega-6 polyunsaturated fatty acids and inhibitory purine nucleotides are high affinity ligands for novel mitochondrial uncoupling proteins UCP2 and UCP3.J Biol Chem 278: 20761-9.

Zadravec D, Tvrdik P, Guillou H, Haslam R, Kobayashi T, et al. (2011) ELOVL2 controls the level of n-6 28:5 and 30:5 fatty acids in testis, a prerequisite for male fertility and sperm maturation in mice. J Lipid Res 52: 245–55.

Zafari AM, Ushio-Fukai M, Akers M, Yin Q, Shah A, et al. (1998) Role of NADH/NADPH oxidase-derived H2O2 in angiotensin II-induced vascular hypertrophy. Hypertension 32: 488-95.

Zagórska A, Dulak J (2004)HIF-1: the knowns and unknowns of hypoxia sensing.Acta Biochim Pol 51: 563-85.

Zaiton Z, Merican Z, Khalid BA, Mohamed JB, Baharom S (1993) The effects of propranolol on skeletal muscle contraction, lipid peroxidation products and antioxidant activity in experimental hyperthyroidism. Gen Pharmacol 24: 195–9.

Zajitschek SRK, Lindholm AK, Evans JP, Brooks RC (2009) Experimental evidence that high levels of inbreeding depress sperm competitiveness. J Evol Biol 22: 1338–45.

Zakeri ZF, Wolgemuth DJ, Hunt CR (1988) Identification and sequence analysis of a new member of the mouse HSP70 gene family and characterization of its unique cellular and developmental pattern of expression in the male germ line. Mol Cell Biol 8:2925–32.

Zaky A, Busso C, Izumi T, Chattopadhyay R, Bassiouny A, et al. (2008)Regulation of the human AP-endonuclease (APE1/Ref-1) expression by the tumor suppressor p53 in response to DNA damage.Nucleic Acids Res 36: 1555-66.

Zalata A, Hafez T, Schoonjans F, Comhaire F (1996)The possible meaning of transferrin and its soluble receptors in seminal plasma as markers of the seminiferous epithelium.Hum Reprod 11: 761-4.

Zalata AA, Christophe AB, Depuydt CE, Schoonjans F, Comhaire FH (1998) The fatty acid composition of phospholipids of spermatozoa from infertile patients. Mol Hum Reprod 4: 111–8.

Zambrano E, Martínez-Samayoa PM, Bautista CJ, Deás M, Guillén L, et al. (2005)Sex differences in transgenerational alterations of growth and metabolism in progeny (F2) of female offspring (F1) of rats fed a low protein diet during pregnancy and lactation.J Physiol 566: 225-36.

Zambrano MM, SiegeleDA, Almiron M, Tormo A,Kolter R (1993) Microbial competition: Escherichia coli mutants that take over stationary phase cultures. Science 259:1757–60.

Zambrano MM, Kolter R (1996) GASPing for life in stationary phase. Cell 86: 181–4.

Zamoner A, Barreto KP, Filho DW, Sell F, Woehl VM, et al. (2007) Hyperthyroidism in the developing rat testis is associated with oxidative stress and hyperphosphorylated vimentin accumulation. Mol Cell Endocrinol 267: 116–26.

Zamoner A, Barreto KP, Filho DW, Sell F, Woehl VM, et al. (2008) Propylthiouracil-induced congenital hypothyroidism upregulates vimentin phosphorylation and depletes antioxidant defenses in immature rat testis. J Mol Endocrinol 40: 125–35.

Zamoum T, Simon JC, Crochard D, Ballanger Y, Lapchin L, Vanlerberghe-Masutti F, Guillemaud T (2005) Does insecticide resistance alone account for the low genetic variability of asexually reproducing populations of the peach-potato aphid Myzus persicae? Heredity 94: 630–9.

Zamudio N, Bourc’his D (2010) Transposable elements in the mammalian germline: a comfortable niche or a deadly trap? Heredity 105: 92–104.

Zar HA, Lancaster JR Jr (2000) Mild hypothermia protects against postischemic hepatic endothelial injury and decreases the formation of reactive oxygen species. Redox Rep 5:303-10.

Zardoya R, Villalba S (2001)A phylogenetic framework for the aquaporin family in eukaryotes.J Mol Evol 52: 391-404.

Zarse K, Schmeisser S, Groth M, Priebe S, Beuster G, Kuhlow D, Guthke R, Platzer M, Kahn CR, Ristow M (2012)Impaired insulin/IGF1 signaling extends life span by promoting mitochondrial L-proline catabolism to induce a transient ROS signal.Cell Metab 15: 451-65.

Zatyka M, Thomas CM (1998) Control of genes for conjugative transfer of plasmids and other mobile elements. FEMS Microbiol Rev 21: 291-319.

Zbinden G (1980) Unscheduled DNA synthesis in the testis: a secondary test for the evaluation of chemical mutagens. Arch Toxicol 46: 139–49.

Zchori-Fein E, Gottlieb Y, Kelly SE, Brown JK, Wilson JM, Karr TL, et al. (2001) A newly discovered bacterium associated with parthenogenesis and change in host selection behavior in parasitoid wasps. Proc Natl Acad Sci USA 98: 12555–60.

Zchori-Fein E, Perlman SJ (2004) Distribution of the bacterial symbiont Cardinium in arthropods. Mol Ecol 13: 2009–16.

Zeh DW, Zeh JA (2000) Reproductive mode and speciation: the viviparity-driven conflict hypothesis. Bioessays 22: 938-46.

Zeh DW, Zeh JA, Ishida Y (2009) Transposable elements and an epigenetic basis for punctuated equilibria. Bioessays 31:715–26.

Zeh JA, Zeh DW (1996) The evolution of polyandry I: intragenomic conflict and genetic incompatibility. Proc R Soc B 263: 1711–7.

Zeh JA, Zeh DW (1997) The evolution of polyandry II: post-copulatory defences against genetic incompatibility. Proc R Soc Lond B 264: 69-75.

Zeh JA, Zeh DW (2001) Reproductive mode and the genetic benefits of polyandry. Anim Behav 61: 1051–63.

Zeh JA, Zeh DW (2005) Maternal inheritance, sexual conflict and the maladapted male. Trends Genet 21: 281–6.

Zelazowska M, Kilarski W, Bilinski SM, Podder DD, Kloc M (2007) Balbiani cytoplasm in oocytes of a primitive fish, the sturgeon Acipenser gueldenstaedtii, and its potential homology to the Balbiani body (mitochondrial cloud) of Xenopus laevis oocytes. Cell Tissue Res 329: 137-45.

Zeldovich KB, Chen P, Shakhnovich EI (2007)Protein stability imposes limits on organism complexity and speed of molecular evolution.Proc Natl Acad Sci USA 104: 16152-7.

Zeleznik AJ, Schuler HM, Reichert LE Jr (1981) Gonadotropin-binding sites in the rhesus monkey ovary: role of the vasculature in the selective distribution of human chorionic gonadotropin to the preovulatory follicle. Endocrinology 109: 356–62.

Zemojtel T, Kielbasa SM, Arndt PF, Chung HR, Vingron M (2009) Methylation and deamination of CpGs generate p53-binding sites on a genomic scale. Trends Genet 25: 63–6.

Zemojtel T, Kielbasa SM, Arndt PF, Behrens S, Bourque G, Vingron M (2011)CpG deamination creates transcription factor-binding sites with high efficiency.Genome Biol Evol 3: 1304-11.

Zenclussen ML, Jensen F, Rebelo S, El-Mousleh T, Casalis PA, Zenclussen AC (2012) Heme oxygenase-1 expression in the ovary dictates a proper oocyte ovulation, fertilization, and corpora lutea maintenance. Am J Reprod Immunol 67: 376–82.

Zendman AJ, Ruiter DJ, Van Muijen GN (2003)Cancer/testis-associated genes: identification, expression profile, and putative function.J Cell Physiol 194: 272-88.

Zeng HT, Ren Z, Yeung WS, Shu YM, Xu YW, et al. (2007) Low mitochondrial DNA and ATP contents contribute to the absence of birefringent spindle imaged with PolScope in in vitro matured human oocytes. Hum Reprod 22: 1681–6.

Zeng X, Zhou J, Vasseur C (2000) A strategy for controlling nonlinear systems using a learning automaton. Automatica 36: 1517–24.

Zenvirth D, Arbel T, Sherman A, Goldway M, Klein S, Simchen G (1992) Multiple sites for double-strand breaks in whole meiotic chromosomes of Saccharomyces cerevisiae. EMBO J 11: 3441–7.

Zenvirth D, Richler C, Bardhan A, Baudat F, Barzilai A, et al. (2003) Mammalian meiosis involves DNA double-strand breaks with 3' overhangs.Chromosoma 111: 369-76.

Zenzes MT, Puy LA, Bielecki R, Reed TE (1999) Detection of benzo[a]pyrene diol epoxide-DNA adducts in embryos from smoking couples: evidence for transmission by spermatozoa. Mol Hum Reprod 5: 125–31.

Zerani M, Amabili F, Mosconi G, Gobbetti A (1991) Effects of captivity stress on plasma steroid levels in the green frog, Rana esculenta, during the annual reproductive cycle. Comp Biochem Physiol A Comp Physiol 98: 491-6.

Zeyl C (2007) Evolutionary genetics: a piggyback ride to adaptation and diversity. Curr Biol 17:R333-5.

Zeyl C, Bell G, Green DM (1996) Sex and the spread of the retrotransposon Ty3 in experimental populations of Saccharomyces cerevisiae. Genetics 143: 1567–77.

Zeyl C, Bell G (1997) The advantage of sex in evolving yeast populations. Nature 388: 465-8.

Zeyl C, de Visser JAGM (2001) Estimates of the rate and distribution of fitness effects of spontaneous mutation in Saccharomyces cerevisiae. Genetics 157: 53-61.

Zeyl C, Mizesko M, de Visser JA (2001)Mutational meltdown in laboratory yeast populations.Evolution 55: 909-17.

Zhai J, Liu J, Liu B, Li P, Meyers BC, et al. (2008) Small RNA-directed epigenetic natural variation in Arabidopsis thaliana. PLoS Genet 4: e1000056.

Zhang G, Li J, Purkayastha S, Tang Y, Zhang H, et al. (2013) Hypothalamic programming of systemic ageing involving IKK-β, NF-κB and GnRH. Nature 497: 211–6.

Zhang J, Webb DM, Podlaha O (2002) Accelerated protein evolution and origins of human-specific features: Foxp2 as an example. Genetics 162: 1825-35.

Zhang J, He X (2005)Significant impact of protein dispensability on the instantaneous rate of protein evolution.Mol Biol Evol 22: 1147-55.

Zhang K, Shang Y, Liao S, Zhang W, Nian H, et al. (2007) Uncoupling protein 2 protects testicular germ cells from hyperthermia-induced apoptosis. Biochem Biophys Res Commun 360: 327–32.

Zhang L, Murphy PJ, Kerr A, Tate ME (1993) Agrobacterium conjugation and gene regulation by N-acyl-L-homoserine lactones. Nature 362: 446-8.

Zhang L, Li WH (2004) Mammalian housekeeping genes evolve more slowly than tissue-specific genes. Mol Biol Evol 21: 236–9.

Zhang L, Charron M, Wright WW, Chatterjee B, Song CS, et al. (2004) Nuclear factor-kappaB activates transcription of the androgen receptor gene in Sertoli cells isolated from testes of adult rats. Endocrinology 145: 781-9.

Zhang L, Han XK, Qi YY, Liu Y, Chen QS (2008) Seasonal effects on apoptosis and proliferation of germ cells in the testes of the Chinese soft-shelled turtle, Pelodiscus sinensis. Theriogenology 69:1148-58.

Zhang N, Zhang J, Cheng Y, Howard K (1996) Identification and genetic analysis of wunen, a gene guiding Drosophila melanogaster germ cell migration. Genetics 143: 1231–41.

Zhang N, Zhang J, Purcell KJ, Cheng Y, Howard K (1997) The Drosophila protein Wunen repels migrating germ cells. Nature 385: 64–67.

Zhang QM, Takemoto T, Mito S, Yonei S (1996) Induction of repair capacity for oxidatively damaged DNA as a component of peroxide stress response in Escherichia coli. J Radiat Res 37: 171-6.

Zhang R, Brennan ML, Shen Z, MacPherson JC, Schmitt D, et al. (2002) Myeloperoxidase functions as a major enzymatic catalyst for initiation of lipid peroxidation at sites of inflammation. J Biol Chem 277: 46116–22.

Zhang R, Wang YQ, Su B (2008) Molecular evolution of a primate-specific microRNA family. Mol Biol Evol 25: 1493-502.

Zhang Y, Lyver ER, Knight SA, Pain D, Lesuisse E, Dancis A (2006) Mrs3p, Mrs4p, and frataxin provide iron for Fe-S cluster synthesis in mitochondria. J Biol Chem 281: 22493-502.

Zhang YZ, Ouyang YC, Hou Y, Schatten H, Chen DY, Sun QY (2008) Mitochondrial behavior during oogenesis in zebrafish: a confocal microscopy analysis. Dev Growth Differ 50: 189–201.

Zhang Z, Hambuch T, Parsch J (2004a) Molecular evolution of sex-biased genes in Drosophila. J Mol Evol 21: 2130-9.

Zhang Z, Short RV, Meehan T, De Kretser DM, Renfree MB, Loveland KL (2004b) Functional analysis of the cooled rat testis. J Androl 25: 57–68.

Zhang Z, Sun H, Dai H, Walsh RM, Imakura M (2009) MicroRNA miR-210 modulates cellular response to hypoxia through the MYC antagonist MNT. Cell Cycle 8: 2756–68.

Zhang Z, Saier MH Jr (2009) A mechanism of transposon-mediated directed mutation. Mol Microbiol 74:29-43.

Zhang Z, Saier MH Jr (2011) Transposon-mediated adaptive and directed mutations and their potential evolutionary benefits. J Mol Microbiol Biotechnol 21: 59-70.

Zhang Z, Wang J, Shlykov MA, Saier MH Jr (2013) Transposon mutagenesis in disease, drug discovery, and bacterial evolution. In: Mittelman D, ed. Stress-Induced Mutagenesis. New York, NY: Springer. pp 59-77.

Zhao T, Li G, Mi S, Li S, Hannon GJ, et al. (2007) A complex system of small RNAs in the unicellular green alga Chlamydomonas reinhardtii. Genes Dev 21: 1190-203.

Zhao X, Ueba T, Christie BR, Barkho B, McConnell MJ, et al. (2003) Mice lacking methyl-CpG binding protein 1 have deficits in adult neurogenesis and hippocampal function. Proc Natl Acad Sci USA 100: 6777-82.

Zhao Y, Epstein RJ (2008) Programmed genetic instability: a tumor-permissive mechanism for maintaining the evolvability of higher species through methylation-dependent mutation of DNA repair genes in the male germ line. Mol Biol Evol 25: 1737–49.

Zheng H, Olive PL (1997) Influence of oxygen on radiation-induced DNA damage in testicular cells of C3H mice. Int J Radiat Biol 71: 275–82.

Zheng K, Xiol J, Reuter M, Eckardt S, Leu NA, et al. (2010) Mouse MOV10L1 associates with Piwi proteins and is an essential component of the Piwi-interacting RNA (piRNA) pathway. Proc Natl Acad Sci USA 107: 11841-6.

Zheng M, Doan B, Schneider TD, Storz G (1999) OxyR and SoxRS regulation of fur. J Bacteriol 181: 4639-4643

Zheng P, Schramm RD, Latham KE (2005) Developmental regulation and in vitro culture effects on expression of DNA repair and cell cycle checkpoint control genes in rhesus monkey oocytes and embryos. Biol Reprod 72: 1359–69.

Zheng S, Turner TT, Lysiak JJ (2006) Caspase 2 activity contributes to the initial wave of germ cell apoptosis during the first round of spermatogenesis. Biol Reprod 74:1026–33.

Zhivotovsky LA (1997) Environmental stress and evolution: a theoretical study. EXS 83: 241–54.

Zhong L, D'Urso A, Toiber D, Sebastian C, Henry RE, et al. (2010) The histone deacetylase Sirt6 regulates glucose homeostasis via Hif1alpha. Cell 140: 280-93.

Zhong W, Priest N (2011) Stress-induced recombination and the mechanism of evolvability. Behav Ecol Sociobiol 65: 493-502.

Zhou BB, Elledge SJ (2000) The DNA damage response: putting checkpoints in perspective. Nature 408: 433–9.

Zhou J, Fandrey J, Schumann J, Tiegs G, Brune B (2003) NO and TNFalpha released from activated macrophages stabilize HIF-1alpha in resting tubular LLC-PK1 cells. Am J Physiol Cell Physiol 284:C439–C446.

Zhou RR, Wang B, Wang J, Schatten H, Zhang YZ (2010) Is the mitochondrial cloud the selection machinery for preferentially transmitting wild-type mtDNA between generations? Rewinding Muller’s ratchet efficiently. Curr Genet 56: 101–7.

Zhou X, Zuo Z, Zhou F, Zhao W, Sakaguchi Y, et al. (2010) Profiling sex-specific piRNAs in zebrafish. Genetics. 2010 Dec;186: 1175-85.

Zhou ZQ, Walter CA (1995) Expression of the DNA repair gene XRCC1 in baboon tissues. Mutat Res 348: 111–6.

Zhou ZQ, Hurlin PJ (2001) The interplay between Mad and Myc in proliferation and differentiation. Trends Cell Biol 11: S10-4.

Zhu BK, Setchell BP (2004) Effects of paternal heat stress on the in vivo development of preimplantation embryos in the mouse. Reprod Nutr Dev 44: 617–29.

Zhu J, He F, Song S, Wang J, Yu J (2008) How many human genes can be defined as housekeeping with current expression data? BMC Genomics 9: 172.

Zhu JK (2009) Active DNA demethylation mediated by DNA glycosylases. Annu Rev Genet 43: 143-66.

Zhu T, Mo H, Wang N, Nam DS, Cao Y, et al. (1993) Genotypic and phenotypic characterization of HIV-1 patients with primary infection. Science 261: 1179–81.

Zhu W, Pao GM, Satoh A, Cummings G, Monaghan JR, et al. (2012a) Activation of germline-specific genes is required for limb regeneration in the Mexican axolotl. Dev Biol 370: 42–51.

Zhu W, Kuo D, Nathanson J, Satoh A, Pao GM, et al. (2012b) Retrotransposon long interspersed nucleotide element-1 (LINE-1) is activated during salamander limb regeneration. Dev Growth Differ 54: 673–85.

Zhu XP, Dunn JM, Phillips RA, Goddard AD, Paton KE, et al. (1989) Preferential germline mutation of the paternal allele in retinoblastoma. Nature 340:312-3.

Zhu X, Li M, Pan H, Bao X, Zhang J, Wu X (2010) Analysis of the parental origin of de novo MECP2 mutations and X chromosome inactivation in 24 sporadic patients with Rett syndrome in China. J Child Neurol 25: 842-8.

Zhuchenko AA, Korol AB (1983) Ecological aspects of the recombination problem. Theor Appl Genet 64: 177–85.

Zhuchenko AA, Korol AB, Gavrilenko TA, Kibenko TY (1986) The correlation between the stability of the genotype and the change in its recombination characteristics under temperature influences. Genetika 22: 966–74.

Zhuchenko AA, Korol AB, Vizir IY, Bocharnikova NI, Zamorzaeva NI (1989) Sex differences in crossover frequency for tomato and thale cress (Arabidopsis thaliana). Soy Genet 24: 1104-10.

Ziech D, Franco R, Pappa A, Panayiotidis MI (2011) Reactive oxygen species (ROS)–induced genetic and epigenetic alterations in human carcinogenesis. Mutat Res 711: 167-73.

Ziegelhoffer EC, Donohue TJ (2009) Bacterial responses to photo-oxidative stress. Nat Rev Microbiol 7: 856–63.

Ziegler-Skylakakis K, Andrae U (1987) Mutagenicity of hydrogen peroxide in V79 Chinese hamster cells. Mutat Res 192: 65-7.

Ziel KA, Campbell CC, Wilson GL, Gillespie MN (2004) Ref-1/Ape is critical for formation of the hypoxia-inducible transcriptional complex on the hypoxic response element of the rat pulmonary artery endothelial cell VEGF gene. FASEB J 18: 986–8.

Zikidis KC, Vasilakos AV (1996) ASAFES2: a novel-fuzzy architecture for fuzzy computing, based on functional reasoning. Fuzzy Set Syst 83: 63-84.

Zilberman D, Cao X, Jacobsen SE (2003) ARGONAUTE4 control of locus-specific siRNA accumulation and DNA and histone methylation. Science 299: 716-9.

Zimmermann RC, Xiao E, Husami N, Sauer MV, Lobo R, et al. (2001) Short-term administration of antivascular endothelial growth factor antibody in the late follicular phase delays follicular development in the rhesus monkey. J Clin Endocrinol Metab 86: 768–72.

Zimmermann RC, Xiao E, Bohlen P, Ferin M (2002) Administration of anti-vascular endothelial growth factor receptor 2 antibody in the early follicular phase delays follicular selection and development in the rhesus monkey. Endocrinology 143: 2496–502.

Zimmermann RC, Hartman T, Kavic S, Pauli SA, Bohlen P, et al. (2003) Vascular endothelial growth factor receptor 2-mediated angiogenesis is essential for gonadotropin-dependent follicle development. J Clin Invest 112: 659-69.

Zinaman MJ, Clegg ED, Brown CC, O'Connor J, Selevan SG (1996) Estimates of human fertility and pregnancy loss. Fertil Steril 65: 503–9.

Zingg JM, Shen JC, Yang AS, H. Rapoport H, Jones PA (1996) Methylation inhibitors can increase the rate of cytosine deamination by (cytosine-5)-DNA methyltransferase. Nucleic Acids Res 24: 3267–75.

Zini A, O’Bryan MK, Magid MS, Schlegel PN (1996) Immunohistochemical localization of endothelial nitric oxide synthase in human testis, epididymis, and vas deferens suggests a possible role for nitric oxide in spermatogenesis, sperm maturation, and programmed cell death. Biol Reprod 55: 935–41.

Zini A, San Gabriel M, Baazeem AA (2009) Antioxidants and sperm DNA damage: a clinical perspective. J Assist Reprod Genet 26: 427–32.

Zinkevich NS, Gutterman DD (2011) ROS-induced ROS release in vascular biology: redox-redox signaling. Am J Physiol Heart Circ Physiol 301: H647–H653.

Zinn AR, Ross JL (1998) Turner syndrome and haploinsufficiency. Curr Opin Genet Dev 8: 322–7.

Zinn KE, Tunc-Ozdemir M, Harper JF (2010) Temperature stress and plant sexual reproduction: uncovering the weakest links. J Exp Bot 61: 1959-68.

Zinser ER, Kolter R (2000) Prolonged stationary-phase incubation selects for lrp mutations in Escherichia coli K-12. J Bacteriol 182: 4361–5.

Zirkin BR, Chen H, Luo L (1997) Leydig cell steroidogenesis in aging rats. Exp Gerontol 32: 529-37.

Zirkin BR, Chen H (2000) Regulation of Leydig cell steroidogenic function during aging. Biol Reprod 63:977-81.

Zoccarato F, Cavallini L, Alexandre A (2004) Respiration-dependent removal of exogenous H2O2 in brain mitochondria: inhibition by Ca2+. J Biol Chem 279: 4166–74.

Zotin AI (1990) Thermodynamic bases of biological processes: physiological reactions and interactions. New York, NY: Walter de Gruyter.

Zs.-Nagy I (1992) A proposal for reconsideration of the role of oxygen free radicals in cell differentiation and aging. Ann NY Acad Sci 673: 142–59.

Zuckerkandl E, Pauling L (1965) Evolutionary divergence and convergence in proteins. In: Bryson V, Vogel HJ, eds. Evolving genes and proteins. New York, NY: Academic Press. pp 97–166.

Zuckerman S (1951) The number of oocytes in the mature ovary. Recent Prog Horm Res 6: 63-109.

Zuckerman S (1971) Beyond the ivory tower: the frontiers of public and private science. New York, NY: Taplinger Pub. Co.

Zufall RA, Robinson T, Katz LA (2005) Evolution of developmentally regulated genome rearrangements in eukaryotes. J Exp Zoolog B Mol Dev Evol 304: 448–55.

Zuk M, Thornhill R, Ligon JD, Johnson K (1990) Parasites and mate choice in red jungle fowl. Am Zool 30: 235-44.

Zuk O, Hechter E, Sunyaev SR, Lander ES (2012) The mystery of missing heritability: Genetic interactions create phantom heritability.Proc Natl Acad Sci USA 109:1193-8.

Source(s) of Funding


None

 

Competing Interests


None

 

Disclaimer


This article has been downloaded from WebmedCentral. With our unique author driven post publication peer review, contents posted on this web portal do not undergo any prepublication peer or editorial review. It is completely the responsibility of the authors to ensure not only scientific and ethical standards of the manuscript but also its grammatical accuracy. Authors must ensure that they obtain all the necessary permissions before submitting any information that requires obtaining a consent or approval from a third party. Authors should also ensure not to submit any information which they do not have the copyright of or of which they have transferred the copyrights to a third party.
Contents on WebmedCentral are purely for biomedical researchers and scientists. They are not meant to cater to the needs of an individual patient. The web portal or any content(s) therein is neither designed to support, nor replace, the relationship that exists between a patient/site visitor and his/her physician. Your use of the WebmedCentral site and its contents is entirely at your own risk. We do not take any responsibility for any harm that you may suffer or inflict on a third person by following the contents of this website.

Reviews
2 reviews posted so far

Review The mutagenesis-selection-cascade theory of sexual reproduction
Posted by Dr. Dipak Kumar Sahoo on 19 Jun 2016 10:26:30 AM GMT Reviewed by Interested Peers

Revie on The mutagenesis-selection-cascade theory of sexual reproduction
Posted by Prof. Kulvinder K Kaur on 26 Sep 2013 04:39:32 AM GMT Reviewed by Interested Peers
This review will not be counted towards final review score for this article and for its inclusion into WebmedCentral Peer Reviewer articles because reviewer did not feel he/she had sufficient experience and knowledge to review the article.

Comments
0 comments posted so far

Please use this functionality to flag objectionable, inappropriate, inaccurate, and offensive content to WebmedCentral Team and the authors.

 

Author Comments
0 comments posted so far

 

What is article Popularity?

Article popularity is calculated by considering the scores: age of the article
Popularity = (P - 1) / (T + 2)^1.5
Where
P : points is the sum of individual scores, which includes article Views, Downloads, Reviews, Comments and their weightage

Scores   Weightage
Views Points X 1
Download Points X 2
Comment Points X 5
Review Points X 10
Points= sum(Views Points + Download Points + Comment Points + Review Points)
T : time since submission in hours.
P is subtracted by 1 to negate submitter's vote.
Age factor is (time since submission in hours plus two) to the power of 1.5.factor.

How Article Quality Works?

For each article Authors/Readers, Reviewers and WMC Editors can review/rate the articles. These ratings are used to determine Feedback Scores.

In most cases, article receive ratings in the range of 0 to 10. We calculate average of all the ratings and consider it as article quality.

Quality=Average(Authors/Readers Ratings + Reviewers Ratings + WMC Editor Ratings)