Original Articles
 

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

Sociobiology, inclusive fitness, kin selection, stochasticity, cooperation, altruism, mutualism, cheating, self-organization, multilevel selection

Heininger K. Duality of stochasticity and natural selection shape the ecology-driven pattern of social interactions: the fall of Hamilton's rule. WebmedCentral ECOLOGY 2015;6(3):WMC004804

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: 30 Mar 2015 03:39:51 AM GMT
Published on: 30 Mar 2015 08:50:45 AM GMT

Abstract


Both competition and cooperation are pervasive at all levels of biological organization. Traditionally, the theory of evolution, understood as “red in tooth and claw” is challenged by the mere existence of cooperation. As a plausible explanation of this conundrum, Hamilton’s rule has dominated sociobiology for 50 years, but its genetic component (relatedness) is increasingly questioned. On the other hand, there is no doubt that social interactions are regulated by ecological (cost-benefit) factors.

Hamilton’s rule, like Darwin’s theory, implicitly assumes a stable environment. Stable environments favor selfish individuals that are selected to maximize their fitness. Hamilton’s rule allegedly explains the paradox of altruism, that selfish individuals forego their own reproductive opportunities to help close kin to reproduce. Kin selection/inclusive fitness was Hamilton’s explanation for this conundrum. But abiotic and biotic environments are not stable, but variable, often unpredictable. Particularly, biotic environments with their manifold conflicts, Red Queen coevolutionary arms races, and density- and frequency-dependence of fitness have often chaotic dynamics. Environmental stochasticity, resulting in uncertain, unpredictable reproductive success, changes fundamentally the rules for the “gamble of life”. A cybernetic model of evolution revealed the duality of stochasticity and natural selection at input and output levels of the evolutionary Black Box, resulting in multilevel selection of social behavior (Heininger, 2013, 2015). From microbes to mammals, cooperation is selected-for in harsh, uncertain and unpredictable environments. In stochastic environments, cooperation trades individual fitness maximization for less variability and greater reliability of evolutionary outcomes. Thus, the evolution of cooperation is a bet-hedging (risk spreading) strategy of risk-averse individuals. The biological default setting of individuals is neither selfish nor cooperative but ecologically context-dependent and dynamic. Competition and cooperation are threshold traits of nonlinear complex systems on a continuum of ecological variables. The frequent kin structure of communities is not the predominant reason for cooperation/altruism but the result of limited dispersal. Limited dispersal and the evolution of cooperation share environmental stochasticity as common cause resulting in spurious relationships. The environmental conditions that favor cooperativity discourage dispersal and promote philopatry, thus shaping the preferential kinship structure of cooperative communities. Under more adverse ecological conditions, kin competition can strongly antagonize the benefits of kin cooperation and inhibit the evolution of cooperation in viscous populations. Not kinship but context-dependent, pleiotropic processes shape the dynamic sociobiological behavior of populations.

The ecological conditions and genetic “fossil record” of social behavior in both prokaryotic and eukaryotic microbes hold the key to understand the evolutionary present of cooperation in higher taxa. In microbes, cooperative behavior is induced by adverse conditions, particularly starvation, leading to social aggregations with the formation of complex patterns such as fruiting bodies and biofilms. In colonies of metabolically stressed clonal cells the competition for scarce resources is decided by a fair lottery. Experiments with bacteria and social amoeba suggest that cell fate ‘decisions’ (either survival as spores or cell death whose remains fuel the metamorphosis of spores) are stochastic, and moreover that these ‘decisions’ are controlled by genetically-encoded probabilities that are evolvable. Behind the “veil of ignorance”, the competitors are not “aware” of their relative position in the competitive hierarchy. This hierarchy is determined stochastically through a variety of cellular processes with inherent noise that render the cells heterogeneous and the lotteries fair.

The vast majority of cooperative systems are characterized by dominance hierarchies with asymmetric conflicts between dominants and subordinates over limited reproductive opportunities. Eusocial societies, the Holy Grail of kin selection theory, are despotic Orwellian societies that only serve the reproductive needs of selfish Big Sisters and Brothers that for this purpose enslave, police, and suppress their worker castes by aggression and chemical agents. In this respect they resemble a unitary metazoan organism with its reproductive monopoly of the germline. Punishment allows the evolution of cooperation (or anything else) in sizable groups; thus neither altruism nor inclusive fitness gains are behaviorally or causally involved in these despotic systems. Intriguingly, queen pheromones are even able to suppress reproductive activity across species boundaries e.g. from honeybees to fruit flies. The existence of this oppressive system clearly argues against the role of inclusive fitness in the evolution of eusociality. Nestmate recognition, the eusocial version of the immune system and histocompatibility complex, has the function to ensure the reproductive monopoly. Within my alternative conceptual framework of colony fitness a joint genetic-physiological-behavioral- ecological hypothesis of eusociality in insects is presented.

In stochastic environments, reproductive success is unpredictable and highly variable. In taxa without parental brood care, particularly insects, survivorship to reproductive maturity is extremely low. Even in eusocial insects with independent colony foundation, the vast majority of attempts to establish a colony will fail. The extent to which immediate gains are preferred over future rewards is known as future discounting. Individuals who grow up in environments where resources are scarce, competition is intense, and mortality is high should discount the future with its uncertain benefits more heavily than individuals who grow up in abundant, supportive, long-lived habitats. On the other hand, an uncertain, unpredictable environment selects for evolutionary gambling, and either conservative or diversifying bet-hedging as individual- or population-level insurance to individual risk. The deterministic Hamilton’s rule is hardly compatible with the stochasticity of reproductive success in uncertain, unpredictable environments. If the basic assumptions of kin selection/inclusive fitness theory would be right, eusociality should have evolved in (i) less adverse environments with less unpredictable reproductive success, and (ii) in taxa with more predictable reproductive success and, hence, less uncertain inclusive fitness benefits.

The insight that cooperation is a selected-for trait in stochastic environments unifies the so-far distinct concepts of cooperation, mutualism, symbiosis and cooperation in mating systems. Cooperation emerges as a self-organized behavior of complex systems. Phenomena such as swarm-behavior and -intelligence and division of labor emerge from the interplay of both stochastic and deterministic forces, generating order from disorder through self-organization. Responding to the uncertainties of lotteries with insurance policies, populations engage in nest/burrow building, social contracts with assured fitness returns and social queuing. Via the law of large numbers evolution generated a form of automatic biological insurance against idiosyncratic risk.

From a systems biology perspective Hamilton’s rule is simplistic, biased by observation selection, static, and parochial (ignoring the world outside its limited scope, i.e., both the huge majority of cooperative behaviors targeted to nonkin and to other species and being blind to epistemological inputs e.g. from complexity theory and neurobiology). Hamilton’s rule was (and many mathematical models still are) shaped by an egalitarian worldview in which autonomous individuals “decide” to forego their reproduction and help their kin to ensure the representation of their genes in the next generation. In nature, however, the vast majority of cooperative systems are characterized by dominance hierarchies with asymmetric conflicts. The “decision” of subordinates to help is not self-determined but enforced by despotic dominants and the prevailing environmental conditions that limit the subordinates’ options for independent reproduction. In a world of nonlinear biological processes and social interactions, complex emergent behavior, and probabilistic theories, Hamilton’s linear and deterministic rule is a plausibility-based anachronism (one of Gould and Lewontin’s “just-so stories”) rooted in Newtonian thinking. Increasingly, “expanded” definitions of altruism and indirect fitness have been used to rescue the concept. The sociobiological definition of altruism, as one of an outcome, does not say how this outcome has been achieved. In sociobiological parlance, definitions such as fitness transfer by force, “enforced altruism”, as a result of mere luck “coin-flipping altruism”, or the mutual exchange of commodities “reciprocal altruism”, pervert the common sense definition of altruism. In its last consequence, kin selection theory is a fascistoid concept, emphasizing the cohesive value of genetic homogeneity for a population tied together by parochial altruism. Parochial altruism promotes group conflict and may coevolve with warfare. On the other hand, mutualism causes partners to become increasingly dependent on each other as a basis for peaceful coexistence in societies.

Table of contents


Abstract
1. Introduction
2. The shortcomings of observation selection and linear thinking
3. What is altruism?
4. Both competition and cooperation are pervasive
   4.1 Competition
      4.1.1 Cell competition
   4.2 Cooperation
      4.2.1 Cooperation or “altruism”: direct or indirect fitness effects?
         4.2.1.1 Cooperative breeders
      4.2.2 Group size
      4.2.3 Limited dispersal
         4.2.3.1 Law of Causality
         4.2.3.2 Limited dispersal and cooperation: spurious relationship?
5. The social “games” of microbes: cooperation, “altruism”, cheating, or exploitation?
   5.1 Myxococcus xanthus
   5.2 Dictyostelium discoideum
   5.3 Microbial cooperation and competition
      5.3.1 Cooperation
      5.3.2 Competition and siblicide
   5.4 Excursion: “Veil of ignorance” and fair lotteries
      5.4.1 Apoptosis as fair lottery with unequal outcome
      5.4.2 Are “self-destructive” acts the phenotype of fair lotteries?
6. The ecological context-dependency of social interactions
   6.1 The dynamics of cooperation and competition
7. Cooperation and competition: threshold traits on a continuum of ecological variables
   7.1 Conflict as a driver of evolutionary innovation
   7.2 The ecology of “cheaters”
      7.2.1 Frequency-dependence of cheater’s payoff
      7.2.2 Social parasitism as bet-hedging strategy?
8. Eusociality
   8.1 Ecological success
   8.2 Superorganism
   8.3 Kin selection and eusociality
      8.3.1 Haplodiploidy hypothesis of eusociality
         8.3.1.1 Split sex ratios
      8.3.2 Monogamy hypothesis of eusociality
      8.3.3 Low relatedness due to polygyny and polyandry
   8.4 Eusociality is an assimilated/accommodated phenotype
   8.5 A joint genetic-physiological-behavioral-ecological hypothesis of eusociality in insects
      8.5.1 Genetic
      8.5.2 Physiological
      8.5.3 Behavioral
      8.5.4 Ecological
9. Reproductive skew
   9.1 Reproductive suppression in vertebrates
   9.2 Reproductive suppression in eusocial insects
   9.3 Excursion: Pleiotropy
10. Resources and social interactions: a tale of murder, affiliation and indifference
   10.1 Scarce resources and social behavior
   10.2 Unpredictable or limited resources and sociality
   10.3 Abundant resources and sociality
   10.4 Resources and sociality in plants
11. The genetics of social interactions
   11.1 Social behavior, a complex trait
      11.1.1 Aggression
   11.2 Insulin/IGF-like signaling pathways
12. The neurobiology of social interactions
   12.1 Vertebrates
      12.1.1 Oxytocin/vasopressin
      12.1.2 Hypothalamic-pituitary-adrenal axis
   12.2 Invertebrates
13. The ecology of kin recognition
   13.1 Kin recognition by olfactory cues
      13.1.1 Hydrocarbons as olfactory cues in nestmate recognition
      13.1.2 Histocompatibility polymorphism
   13.2 Kin discrimination and nepotism
   13.3 Excursion: An Orwellian Big Brother society needs an effective system of communication and surveillance
14. Stochasticity and selection: duality in evolution
   14.1 Cooperation and multilevel selection
15. Stochasticity and cooperation
   15.1 The unpredictability of reproductive success in stochastic environments
      15.1.1 Discounting future uncertain benefits
   15.2 Cooperation as bet-hedging response to environmental stochasticity
      15.2.1 Empirical evidence
         15.2.1.1 Cooperative breeding
      15.2.2 Evidence from models
   15.3 Natural selection stabilizes cooperation
16. Stochasticity, complex systems and self-organization
   16.1 Stochasticity of swarms
      16.1.1 Swarm behavior
      16.1.2 Swarm intelligence
   16.2 Division of labor due to self-organization
17. Stochasticity, lotteries and insurance
   17.1 Swarm formation as insurance
   17.2 Fortress defense
   17.3 Social queuing as lottery
   17.4 Parental care and assured fitness returns
18. Sexual reproduction as cooperative behavior
19. Prosocial intent and anthropocentric arguments
20. Kin selection and inclusive fitness: the fall of Hamilton’s rule
   20.1 Attempts to save Hamilton’s rule
21. Does altruism exist?
22. Conclusions
23. Abbreviations
24. References

1. Introduction


The challenge to any approach purporting to replace inclusive fitness theory is to explain the same phenomena without using the insights or concepts of the theory.
Bourke (2011a)

Summary
At its heart, evolutionary theory has an economic algorithm, a cost–benefit calculation. The theory of evolution by Natural Selection predicts that individuals will be selfish rather than cooperative. But the world is cooperative and this dilemma has vexed biologists since Darwin. Hypotheses regarding the origin and maintenance of social behavior emphasize the role of ecology and relatedness. While the role of ecology in the evolution of social behavior is not in doubt, recent controversy concerns the role of relatedness.

A growing consensus suggests that ecological/evolutionary and economic theories are ultimately indistinguishable (Boulding, 1978; Hirschliefer, 1977; Real & Caraco, 1986; Noë & Hammerstein, 1994, 1995; Noë et al., 2001; Gandolfi et al., 2002; Hodgson, 2002; Orr, 2007; Yaari & Solomon, 2010; Okasha & Binmore, 2012). 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 (Maderspacher, 2011). Costs represent the flipside of direct benefits (e.g., see Lee & Hays 2004) For a focal genotype of interest to increase in frequency in a population, carriers must, on average, end up with more net direct fitness benefits than average population members. This net direct fitness must account not only for any costs and/or benefits to the focal carrier due to its own behavior associated with the trait, but also for any fitness benefits received from other individuals (related or not; Queller, 1985; Fletcher & Zwick, 2006). In particular, any trait that causes carriers via their own behaviors to put themselves at a disadvantage compared with those they interact with, will only increase if the benefits received from others are sufficient to make up for this disadvantage (Fletcher & Doebeli, 2009).

Robert May began his last presidential address to the Royal Society on 30th November 2005 by saying: “The most important unanswered question in evolutionary biology, and more generally in the social sciences, is how cooperative behaviour evolved and can be maintained in human or other animal groups and societies”. Darwin’s theory of evolution by Natural Selection predicts a nature “red in tooth and claw” (Ruse, 1999). In such a nature individuals will be selfish rather than cooperative (Darwin, 1859; Fisher, 1930). How then, can genetically prescribed selfless behavior arise by natural selection, which is seemingly its antithesis? This problem has vexed biologists since Darwin, who in The Origin of Species declared the paradox—in particular displayed by ants—to be the most important challenge to his theory. The hypotheses regarding the origin and maintenance of social behavior can be categorized into three classes (Slobodchikoff & Shields, 1988): (i) Genetic hypotheses assume that high levels of kinship among group members are necessary and sufficient to explain the kind of society being investigated (Hamilton, 1964; Alexander, 1974; Trivers, 1985). (ii) Ecological hypotheses assume that nongenetic environmental factors are necessary and sufficient to explain an observed social system (Crook, 1965, 1970; Lack, 1968, Emlen, 1982a). (iii) A phylogenetic hypothesis assumes that a social system evolved in the context of conditions occurring earlier in the evolutionary history of the lineage in question and does not represent an adaptation to current conditions (Wilson EO, 1975).

Allegedly, Hamilton's rule (Hamilton, 1963, 1964) provides a tool for understanding a range of social interactions, including altruism, aggression, selfishness and spite. It states that altruism (or less aggression) is favored when rb – c > 0, where c is the fitness cost to the altruist, b is the fitness benefit to the beneficiary and r is their genetic relatedness. The basic idea of the kin selection theory is that of inclusive fitness, a concept that breaks through the limitation imposed by the so-called Darwinian fitness, in which for a characteristic to have adaptive value it must favor its bearer’s ability to reproduce. According to inclusive fitness theory, a gene which codes for behavior that is costly to the individual who carries it, but benefits his genetic relatives, e.g. sharing food with siblings, will increase in frequency by natural selection—because the individual’s relatives are likely to carry copies of the gene in question themselves. Altruism can evolve, Hamilton concluded, so long as the cost incurred by the altruist is offset by a sufficient amount of benefit to sufficiently closely related relatives. Kin selection theory is widely accepted among evolutionary biologists as the most plausible way of explaining the evolution of altruism from a Darwinian perspective (Cronin, 1991; Okasha, 2002). In the kin selection equation (Hamilton, 1964) relatedness is deeply intertwined with ecology (cost-benefit) so that both are essential (Crozier, 2008). While the role of ecology in the evolution of social behavior is not in doubt, recent controversy concerns the role of relatedness. Nowak et al. (2010) argued that the narrow focus on relatedness often fails to characterize the underlying biology and prevents the development of multiple competing hypotheses. This work has received some support (Doebeli, 2010; Gadagkar, 2010; van Veelen et al., 2010a) but, not surprisingly, has been angrily rebuffed by more than 150 prominent advocates of inclusive fitness theory (Abbot et al., 2011; Boomsma et al., 2011; Bourke, 2011a; Ferriere & Michod, 2011; Gardner et al., 2011; Herre & Wcislo, 2011; Marshall, 2011; Ratnieks et al., 2011; Rousset & Lion, 2011; Strassmann et al., 2011a). Edward O. Wilson (in response to R. Dawkins’ recension of “The Social Conquest of Earth” [Prospect Magazine 195, 24th May 2012]) answered that “It should be born in mind that if science depended on rhetoric and polls, we would still be burning objects with phlogiston and navigating with geocentric maps.”

In the following, I will argue that, from a systems biology perspective, the kin selection/inclusive fitness theory is simplistic, biased by observation selection, static and parochial. The result of my comprehensive literature study is an ecological-evolutionary scenario in which the omnipresent dualism of environmental stochasticity and natural selection shaped the evolutionary pattern of cooperation and competition. The complexity of the sociobiological regulation 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 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. The shortcomings of observation selection and linear thinking


It is theory which decides what we can observe
Einstein to Heisenberg, 1926 (Heisenberg, 1971)

Mathematical descriptions of nature are not fundamental truths about the world, but models. There are good models and bad models and indifferent models, and what model you use depends on the purposes for which you use it and the range of phenomena which you want to understand […] reductionist rhetoric ... claims a degree of correspondence between deep underlying rules and reality that is never justified by any actual calculation or experiment.
Cohen & Stewart, 1994, p. 410

Summary
The reductionist approach to science today remains largely the dominant model. Reductionist and linear thinking have brought tremendous progress in our knowledge of biological processes and their networks. But reductionist thinking fails in the quest to understand complex systems. The observation selection related to the reductionist approach fails to appreciate biological phenomena in their wider context.

The nineteenth-century American poet John Saxe told the story of the blind men from Industan and the elephant (1892):

It was six men from Industan, to learning much inclined,
Who went to see the elephant (though all of them were blind),
That each by observation might satisfy his mind…

Not surprisingly, each blind man felt a different part of the elephant, so that the one touching its legs thought they were tree trunks, the one touching its tail thought it was a snake, and so on (Gandolfi et al., 2002). By the end, they

Disputed loud and long, each in his opinion stiff and strong,
Though each was partly in the right, and all of them were wrong.

The same issue was approached by Bostrom (2003) from a different perspective. He asked: “How big is the smallest fish in the pond? You take your wide-meshed fishing net and catch one hundred fishes, every one of which is greater than six inches long. Does this evidence support the hypothesis that no fish in the pond is much less than six inches long? Not if your wide-meshed net can’t actually catch smaller fish.
The limitations of your data collection process affect the inferences you can draw from the data. In the case of the fish-size-estimation problem, a selection effect—the net’s being able to sample only the big fish—invalidates any attempt to extrapolate from the catch to the population remaining in the water. Had your net had a finer mesh, allowing it to sample randomly from all the fish, then finding a hundred fishes all greater than a foot long would have been good evidence that few if any fish remaining were much smaller.
In the fish net example, a selection effect is introduced by the fact that the instrument you used to collect data sampled from only a subset of the target population. Analogously, there are selection effects that arise not from the limitations of the measuring device but from the fact that all observations require the existence of an appropriately positioned observer. These are known as observation selection effects.”

Traditional linear thinking approaches work against an understanding of how the different parts of a system work together and underplay or ignore the multifaceted nature of complex problems. Gould and Lewontin (1979) criticized the adaptationist “just-so stories”, that break an organism into unitary ‘traits’, either behavioral or physical, and propose an adaptive story for each trait considered separately. Instead of speculations about how a trait might possibly have been selected they advocated a holistic approach, to analyze organisms as integrated wholes. Naturally, with any research program that requires theory to be integrated with data, there is an inevitable tension between experimental biologists, who deal daily with the complexity of real biological systems, and theoretical biologists, who “simplify, simplify, simplify” in the name of tractability (Rouzine et al., 2001). All models, by definition, are simplified representations of natural systems or processes and out of necessity, must be abstractions from reality. It has been stressed (Kingsland, 1985; Clark & Yoshimura, 1993) that only the naïve believe that any particular mathematical model can provide more than a partial view of the complexity inherent in any biological system. Thus, all models are wrong in some respect by design (or more commonly, by accident) (Bentley, 2009). Theoreticians noted that when only a few components of a complex system are analyzed in isolation, conceptual mistakes commonly arise (e.g.Frank, 1996; Melbourne & Hasting, 2008; Stover et al., 2012). Byrne & Callaghan (2014) even noted that the development of systems of equations, however sophisticated, may well have very little to tell us about the social world. In the same vein of thought, Wilbur et al. (1974) remarked that attempts to explain life histories as outcomes of single selective pressures, however simple and appealing, have obscured rather than elucidated the evolution of life histories. Most studies of the genetics of adaptation rely on a separation of timescales of evolution and ecology (Orr, 2010). In many cases, this might be a severe oversimplification (e.g. Hairston et al., 2005; Heininger, 2013). Indeed, recent developments on the establishment of beneficial mutations in changing environments reveal that this separation can lead to qualitatively different results (Uecker & Hermisson, 2011).

Hamilton’s rule is an extreme example of linear thinking. The early models of social behavior assumed that altruistic and selfish behaviors are caused directly by corresponding genes, which means that the only way for groups to vary behaviorally is for them to vary genetically. Although this was assumed in the models primarily to simplify the mathematics, hardly anyone regards such strict genetic determinism as biologically realistic (Wilson & Wilson, 2007). Van Veelen et al. (2010b) concluded that the tendency to form groups and the tendency to cooperate interact; either looking at the two traits in isolation or allowing them to evolve together can give predictions that differ from each other (van Veelen et al., 2010b). In a similar vein of thought, Mitchell (2009, p. 114) argued for an “integrative pluralism which attempts to do justice to the multilevel, multicomponent, evolved character of complex systems”. Models of social queuing (see chapter 17.2) that take into account even the smallest individual quality differences, which are probably plentiful in nature, result in individuals using queuing strategies completely different from those in models that assume qualitatively equal competitors (van de Pol et al., 2007). Generally it can be stated that even the most extensive models do not capture the complexity that can be demonstrated in real systems (Clark & Yoshimura, 1993; Sheldon, 2002).

The reductionist approach to science today remains largely the dominant model. It fosters the detailed study of limited domains in individual subdisciplines within the vast tree of science. The decline of Partula snails on a Pacific island, the flow of viscoelastic fluids within porous media, antibiotic resistance in the dihydropteroate synthase enzyme, and the search for the Higgs boson are examples of the degree of specialization in research. That the traditional epistemological approach of reductionism is not sufficient to explain complexity and self-organization is a view increasingly shared among scientists (Coveney, 2003). Systems thinking, on the other hand, offers an integrative way of appreciating all the major dimensions of a complex problem (Bosch et al., 2013). Theoretically, systems biology adopts a holistic, integrative view (Institute of Systems Biology website: www.systemsbiology.org). In practice, however, it is usually large-scale reductionism: it takes a bottom-up approach, using information from complex interactions among lower level (molecular) processes to explain the functioning of cells and organisms. This is crucially different from life history evolution, which uses a top-down approach, starting at the level where selection acts (typically the individual organism), and moving down the organisational hierarchy to eventually find a mechanistic explanation. Both approaches have their advantages and disadvantages and should be used in a complementary manner (Pijpe, 2007).

In this work, to remain in the metaphor of Saxe, I do not argue that the “snake” has hairs and cannot be a snake. Instead, I embark to retrieve as many data as available to identify the huge, context-dependent, ecological scenario into which cooperation is embedded, finally trying to discern the silhouette and body of the “elephant”.

3. What is altruism?


Summary
The sociobiological definition of altruism, characterized by the costly transfer of fitness, is, in its actual version, the one of an outcome. It does not specify how this outcome has been achieved. This led to the strange situation that in sociobiological parlance definitions are used such as fitness transfer by force, “enforced altruism”; as a result of mere luck “coin-flipping altruism”; or the mutual exchange of commodities “reciprocal altruism”, that pervert the common sense definition of altruism as is e.g. expressed by the Oxford Dictionary definition. As in sociobiology “altruism” is an anthropocentric, intentional term replete with moral implications and teleological connotations, I will use it in quotation marks when I have to use it to refer to the argumentation of other authors but apart from that will avoid the term whenever possible.

The Oxford Dictionary defines altruism as: “disinterested and selfless concern for the well-being of others”. A common definition in sociobiology describes altruism as behavior that simultaneously entails fitness costs to the behaving individual and fitness benefits to individuals on the receiving end of the behavior (Bourke & Franks, 1995; Sober & Wilson, 1998; Kerr et al., 2004). The term “altruism” has been used in different ways by different authors, generating considerable confusion and ambiguity (Grafen, 1984; Nunney, 1985; Wilson, 1990; Uyenoyama & Feldman, 1992; Wilson & Dugatkin, 1992). Moreover, the behavior labeled ‘‘cooperation’’ by evolutionary game theorists is the same as that discussed in the evolution of altruism literature (Sober & Wilson, 1998; Worden & Levin, 2007). A fundamental difficulty with the evaluation of acts of commodity/service transfer is that the outside observer can only observe the outcome of fitness transfer but has no knowledge of the specific motives and mechanisms. Basically, a net transfer of commodities can take place due to voluntary intent, cheating, coercion, or be the result of a lottery. For example, altruistic acts involve the actor voluntarily donating fitness to beneficiaries. Parasitic acts, on the other hand, involve the actor extracting benefit from others at net cost to the donors (Doncaster et al., 2013). Both behaviors may have the same direct net-cost transferral of fitness from donor to beneficiary; the key difference between parasitism and altruism is thus who drives the interaction. Identifying the evolutionary driver is not always straightforward in practice, yet it is crucial in determining the conditions necessary to sustain such fitness exchange (Doncaster et al., 2013).

Social interactions are often represented by a quadrant system (e.g. West et al., 2007c), the quadrants being defined by the positive and negative effects on both actor and recipient. Hamilton's rule stands and falls with the role of kinship in the manifestation of altruism; two of the quadrants representing mutualism and selfishness clearly do not require kinship for their manifestation. And spite is an unproven theoretical concept (Hamilton, 1970) that allegedly has some empirical support (Blackman, 2004); but there are thousands of counterexamples in which, under the appropriate ecological conditions, aggression and killing are either directed indiscriminately against kin and nonkin (see chapters 5.3.2 and 10.1) or even preferentially against kin (Dunn et al., 2014). One of the major flaws of the kin selection theory is that Hamilton classified fitness transfers as act of altruism without intimate knowledge of the underlying mechanism(s). One of the objectives of this work is to correctly classify the mechanisms behind the fitness transfer/exchange and delimit cooperation from altruism. Altruism is an intentional term (Grafen, 1999) replete with moral implications and teleological connotations. Cooperation is less tainted by anthropocentric notions. Therefore, I will put “altruism” in quotation marks whenever I have to use it to refer to the argumentation of other authors but apart from that avoid the term whenever possible.

4. Both competition and cooperation are pervasive


Natural selection is demanding, exacting, relentless. It is intolerant of weakness, indifferent to suffering … One might expect organisms shaped by such a force to bear its stamp… Natural selection would surely see off chivalrous self-sacrifice, selfishness should win the day.
Cronin, 1991

Summary
The competitive advantage of organisms depends on their stress resilience and acquisition of limited and contested resources. As a rule, competition in nature is asymmetrical, establishing dominance hierarchies. 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.
Cooperation including mutualism, and symbiosis is abundant at all levels of biological organization. The benefits derived from cooperations include increased feeding success, access to resources that are unavailable to solitary individuals, increased predator protection, ability to outcompete conspecifics, or ability to escape harsh environmental conditions. Cooperative breeders exert costly parenting efforts to contribute to the developmental success of an infant or juvenile that is not their own offspring. Two major selective pressures, predation risk and resource availability in both space and time influence group size.In most species, the costs and benefits associated with group living vary systematically with group size.
Hamilton argued that limited dispersal would lead to what he called population viscosity, elevating local relatedness sufficiently to allow altruism towards neighbors in general.Habitat heterogeneity, availability and distribution induce a cost for dispersing individuals, such as substantial increases in mortality and the risk to end up in unsuitable habitats, and thus select against dispersers.Likewise, adverse, uncertain, and unpredictable environments favor cooperative behavior (see chapters 6 and 15). The environmental conditions that favor cooperativity also discourage dispersal and promote philopatry, thus shaping the preferential kinship structure of cooperative communities. Thus, environmental heterogeneity and stochasticity may be a shared causal factor for limited dispersal and cooperation resulting in a spurious relationship.

Biologists operating within the paradigm of evolution by natural selection have understandably found it easier to think about predation, competition and parasitism than cooperation and mutualism. A naive evolutionist would not expect individuals, which are not closely related to be ‘nice’ to each other (Wilkinson & Sherratt, 2001).Individuals give for two reasons. One is to get a benefit back. The other is to avoid a cost. “Cooperation” theories stress mutual benefits.“Conflict” theories stress costs (Betzig, 2004). 

4.1 Competition

In most natural populations, the reproductive potential far exceeds the environmental opportunity, and natural selection proceeds by culling to what the habitat can support (King, 1967). As Smith (2011) put it: “In some respects natural selection is a quite simple theory, arrived at through the logical integration of three propositions (the presence of variation within natural populations, an absolutely limited resources base, and procreation capacities exceeding mere replacement numbers) whose individual truths can hardly be denied.” In fact, that populations outgrow resources is the central idea of Malthus’s An Essay on the Principle of Population (1798), that led Darwin to the conclusion that this pressure, analogous to breeder’s artificial selection, was a natural form of selection (Ruse, 2009). According to MacArthur and Wilson (1967, p. 149): “Evolution … favours efficiency of conversion of food into offspring”. How resources affect individual survival and reproductive success can be described by the fitness function, w(x), whose value is the expected number of offspring born to individuals with x units of resource (Rogers, 1992). In a similar ecological context, an energetic definition of fitness was put forward. According to the formulation 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 et al., 2006).

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”. 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. On all levels of biological organization, competition for limited resources is a source of conflict and arguably the most pervasive motor of evolution (Darwin, 1859; MacArthur & Levins, 1964; MacArthur, 1970; Pianka, 1974a; Lawlor & Maynard Smith, 1976; Brown, 1981; Davies, 1982; Niemitz, 2002; Winther, 2005; Kopp & Hermisson, 2006; MacLean & Gudelj, 2006; Fisher & Hoekstra, 2010; Heininger, 2012). As a rule, competition in nature is asymmetrical (Brooks & Dodson, 1965; Lawton, 1981; Lawton & Hassell, 1981; Connell, 1983; Schoener, 1983; Karban, 1986; Weiner, 1990; Callaway & Walker, 1997; Law et al., 1997; Ferriere et al., 2002), establishing dominance hierarchies (Weiner, 1990; Erlandsson, 1992; Shelley et al., 2004; Rychlik & Zwolak, 2006). Competitive asymmetry, which leads to increased individual variability in size, has been seen as one of the major processes that secure the existence of reproductive individuals, stabilize population dynamics and assure the persistence of populations (Aikio & Pakkasmaa, 2003). Natural selection generated by asymmetric competition is likely to be a persistent and continuing phenomenon in communities.

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 & Szathmáry, 1995; Szathmáry & Maynard Smith, 1995; Frank, 1995, 2003, 2009; Ratnieks et al., 2006; Gardner & 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 & Buckling, 2005, 2011), and cheat on the others (Vulic & Kolter, 2001), a phenomenon that also may underlie cancerogenesis (Heininger, 2001; Burt & Trivers, 2006). However, competition is also a strong coevolutionary force resulting in the selection of fitter individuals (Lawlor & Maynard Smith, 1976; Futuyma & Slatkin, 1983; Stephens & Krebs, 1986; Zambrano et al., 1993; Grover, 1997; Svanbäck & Bolnick, 2005; Araújo et al., 2011; Heininger, 2012). 

4.1.1 Cell competition

The possibility that cells of multicellular organisms may also compete with one another has been postulated several times over the past two centuries (Díaz & 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, 2003). 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 & Koufopanou, 1991; Maynard Smith & Szathmáry, 1995; Grosberg & Strathmann, 1998, 2007; Ostrowski & Shaulsky, 2009). Thus, an essential constituent of the cellular cooperation paradigm was the alleged clonality or high relatedness of cells (Queller, 2000; Grosberg & Strathmann, 2007; Fisher et al., 2013). However, there is now evidence for a substantial amount of genetic diversity and genetic mosaicism both between somatic and germline cells (Gill et al., 1995Muotri et al., 2005; Flores et al., 2007; Lam & Jeffreys, 2007; Bruder et al., 2008; Liang et al., 2008; Piotrowski et al., 2008; Coufal et al., 2009; Frank, 2010; Quinlan & Hall, 2012; Li et al., 2013). Cell competition was discovered in the imaginal discs of D. melanogaster 40 years ago (Morata & 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 & Ripoll, 1975; Simpson, 1979; Simpson & Morata, 1981; Moreno et al., 2002; Lolo et al., 2012), despite the fact that they would have been viable on their own (Morata & Ripoll, 1975; Simpson, 1979; Simpson & Morata, 1981; Moreno et al., 2002; de la Cova et al., 2004; Li & 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 & Krasnow, 2006) and in postmitotic epithelial tissue repair (Tamori & Deng, 2013). It seems unlikely that such an effective mechanism to select for cell fitness should be confined to flies (Díaz & 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 & 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). Cellular selection is the ultimate consequence when repair systems fail or are overwhelmed. 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 & Basler, 2004). Consistent with the role of these signaling pathways, competition intensity increases in high-intensity competitive environments (Chesson & Huntly, 1997; Violle et al., 2010; Miller et al., 2011). In all taxa, 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 & Moreno, 2005; Khare & Shaulsky, 2006; Morata & Martin, 2007; Johnston, 2009; Green, 2010; Baker, 2011; de Beco et al., 2012; Vivarelli et al., 2012).

Lynn 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, establishing a feedback cycle. Variation can be caused both by a variety of molecular biological processes (see Heininger, 2013, 2015). 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. 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).

Quality control operates by selecting for performance in cellular functions and eliminating inferior units. 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 & Cuervo, 2011), the ubiquitin/proteasome pathway (Sutovsky et al., 2001; 2002; Kostova & Wolf, 2003; Sutovsky, 2003; Thompson et al., 2003; Kwon et al., 2005; Taylor & Rutter, 2011), autophagy (Jin & White, 2007; Yorimitsu & Klionsky, 2007; Lee JY et al., 2010; Lee & Yao, 2010; Arias & Cuervo, 2011; Murrow & Debnath, 2013), mitochondrial turnover (Tatsuta & Langer, 2008; Twig et al., 2008; Dagda & Chu, 2009; Luce et al., 2010), the endoplasmic reticulum (Ellgaard & Helenius, 2003; Jørgensen et al., 2003; Kostova & Wolf, 2003; Kleizen & Braakman, 2004; Groenendyk & Michalak, 2005; Buchberger et al., 2010), and apoptosis (Yin et al., 1998, 2002; Meier et al., 2000; Groenendyk & Michalak, 2005; Jin & White, 2007; Igaki, 2009). Cellular selection within multicellular organisms does not only occur within the immune system and between cancer cells (Nowell, 1976; Kisielow & von Boehmer, 1995; McLean et al., 1997; Breivik & Gaudernack, 1999; von Boehmer et al., 2003; Vineis, 2003; Frank & Nowak, 2004; Merlo et al., 2006; Moreno, 2008; Vermeulen et al., 2008; Kim et al., 2011; Tamori & Deng, 2011; Thomas et al., 2013) but, as predicted by Roux (1881), is ubiquitous in plants and animals, particularly during development (Purves, 1980; Whitham & Slobodchikoff, 1981; Buss, 1983; Antolin & Strobeck, 1985; Kupiec, 1986, 1996, 1997; Sutherland & Watkinson, 1986; Edelman, 1987; Michaelson, 1987, 1993; Klekowski, 1988, 2003; Gill et al., 1995; Otto & Orive, 1995; Møller & Pagel, 1998; Otto & Hastings, 1998; Deppmann et al., 2008; Clarke, 2011; Tamori & Deng, 2011; de Beco et al., 2012). Like in the ecological context, the cellular selection regime is driven by competition for limited resources (McLean et al., 1997; De Boer et al., 2001; Gaudin et al., 2004; Davies et al., 2012; Wright & Bourke, 2013) such as trophic factors (Harris et al., 1997; van Ooyen & Willshaw, 1999; van Ooyen, 2001) and/or some other scarce resources that are needed to promote metabolism (Montague, 1996; Thomaidou et al., 1997).

Darwin imagined, in his last paragraph of the Origin of Species, a tangled bank of competing organisms, and there is ample evidence 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, 2002, 2012; Weiss, 2006). Within-plant competition has been demonstrated in pine trees (Honkanen & Haukioja, 1994) and in annual pea plants (Sachs & Novoplansky, 1997). It has been proposed that clones and individual plants are formed by iterated, semi-autonomous modules (e.g. ramets or shoot modules) that may respond independently to local conditions. The branch-competition hypothesis (Sachs & Novoplansky, 1997) specifically predicts that an inferior module is left out of support if more viable sinks are available (de Kroon et al., 2005). 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 based on cell competition is the ‘selective stabilization of synapses’ (Changeux et al., 1973; Changeux & Danchin, 1976), later confirmed and even explicitly named ‘neural Darwinism’ (Edelman & Mountcastle, 1978; Edelman, 1987). 

4.2 Cooperation

Cooperation is abundant at all levels of biological organization (Hammerstein, 2003; Nowak & Highfield, 2011; Weiss et al., 2011). “Cooperation is so widespread, so much widespread, that it is puzzling why scientists were not willing to easily acknowledge its ubiquitousness and importance” (Miramontes & DeSouza, 2014). Cooperation and social phenomena are present in humans, in primates and in social insects–the common examples usually given– but it is also present in unexpected places such as in plants (Callaway & Walker, 1997; Biernaskie, 2011) or bacteria (Griffin et al., 2004; Cordero et al., 2012) or even as emergent phenomena in artificial societies of robots or other creatures of the cyberspace (Maris & te Boekhorst, 1996; Langton, 1997). In some species, social interactions are confined to interactions between the sexes during mating. Species at the other extreme have complex societies in which individuals live in intimate association with nestmates, and social interactions are fundamental to all aspects of life. Interactions between organisms exist along a continuous gradation, and the lines between cooperation, mutualism, commensalism, and parasitism are not neatly delineated and rather vague (Starr, 1975; Lewis, 1985; Ewald, 1987; Hochberg et al., 2000; Thompson & Cunningham, 2002; Neuhauser & Fargione, 2004; McCreadie et al., 2005; Leung & Poulin, 2008; Pérez-Brocal et al., 2013). Facilitation has been defined as an interaction in which the presence of one species alters the environment in a way that enhances growth, survival or reproduction of a second, neighboring species. According to some definitions, facilitation can be mutualistic, antagonistic or commensal (Bronstein, 2009). These interactions can be considered as a continuous spectrum of dependence, ranging from beneficial to detrimental outcomes (Ewald, 1987; Pérez-Brocal et al., 2013).

Even linkages among mutualism, predation and competition in natural systems have been recognized (Crowley & Cox, 2011; Assaneo et al., 2013; Afkhami et al., 2014). “Cooperation” includes any behaviors (joint resource acquisition, information exchange, communal brood care, predator defense, etc.) that, despite any individual costs, have a net beneficial effect on group members (Mesterton-Gibbons & Dugatkin, 1992; Avilés et al., 2002; Khamis et al., 2006; Garay, 2009; Krams et al., 2010). Noë (2006) draws a definitive line between mere sociality, the tendency of conspecifics to aggregate and pursue their own interests in the context of a group, and cooperation, which requires an interaction or series of interactions between or among individuals that carries a cost for the agent but which, on average, results in a net gain for all participants of the interaction. Noë intends this reading of cooperation to include ‘all other terms that have been used for mutually rewarding interactions and relationships: reciprocity, reciprocal altruism, mutualism, symbiosis, collective action and so forth’. This definition will be adopted here. “Altruism”, by contrast, characterizes action by one individual that benefits another (or group) at the cost of the agent’s own survival, wellbeing or reproduction.

Cooperation is not limited to kin and nonkin members of the actors’ species but occurs also between members of different species. Mutualism, defined as an interaction in which two or more species benefit each other, ranges from specific obligate associations to facultative interactions among free-living species (Boucher et al., 1982; Boza & Scheuring, 2004, Holland & Bronstein, 2008). Both mutualism (Boucher et al., 1982; Boucher, 1985; Doebeli & Knowlton, 1998; Frank, 1998; Herre et al., 1999; Yu, 2001; Clutton-Brock, 2002; Bronstein et al., 2006; Leigh, 2010) and long-term mutualisms, i.e. symbioses, (Kropotkin, 1902; Margulis, 1970, 1981, 1998; Harley & Smith, 1983; Boucher, 1985; Smith & Douglas, 1987; Bronstein, 1994, 2009; Klepzig et al., 2001; Ryan, 2002; Kooijman et al., 2003; Rand et al., 2004; Blüthgen et al., 2007; Douglas, 2010; Cain et al., 2011; Mittelbach, 2012) are pervasive in nature. Douglas H. Boucher (1985), in an edited volume on mutualism, pointed out that there is a long-standing debate among ecologists over the relative importance of competition and co-operation in nature, which can be traced back at least to the 1920s. He noted the remarkable fact that, despite a general bias over the years in favor of competition as the basic organizing principle of nature and a concomitant preference among theoretical ecologists for using the famed Lotka-Volterra competition model in their analyses, in fact a cooperative version of the model (involving a simple sign change) has been reinvented (evidently independently) at least 29 times since 1935 (Corning, 1997).

Mutualisms are known in all kingdoms of organisms, and there is a tendency for the partners to come from different kingdoms (Briand & Yodzis, 1982). This is particularly true for obligate and symbiotic mutualisms, and may simply be a reflection of nutritional complementarity. Some taxa seem particularly likely to enter into mutualisms--e.g. NostocTrebouxiaSymbiodinium, and Chlorella (Margulis, 1981), and at a higher level, ants, coelenterates, and legumes (Boucher et al., 1982). Some 319 species of hummingbirds live almost entirely on nectar, a dependence that has led to specialized joints in their wings that enable them to hover with pinpoint accuracy over the appropriate flower. Flower and bird are partners in a mutualistic exchange of food for the bird and assistance with fertilization for the flower. In an even tighter partnership, at least one quarter of all described fungi, an estimated 12,000 to 20,000 species, enter into associations with 40 genera of photosynthetic algae to form some 13,500 species of lichens. This strategy is so successful that lichens are the most enduring and widespread of life forms (Ryan, 2006). Virtually all species of ruminants, including some 2,000 termites, 10,000 wood-boring beetles and 200 Artiodactyla (deer, camels, antelope, etc.) are vitally dependent upon the services provided by endosymbiotic bacteria or fungi for the breakdown of the cellulose in plants into usable cellulose hydrolysis products (Price, 1991; Lynd et al., 2002). After >100 years of research it is reasonable to conclude that most, if not all, multicellular life on earth is symbiotic with microorganisms (Rodriguez & Redman, 2008). Host-associated microorganisms contribute enormously to the development of their host’s immune system, nutrition, digestion, reproduction, and general wellbeing (Currie, 2001; Vance, 2001; McFall-Ngai, 2002; Dillon & Dillon, 2004; Mazmanian et al., 2005; Taylor et al., 2005; Ley et al., 2006a; Martin et al., 2008; Pais et al., 2008; Chaston & Goodrich-Blair, 2010; Kinross et al., 2011; Mattila et al., 2012). It has been argued that access to mutualistic endosymbiotic microbes is an underappreciated benefit of group living (Lombardo, 2008). A species that is relatively efficient at acquiring one resource would benefit from specialization on acquisition of that resource accompanied by trade for the other resource. The theory of relative advantage extends this prediction to show that specialization and trade confer an advantage even for species that are relatively poor resource competitors for both resources (Schwartz & Hoeksema, 1998). The formation of cooperative and mutualistic partnerships has so much in common with the formation of reproductive partnerships that the formulation of a single basic model seems warranted (Noë & Hammerstein, 1994). In all forms of reproductive and mutualistic pair formation, and in many cases of intra-specific cooperation, the partners belong to two distinct classes, e.g. males and females, figs and fig wasps, or breeders and helpers (Noë & Hammerstein, 1994). Because mutualism causes partners to become increasingly dependent on each other, it is a basis for peaceful coexistence in societies (Clutton-Brock, 2002). In promoting peaceful coexistence, mutualism would be antagonized by kin selection (Zahavi, 1995; Clutton-Brock, 2002).

Most attention on the phenomenon of cooperation has been focused on interactions between animals. However, the same phenomenon occurs at all levels of biological organisation (Hamilton, 1972; Leigh, 1991; Maynard Smith & Szathmary, 1995; Lehmann & Keller, 2006; West et al., 2007a). Cooperative assembly of unicellular organisms into multicellular aggregates, mounds and fruiting bodies may have occurred many times in evolution and is common in nature (Bonner, 2000). The structural complexity and degree of organization of microbial multicellular structures vary from a simple single-layer biofilm and simple aggregates to complicated structures like the fruiting bodies of myxobacteria and slime molds, complex natural biofilms (see chapter 5.3.1) and the colonies of various microbes (Palková & Váchová, 2006; Annesley & Fisher, 2009; Velicer & Vos, 2009; Cáp et al., 2012; Elias & Banin, 2012).

Multicellularity has arisen multiple times (Bonner, 1998; Cavalier-Smith, 1998; Baldauf, 2003; Adl et al., 2005; Keeling et al., 2005; Parfrey et al., 2010). A new individual entity emerges from the interaction of previously independent components. These transitions include the evolution of chromosomes, eukaryotes, sexual reproduction, multicellular organisms, and social insects (Maynard Smith & Szathmáry, 1995; Bourke, 2011b). Separate genes, which make up the genome, cooperate in what has been termed the ‘parliament of the genes’ (Leigh, 1971; Corning, 1996). Molecular interactions are organized in cooperative molecular networks (Weiss & Buchanan, 2009; Barabasi et al., 2011; Foster, 2011). The very existence of multicellular organisms relies upon cooperation between the eukaryotic cells that make them up. The mitochondria upon which these eukaryotic cells rely were once free-living prokaryotic cells but now live cooperative lives. The tree of life is dominated by single-celled microorganisms that appear to perform a huge range of cooperative behaviors (West et al., 2006). For example, the growth and survival of bacteria depend upon excreted products that perform a variety of functions, such as scavenging nutrients, communication, defence and movement. The benefits of such extracellular products can be shared by neighboring cells and hence they represent a ‘public good’ that is open to the problem of exploitation (West et al., 2006). Almost all of the major evolutionary transitions from replicating molecules to complex animal societies have relied upon solving the problem of cooperation (Maynard Smith & Szathmáry, 1995).

In animals, selection for sociality may increase the fitness of the individual as a function of group cooperation (Whitehouse & Lubin, 2005). The benefits derived from cooperations include increased feeding success, access to resources that are unavailable to solitary individuals (Slobodchikoff, 1984; Raffa & Berryman, 1987; Dugatkin, 1997; Wyatt, 2003; Whitehouse & Lubin, 2005; Bonsall & Wright, 2012; Platt et al., 2012), increased predator protection, ability to outcompete conspecifics, or ability to escape harsh environmental conditions. Because of the efficiency of cooperation, the total productivity attained by the group tends to be larger than the sum of individual contributions (Toyoizumi, 2009). Information exchange and group foraging, for instance, are thought to allow the exploitation of patchy and ephemeral resources in colonial breeding birds (e.g., Brown et al., 1990; Berg et al., 1992; Wiklund & Andersson, 1994; Brown & Brown, 1996; Rolland et al., 1998). Female bark beetles (Dendroctonus montanus) use aggregation pheromones to call in conspecifics to overcome the defenses of live trees and to gain access to a resource that is not available to solitary beetles (Raffa & Berryman, 1987; Wyatt, 2003). The utilization of live trees by the tree-killing bark beetles might have followed the development of better communication systems (pheromones) to attract conspecifics during group attack (Raffa & Berryman, 1987). As female density increases, individuals have higher reproductive successes in terms of pupae per attack (until the effects of intraspecifc competition increase in severity) (Raffa & Berryman, 1983; Berryman et al., 1985). Cooperative actions have also been implicated in the capture of prey by social spiders (Whitehouse & Lubin, 2005). Social and subsocial spiders cooperate in the capture of insects that are too large for single individuals to subdue (Nentwig, 1985; Ward, 1986; Rypstra, 1990; Rypstra & Tirey, 1991; Pasquet & Krafft, 1992; Jones & Parker, 2000; Kim et al., 2005). The social spider, Anelosimus eximius, builds communal webs and group foraging allows capture of prey of increasing size. This offsets the decline in the number of prey caught per individual as web size and hence colony size increases. Maximum resource intake occurs at intermediate spider densities (Yip et al., 2008). Individual success rates in lion foraging are also correlated with ecological attributes such as prey type and group structure (Scheel & Packer, 1991). Groups of lions and other predators have higher success foraging than individuals (Dugatkin, 1997; Beauchamp, 2013) but this is open to exploitation by lions that refrain to engage in specific hunting bouts (Scheel & Packer, 1991).

Communal nesting in fish (Tyler, 1995) and bees (Kukuk & Sage, 1994; Kukuk et al., 1998) apparently reduces loss of offspring due to predation. Unrelated and previously unknown potential immigrants were more readily accepted if groups of the cooperatively breeding cichlid, Neolamprologus pulcher, were exposed to fish predators or egg predators than to herbivorous fish or control situations lacking predation risk (Zöttl et al., 2013b). Cooperation among queens in pleometrotic ant species (e.g., Bartz & Hölldobler, 1982; Rissing & Pollock, 1987; Rissing et al., 1989; Tschinkel, 1992a; Bernasconi & Strassmann, 1999) or among males in lion coalitions (Packer et al., 1991) allows the successful takeover of other conspecific associations. The exploitation of hosts by opportunistic pathogenic bacteria, such as Bacillus thuringiensis, involves sharing the exploits of toxin production from multiple individuals as single individuals are incapable of overcoming host defenses. The exploitation of toxin producers by cheats (non-toxin producing strains) has consequences for pathogen virulence (Raymond et al., 2007, 2009), host-pathogen epidemiology (Bonsall, 2010) and the evolution of pathogen strain specificity. By grouping, animals may reduce their individual risk of exposure to parasites by allogrooming (Hart & Hart, 1992; Mooring & Hart, 1992; Johnson et al., 2004), potentially increasing their reproductive success (Hillegass et al., 2010). In environments where night temperatures are low, birds can benefit from roosting together, thereby saving energy and reducing loss of body mass (du Plessis et al., 1994; McKechnie & Lovegrove, 2001; McGowan et al., 2006; Hatchwell et al., 2009). By huddling together in dense colonies during the bitterly cold Antarctic winter, emperor penguins (Aptenodytes forsteri), are able to share precious body heat and provide insulation for one another, thereby reducing their individual energy expenditures by up to 50 percent (Le Maho, 1977). These joint benefits are expected to be large enough to compensate for inevitable costs of group living, such as competition for local resources or increased parasite loads (Alexander, 1974). 

4.2.1 Cooperation or "altruism": direct or indirect fitness effects?

The evolution of cooperation/“altruism” (regarding the ambiguity of the terms see chapter 3) has often been attributed primarily to kin selection (whereby individuals gain “indirect” benefits to their fitness by assisting collateral relatives). Importantly, the distinction between acts of cooperation and “altruism” is crucially dependent on the comprehensive elucidation of direct fitness effects. The recent past has seen several examples where meticulous long-term field observations recognized direct fitness benefits of cooperative breeding (see chapters 4.2.1.1, 6 and 17.3) thus allowing behavior that was previously assumed to be “altruistic” to be re-categorized as cooperative. Jennifer Smith (2014) has given a comprehensive account of the direct benefits of communal living in mammals, often independent of any kin selection-related effects: 
“Long-term studies on free-living mammals suggest that exchanges of helpful behaviors, most of which occur among kin, have cumulative direct fitness consequences for individuals (Silk & House, 2011). The accumulation of social acts, such as grooming and long-term associations, enhances both the longevity and offspring survival for the vast range of mammals. Fitness consequences of sociality have now been documented in mammalian species including humans (House et al., 1988), baboons (Silk et al., 2003; Silk et al., 2010), house mice (Weidt et al., 2008), laboratory rats, Rattus norvegicus (Yee et al., 2008), horses, Equus caballus (Cameron et al., 2009), dolphins, Tursiops aduncus (Frère et al., 2010), rock hyraxes, Procavia capensis (Barocas et al., 2011) and yellow-bellied marmots (Armitage & Schwartz, 2000; Wey & Blumstein, 2012).

Nonkin cooperation of the same species also yields direct immediate or delayed benefits. For example, spotted hyaenas withhold aggression from unrelated adult females with whom they exchange other commodities important for survival (Smith JE et al., 2007). Vervet monkeys (Seyfarth & Cheney, 1983) and baboons (Cheney et al., 2010) also solicit cooperation from recent, unrelated grooming partners, presumably because of the direct benefits that donors receive from helping nonkin. Langergraber et al. (2007) used molecular genetics to tease apart the relative effects of direct and indirect benefits in philopatric male chimpanzees at Ngogo in Kibale National Park, Uganda. Interestingly, the majority of highly affiliative and cooperative dyads (e.g. pairs that formed coalitions at the highest hourly rates) were unrelated or distantly related. A recent meta-analysis by Schino and Aureli (2010) provided similar insights about allogrooming in nonhuman primates. By comparing the relative effects of kinship and reciprocity, they found that when both factors were evaluated simultaneously, the effects of reciprocity exceeded those of kinship in explaining grooming patterns. Similarly, meerkats gain direct benefits from sentinel behavior (Clutton-Brock et al., 1999). That is, rather than guarding only being favored by indirect benefits gained from helping kin, meerkats gain direct benefits from guarding; sentinels guard to reduce their own predation risk if no other animal is on guard and if they have recently eaten.”

There is increasing evidence that cooperation between unrelated individuals is wide-spread (Stacey & Koenig, 1990; Cockburn, 1998; Avilés et al., 2002; Clutton-Brock, 2002, 2009a; Dugatkin, 2002; Mesterton-Gibbons & Dugatkin, 1992). Studies have failed to find relatedness levels in accordance with expectations, e.g. in kingfishers (Reyer, 1984), dunnocks (Davies, 1992), fairy-wrens (Dunn et al., 1995), eider ducks (Öst et al., 2005), subdesert mesite Monias benschi (Seddon et al., 2005), Taiwan yuhinas (Shen, 2009), manakins (McDonald & Potts, 1994), mongooses (Creel & Waser, 1994), hyenas (Van Horn et al., 2004), coalitions of male lions (Packer & Pusey, 1982), wolves (Vucetich et al., 2004), chimpanzees (Vigilant et al., 2001; Lukas et al., 2005), greater spear-nosed bats (Bohn et al., 2009), big brown bats (Metheny et al., 2008), vampire bats (Wilkinson, 1985), oceanic delphinid species (Hayano et al., 2004; Karczmarski et al., 2005; Viricel et al., 2008; Mirimin et al., 2011), guppies (Russell et al., 2004), cichlid Neolamprologus pulcher (Dierkes et al., 2005), clown anemonefishes (Buston et al., 2007), the paper wasp Polistes dominulus (Queller et al., 2000; Zanette & Field, 2008), communal bees (Kukuk & Sage, 1994; Danforth et al., 1996; Paxton et al., 1997), and ant foundresses (Hagen et al., 1988; Sasaki et al., 1996; Bernasconi & Strassmann, 1999; Helms Cahan & Helms, 2012).

The literature often attributes non-human primate altruism and cooperation to kin selection, thus calling human cooperation with non-relatives a ‘huge anomaly’ in the animal kingdom (Fehr & Fischbacher, 2003; Gintis et al., 2003; Boyd, 2006; Melis & Semmann, 2010). Even though there is ample evidence that this claim does not hold for captive chimpanzees (de Waal, 1982, 1992, 1997a; Koyama et al., 2006), it has only recently been effectively countered for wild chimpanzees. DNA data from the field demonstrate that most of the cooperative relationships among male chimpanzees are of a reciprocal nature and concern individuals without family ties (Mitani, 2006; Langergraber et al., 2007). Bonobos may show the same pattern, since females maintain a close cooperative network that allows them to collectively dominate the males (Furuichi, 1997; de Waal, 1997b) despite the fact that females are also the migratory sex, hence largely unrelated within each community (Kano, 1992). It seems, then, that both of our closest relatives are marked by frequent cooperation among non-relatives. Taken together, these findings suggest that kin selection is not the primary reason for animals to group together (Clutton-Brock, 2002; Valsecchi et al., 2002; Avilés et al., 2004; Russell et al., 2004; Spong & Creel, 2004; Van Horn et al., 2004; Vucetich et al., 2004; Lukas et al., 2005). Enduring relationships between nonkin, same-sex individuals occur not only in humans and non-human primates (Langergraber et al., 2007, 2009; Lehmann & Boesch, 2009; Mitani, 2009; Holt-Lunstad et al., 2010; Schulke et al., 2010; Seyfarth & Cheney, 2012) but also in various other mammals (McShea, 1990; Packer et al. 1991; Möller et al., 2001; Parker & Lee, 2003; Weidt et al., 2008; Cameron et al., 2009; Seyfarth & Cheney, 2012). 

4.2.1.1 Cooperative breeders

Cooperative breeding behavior has been defined as displayed by individuals exerting costly parenting effort to contribute to the developmental success of an infant or juvenile that is not their own offspring (Emlen, 1984; Stacey & Koenig, 1990). Cooperative breeders are typically divided into two general types of cooperative societies: plural breeders and singular breeders (Brown, 1987; Hauber & Lacey, 2005). In plural (communally) breeding societies, all adult group members are reproductively active and alloparental care (i.e., care of non-offspring young) is typically performed by individuals with offspring of their own. In contrast, in singular breeding societies, production of offspring is typically limited to a single male and female per group who are assisted by a variable number of non-breeding adult alloparents (‘‘helpers at the nest’’). These categories are not absolute and populations or even social groups of conspecifics may shift breeding systems (i.e., singular, plural, or non-cooperative) between or, in some cases, within years (e.g., white-fronted bee-eaters Merops bullockoides: Emlen & Wrege, 1992; long-tailed tits Aegithalos caudatus: McGowan et al., 2003; carrion crows Corvus corone: Baglione et al., 2002; house mice Mus musculus, yellow bellied marmots Marmota flaviventris, and voles Microtus spp.: Solomon & French, 1997). There is abundant evidence for a preferential help to close kin (Reyer, 1984;Curry, 1988a; Stacey & Koenig, 1990; Emlen, 1991; Russell & Hatchwell, 2001; Baglione et al., 2003; Dickinson & Hatchwell, 2004; Hatchwell, 2009; Nam et al., 2010; McDonald & Wright, 2011; Hatchwell et al., 2014). However, cooperative breeding also occurs in some species that do not live in family groups, and more recent work has shown examples from species living in a variety of social groups where the huge variation between individuals in helping effort does not correlate with differences in genetic relatedness (Malcolm & Marten, 1982; du Plessis, 1993; Creel & Waser, 1994; Piper, 1994; Dunn et al., 1995; Hoogland, 1995; DeLay et al., 1996; Keller, 1997; Magrath & Whittingham, 1997; Heinsohn & Legge, 1999; Clutton-Brock et al., 2000, 2001a; Queller et al., 2000; Cant & Field, 2001, 2005; Clutton-Brock, 2002; Griffin & West, 2002, 2003; Field et al., 2006; Leadbeater et al., 2010; Hanson, 2013; Zöttl et al., 2013a). It has been suggested that by helping to care for unrelated young, individuals may take advantage of a kin-recognition mechanism based on associations learned by nestlings while being fed. The deceived young later may offer assistance according to its perceived relatedness to the former helper (Connor & Curry, 1995). This mechanism, termed kinship deceit, may be a form of bet-hedging in cooperative breeding systems where mortality is high, where breeders can benefit from contributions by helpers, and where helpers normally assist relatives (Connor & Curry, 1995). In 213 species of cooperatively breeding birds for which data are available, the majority of species (55%) nest in nuclear family groups, but cooperative breeding by unrelated individuals is more common than previously recognized: 30% nest in mixed groups of relatives and non-relatives, and 15% nest primarily with non-relatives. In obligately cooperative species, unaided pairs cannot rear young (e.g. dwarf mongooses, Helogale parvula: Creel, 1990a; African wild dogs, Lycaon pictus: Malcolm & Marten, 1982). Obligate cooperative breeders are far more likely to breed with non-relatives than are facultative cooperators, indicating that when constraints on independent breeding are sufficiently severe, the direct benefits of cooperative breeding—specifically, increased survival, territory inheritance and access to current or future mating opportunities—are frequently sufficient to maintain social nesting even when genetic relatedness is low (Riehl, 2013). The indirect benefits of cooperative behavior may often have been overestimated while the direct benefits of helping to the helper’s own fitness have probably been underestimated (Clutton-Brock, 2002). In general, cooperative breeders fledge their young significantly earlier and raised more broods per season than non-cooperative bird species of similar body mass. Within phylogenetic families, cooperative species have shorter nestling periods, and the duration of the nestling period tends to decline as cooperative group size increased (Ridley & van den Heuvel, 2012).

4.2.2 Group size

Although there are many ways to describe sociality, it is social group size that delineates the state space in which social systems and social complexity can evolve (Terborgh, 1983; van Schaik & van Hooff, 1983; Terborgh & Janson, 1986; Janson, 1992; Pollard & Blumstein, 2008). The general theoretical framework for explaining variation in group size assumes that there are fitness consequences associated with group size and that individuals maintain membership in groups of favorable size to maximize fitness (Wrangham, 1980). In most species, the costs and benefits associated with group living vary systematically with group size.

The individual fitness of group members as a function of group size may either consistently increase, consistently decrease or have a maximum at an intermediate group size (Brown et al., 1990). In 1931, Warder Clyde Allee proposed that populations at low numbers are affected by a positive relationship between population growth rate and density, which increases their likelihood of extinction, an idea that he later extended in his 1949 book on animal ecology (Allee, 1931; Allee et al., 1949). There is widespread evidence for the Allee effect in mammals (e.g. Komers & Curman, 2000), birds (e.g. Green, 1997) and fish (e.g. Liermann & Hilborn, 1997). Besides inbreeding depression and demographic stochasticity, a reduction in intraspecific cooperation might lead to inverse density dependence of population growth (Courchamp et al., 1999a). Factors causing an Allee effect may include dilution of predation risk at increasing group sizes (Hamilton, 1971; Treherne and Foster, 1982; Heg et al., 2004), increased predation at low population densities that follows reduced vigilance (Kenward, 1978; Berg et al., 1992), interspecific kleptoparasitism (Carbone et al., 1997), and difficulties in finding a mate (Kuussaari et al., 1998; Berec et al., 2001). The high rates of group extinction observed in obligate cooperative breeders are generated by a need for a critical number of helpers, which produce an Allee effect (Courchamp et al., 1999b; Courchamp & Macdonald, 2001). African wild dogs, Lycaon pictus, live in groups of up to 20 adults and their dependent young (Creel & Creel, 1995). The hunting strategy of the group usually requires a critical size to be energetically efficient (Fanshawe & Fitzgibbon, 1993; Creel & Creel, 1995; Creel, 1997). A threshold group size might also be required for hunting because of kleptoparasitism by hyenas, which can be energetically very costly to small groups of wild dogs (Fanshawe & Fitzgibbon, 1993; Carbone et al., 1997, 2005; Gorman et al., 1998). In addition, helpers are required by the breeding female: litters are very large (up to 20 pups), and the breeding female, then the pups, need to be fed by other members of the group (Malcolm & Marten, 1982; Woodroffe et al., 1997). Group members also help by chasing predators from the den area, and by staying at the den to protect the pups while the pack is hunting (Malcolm & Marten, 1982; Woodroffe et al., 1997). There is a critical minimum threshold, below which packs face an increasing probability of extinction – an Allee effect with consequences for the conservation of this species, and of other obligate cooperators (Courchamp & Macdonald, 2001). Theoretical models (Avilés, 1999; Kokko et al., 2001a) and empirical examples (Caraco & Wolf, 1975; Nudds, 1978; Buss, 1981; Itô, 1987; Raffa & Berryman, 1987; Heinsohn, 1992; Cash et al., 1993; Komdeur, 1994; Wiklund & Andersson, 1994; Booth, 1995; Jeanne & Nordheim, 1996; Avilés & Tufiño, 1998) show that as a result of cooperation some components of individual fitness may increase as a function of colony size and underlie the positive correlation seen between the number of helpers and the number of offspring (Creel & Creel, 1991) raised in most species with communal care (coyotes: Bekoff & Wells, 1982; birds: Brown, 1987; lions: Bygott et al., 1979; wild dogs: Malcolm & Marten, 1982; jackals: Moehlman, 1979; dwarf mongooses: Rood, 1990; Creel & Creel, 1991; badgers: Kruuk, 1989). Long-term fitness considerations can explain group size regulation in meerkats. Group size distributions expected from predicted dispersal and eviction strategies matched empirical distributions most closely when emigrant survival was approximately that determined from the field study (Stephens et al., 2005).

The many factors that influence, and are influenced by, group size include predation risk (Alexander 1974; Altmann, 1974; van Schaik, 1983; van Schaik & van Hooff, 1983; Janson, 1998), resource availability and competition (Wrangham,1980; van Schaik, 1983; van Schaik & van Hooff, 1983), foraging patch size and heterogeneity (Altmann, 1974; Leighton & Leighton, 1982; Johnson DDP et al., 2002), disease/parasite risk (Freeland, 1976; Altizer et al., 2003), body mass (Jorde & Spuhler, 1974; Clutton-Brock & Harvey, 1977; van Schaik, 1983; Janson & Goldsmith, 1995), diet (Clutton-Brock, 1975; Clutton-Brock & Harvey, 1977; Janson & Goldsmith, 1995; Williamson & Dunbar, 1999), phylogenetic inertia (Di Fiore & Rendall, 1994), life history (Wittenberger, 1980), mating opportunities (Lindenfors et al., 2004) and travel costs (Chapman & Chapman, 2000; Sernland et al., 2003). Most of these factors can be subsumed under two major categories of selective pressures: predation risk and resource needs (Dunbar, 1988; Krause & Ruxton, 2002; Hoare et al., 2004; Caro, 2005). The classic view on the evolution of group size is that observed group sizes reflect these two major factors (Clutton-Brock & Harvey, 1977; Terborgh, 1983; van Schaik & van Hooff, 1983; Terborgh & Janson, 1986; Rodman, 1988; Janson, 1992; Wrangham et al., 1993; Janson & Goldsmith, 1995; Müller & Thalmann, 2000; Pollard & Blumstein, 2008). There is empirical evidence that group size increases with increasing risk of predation, and conversely, that more solitary individuals are found when the predation risk is low, e.g. in passerine bird colonies (Wiklund & Andersson, 1994), African penguin groups (Mori, 1999), groups of rodents (Ebensperger & Wallem, 2002), frog tadpoles (Spieler, 2003) and fish shoals (Magurran & Pitcher, 1987; Hoare et al., 2004). Evidence that dominant breeders accept additional helpers more readily with increasing benefits of help exists from pied kingfishers (Ceryle rudis), where experimental clutch enlargement increased helper presence (Reyer, 1980). An opposite regulatory effect was observed in the clown anemonefish (Amphiprion percula), where dominant group members prevent immigration when the group is saturated (Buston, 2003a). Indirect evidence for group size adjustment to demand exists from baywings (Agelaioides badius), where more helpers were present when nests were parasitized by cowbirds (Molothrus spec.), which resulted in a higher begging intensity in the nest (Ursino et al., 2011). In the cooperatively breeding cichlid, N. pulcher, evicted offspring can be reaccepted when competition for the territory is experimentally increased, suggesting that breeders in this species are able to estimate their need for help (Taborsky, 1985).

It is the pattern of resource availability in both space and time that influences group size (Crook, 1964; Clutton-Brock, 1975; Macdonald, 1983; Johnson DDP et al., 2002). Increases in overall resource abundance leads to increases in habitat quality per unit area, thus resulting in higher animal densities, and typically smaller ranging areas (Reiss, 1988; Powell et al., 1997). Resource abundance in itself does not necessarily affect group size, however, because rich territories are typically contracted (or split), such that individuals maintain the former per-capita intake rate (Kruuk & Macdonald, 1985). But if resources are heterogeneous in space or time, then this is no longer necessarily true – proportional increases in average resource abundance might not enable the territory to shrink, because larger areas are still needed to encompass the temporal and spatial variability of these resources (Johnson DDP et al., 2002). Food scarcity will decrease optimal group size when large groups face more intense intragroup food competition (so that there is a ‘‘many eyes—many mouths’’ trade-off) that is not compensated by their intergroup competitive advantages (Giraldeau, 1988; Ritz, 1997). This can occur when scarcity results in less variation in food patch quality such that even the best patches are not worth defending or when food sources are less defensible (Wrangham, 1980). Food scarcity will increase optimal group size when resource defense benefits outweigh high intragroup feeding competition in large groups. This can occur if food scarcity increases the supply or defensibility of high-quality food patches, thus increasing intergroup competition (Pride, 2005). With greater group carrying capacities (more abundant resources), grouping and cooperative tendencies evolved in opposite directions; as the group carrying capacity increased, greater grouping but lower cooperative tendencies evolved (Avilés et al., 2002). The latter result is in agreement with findings of N-player Prisoner’s Dilemma games (e.g., Boyd & Richerson, 1988) or game-theoretic models of cooperative hunting (Packer & Ruttan, 1988) that have shown that larger groups are less stable and more susceptible to cheating. Because the costs and benefits of grouping vary as a function of group size, it can be expected that individuals will modify their choice of group size (by joining or leaving groups) as ecological conditions change to maximize their fitness (Pulliam & Caraco, 1984). One problem with this possibility is that individuals within a group may not be able to exclude solitary individuals, and consequently individuals will continue to join a group exceeding the ‘‘optimal group size’’ (Sibly, 1983). Thus solitary individuals, by joining a group, may increase their own fitness, but decrease that of all other group members. Theoretical considerations have revealed that optimal group size and its obvious direct-fitness benefits, but not relatedness, should determine the decision of solitary individuals to join a group (Rannala & Brown, 1994). One safe strategy for solitaries would be to only join groups that are below the optimum group size (in terms of direct fitness), and for group members of only allowing solitaries to join if the group is below the optimum size (Rannala & Brown, 1994). When the costs of grouping become greater than the benefits individuals should be expected to leave (Krause & Ruxton, 2002; Couzin & Krause, 2003). Groups fission when they become large in badgers (Da Silva et al., 1993; Newman, 2000) and other species (Macdonald, 1979; Dunbar, 1989; Baker et al., 2000). Increasing group size commonly depresses fecundity and increases mortality of females and their offspring (Clutton-Brock et al., 1982; van Schaik et al., 1983a; Clutton-Brock, 2002, 2009b; Silk, 2007; Clutton-Brock et al., 2008; Clutton-Brock & Huchard, 2013). Where reproductive competition is intense, increases in group size are often associated with increased rates of abortion, infanticide and eviction (or dispersal), which progressively restrict recruitment and constrain the upper limits of group size (Hoogland, 1995; Stephens et al., 2005; Kappeler & Fichtel, 2011). In contrast, where increasing group size has little effect on the intensity of breeding competition between group members, females may form large groups whose size is ultimately limited by the effects of competition for resources on fecundity and survival (Prins, 1996; Moss & Lee, 2011).

4.2.3 Limited dispersal

4.2.3.1 Law of Causality

The Law of Causality states: Every event must have a cause (Hughes & Lavery, 2004). Therefore we explain particular events and general patterns by identifying the causal factors involved. Ordering two or more events in a causal order is crucial for a scientific understanding. Another order of events is their temporal order. While the temporal order is observable, outside of a controlled scientific experiment the causal order is not. This is because a complete causal account specifies the necessary and sufficient conditions for something to occur and both of these conditions involve counterfactual statements (Damer, 1995; Hughes & Lavery, 2004). Counterfactual statements are about what would have happened had the purported necessary and sufficient conditions not been satisfied. These possibilities are, by definition, not observable. For these reasons, the concept of causation must be carefully distinguished from the concept of correlation. Two events that regularly occur at the same time or in the same sequence may be both correlated and related as cause and effect or they may be correlated without being in a direct causal relationship. A correlation is observed when different events occur at the same time or occur regularly in the same sequence. With causation, one event (the cause) is responsible for, or brings about, another event (the effect). We can see the need for this distinction by considering one of the causal fallacies, common cause. Someone might notice that a sore throat is always accompanied by sinus congestion (a correlation). On the basis of this observed correlation, the sick person might fallaciously believe that the sore throat causes the sinus congestion. But really, the sore throat and the sinus congestion are both caused by a third factor, namely, a cold virus. So while the sore throat and sinus congestion have a common cause, neither causes the other (Hughes & Lavery, 2004). There is a correlation but no causal relationship between the sore throat and the sinus congestion, while both are epiphenomena of their common cause, the cold virus. Hence, a spurious relationship is fallaciously assumed when two occurrences have no causal connection, yet it may be inferred that they do, due to a certain third, unseen factor (referred to as a "confounding factor"). Hence, causality evaluations should be based on three criteria: (i) Correlation: cause and effect must vary together; (ii) Time sequence: the cause must come before the effect; (iii) Non-spuriousness: the relationship between two events cannot be explained by any third variable.

4.2.3.2 Limited dispersal and cooperation: spurious relationship?

In his original papers on inclusive fitness theory, Hamilton pointed out that a sufficiently high relatedness to favor altruistic behaviors could accrue in two ways — kin discrimination or limited dispersal (Hamilton, 1964, 1971, 1972). He argued that limited dispersal would lead to what he called population viscosity, elevating local relatedness sufficiently to allow altruism towards neighbors in general (Queller, 1994a). There is a huge theoretical literature on the possible role of limited dispersal in kin selection (e.g. West et al., 2002a; Lehmann et al., 2008; Platt & Bever, 2009) as well as experimental evolution tests of these models (Griffin et al., 2004; Diggle et al., 2007b; Kümmerli et al., 2009b). The first study to challenge Hamilton’s ideas about limited dispersal was a computer simulation based on group-selection theory (Wilson et al., 1992). Discussions of population viscosity and the evolution of cooperation have emphasized the potential for kin competition to limit the evolution of cooperation in viscous populations (West et al., 2002a; ElMouden & Gardner, 2008; Grafen & Archetti, 2008; Platt & Bever, 2009). Whereas cooperative individuals are more likely to benefit kin in viscous populations, they also compete for limiting resources with these same kin. Early theoretical work found that such kin competition can strongly antagonize the benefits of kin cooperation and inhibit the evolution of cooperation in viscous populations (Grafen, 1984; Taylor, 1992a; Queller, 1992a, 1994a; Wilson et al., 1992; West et al., 2002a). Consistent with this, empirical studies have failed to find a relationship between relatedness and aggressiveness in fig wasps (West et al., 2001) and bruchid beetle larvae (Smallegange & Tregenza, 2008), suggesting that the effects of kin competition might negate any kin-selected benefits associated with being less aggressive toward kin. E.O. Wilson and others argue that kin selection acts more often as a ‘dissolutive’ than a binding force within groups (Wilson & Hölldobler, 2005; Wilson & Wilson; 2007; Nowak et al., 2010). The logic in Hamilton’s argument (1964, 1971, 1972) is based on the assumption that selfishness is the evolutionary default state of every individual. Genetic relatedness may be a means to overcome this selfish “programming”. But if the selfish/collaborative setting of each individual is plastic depending on the ecological conditions (as I will show below), the logic of the argument no longer holds.

The diversity of dispersal strategies is expected to be shaped by species specific characteristics and interspecific competition, which can balance the relative benefits and costs of dispersing, in interaction with the environment (Büchi & Vuilleumier, 2012). Some authors emphasize that the cost on direct fitness is small if personal reproduction after dispersing would have been unlikely, which happens if habitats are saturated and gaining a breeding position elsewhere is, therefore, difficult (the “ecological constraint” hypothesis; Selander, 1964; Brown, 1969; Emlen, 1982a, 1995). Others put more emphasis on the benefits that natal philopatry can bring about (Stacey & Ligon, 1987, 1991; Zack & Stutchbury, 1992) or on life-history traits, such as longevity, that predispose species to become cooperative (Arnold & Owens, 1998; Hatchwell & Komdeur, 2000). In some species, delayed dispersal occurs without offspring helping their parents (Gayou, 1986; Veltman, 1989; Ekman et al., 1994; Green & Cockburn, 2001; Kokko & Ekman, 2002). This suggests that direct benefits can suffice to explain delayed dispersal and serves as a useful reminder that the decision to help does not automatically follow from the decision to stay (Emlen, 1982a, 1982b; Brown, 1987; Kokko et al., 2001a). A variety of direct benefits can favor philopatry (Kokko & Ekman, 2002): subordinates may breed (females) or gain paternity (males) despite the presence of a dominant (Arnold, 1990a; Brown & Brown, 1990; Rabenold et al., 1990; Jennions & Macdonald, 1994; Sherman et al., 1995; Laranzo-Perea et al., 2000), they may eventually inherit the dominant position within the group (territorial inheritance; Woolfenden & Fitzpatrick, 1978; Zack & Stutchbury, 1992; Russell & Rowley, 1993; Ragsdale, 1999; Queller et al., 2000), or they may gain a breeding position elsewhere after having spent time in the group (Walters et al., 1988; Zack, 1990; Zack & Stutchbury, 1992; Russell & Rowley, 1993; Green & Cockburn, 2001). These advantages may co-occur with a “safe haven” mechanism, due to improved survival of offspring in the natal territory (Ekman et al., 2000). If individuals have to wait for breeding positions, survival during the waiting period becomes an important predictor of fitness (Faaborg & Bednarz, 1990; Ekman et al., 2000; Green & Cockburn, 2001). Philopatric Siberian jay (Perisoreus infaustus) offspring have an odds ratio of being killed by predators 62% lower than offspring that dispersed promptly after independence to join groups of unrelated individuals (20.6% versus 33.3% winter mortality) (Griesser et al., 2006).

Dispersers typically have a phenotype allowing them to colonize unoccupied space, but this same phenotype is selectively disadvantageous under crowded conditions (e.g. Duckworth & Kruuk, 2009). Dispersal from the natal group often has high costs unless dispersing females can join other breeding groups (Berger, 1987; Van Vuren & Armitage, 1994; Nunes, 2007; Ronce, 2007; Doligez & Part, 2008; Strier, 2008; Clutton-Brock & Lukas, 2012). Dispersing meerkats have lower rates of weight gain while foraging than residents, levels of glucocorticoids rise and they suffer higher parasite load (Young & Monfort, 2009). As well as affecting survival, the energetic costs of dispersal may delay breeding and reduce reproductive potential (Ronce, 2007; Fisher et al., 2009). Under the risk of predation, subordinates of the cooperatively breeding cichlid Neolamprologus pulcher stay at home rather than leave and breed independently (Heg et al., 2004). Individuals that disperse to unfamiliar areas may also be more vulnerable to predators: for example, experimental studies show that dispersing white-footed mice are more susceptible than residents to predation by owls (Metzgar, 1967). In addition, they are often likely to be attacked by members of resident groups, sometimes with fatal consequences (Fritts & Mech, 1981; Packer & Pusey, 1982; Boydston et al., 2001; Creel & Creel, 2002). High rates of mortality in dispersers may be particularly common in carnivores, where attacks by residents are likely to be dangerous (Waser, 1996): for example, grey wolves making extra-territorial forays die at five times the rate of residents (Messier, 1985) while mortality rates in dispersing African wild dogs are 2.7 times higher than those of residents in females and 6.8 times higher in males (Creel & Creel, 2002). Several studies have also shown that dispersal is associated with substantial increases in mortality in many other species (Errington, 1963; Van Vuren & Armitage, 1994): for example, in red howler monkeys, 43–52% of dispersing females are suspected or known to die (Crockett & Pope, 1993).

Uncertainty can be measured as the variance of a distribution of environmental quality, and adversity as the mean (Andras et al., 2003). Both adversity and uncertainty have been conceptualised as aspects of environmental ‘risk’ (Daly & Wilson, 2002). Risky environments can be expected to increase the costs of dispersal. As well as suffering the direct costs of moving between groups, dispersing females lose the potential benefits of associating with kin and this may have an important influence on the probability that they will disperse (Lambin et al., 2001; Perrin & Goudet, 2001; Silk, 2007). Studies have suggested positive effects of kin on philopatry and sociality in a wide variety of systems, including spiders (Jones & Parker, 2002), lizards (Davis, 2011), rodents (reviewed in Sherman, 1981; Lambin et al., 2001), primates (reviewed in Silk, 2002), and birds (Eikenaar et al., 2007; Hatchwell, 2009). In various mammals associating with kin can have important benefits to breeding success and/or survival (Fairbanks & Mcguire, 1987; Arnold, 1990a, b; Crockett & Pope, 1993; Lambin & Krebs, 1993; Moses & Millar, 1994; Lambin & Yoccoz, 1998; Pusenius et al., 1998; Pomeroy et al., 2000, 2001; Pavelka et al., 2002; Hacklander et al., 2003; Cheney et al., 2004; Silk, 2007; Krebs et al., 2007; Fedigan et al., 2008; Clutton-Brock & Lukas, 2012). Unconditional dispersal propensity, the baseline tendency for individuals to relocate independent of their current environment, can be favored in a stable environment as a means of reducing kin competition (Hamilton & May, 1977; Comins et al., 1980; Motro, 1983; Frank, 1986; Gandon & Michalakis, 1999; Perrin & Goudet, 2001) or the cost of demographic stochasticity (Travis & Dytham, 1998; Parvinen et al., 2003), but is disfavored by variation in fitness over space (Hastings, 1983; Holt, 1985). Ecological drivers of dispersal are environmental spatiotemporal variability and stochasticity, i.e. habitat heterogeneity, availability and distribution (Gadgil, 1971; Comins et al., 1980; Levin et al., 1984; McPeek & Holt, 1992) that induce a cost for dispersing individuals, as they face the risk to end up in unsuitable habitats (Hastings, 1983; Morris, 1991), and thus select against dispersers. In an uncertain, unpredictable environment it should be selectively favored to reduce uncertainty at least to some degree by associating with familiar individuals. Multi-species models indicate that the fraction of a community richness that comprises mutualistic versus exploitative relationships is greatest with low dispersal ability (Filotas et al., 2010a), consistent with empirical work in arid systems (Kéfi et al., 2008). Function, diversity and interspecific interactions of locally linked communities undergo a phase transition with changes in the rate of species dispersal. Low spatial interconnectedness favors the spontaneous emergence of strongly mutualistic communities which are more stable but less productive and diverse. On the other hand, high spatial interconnectedness promotes local biodiversity at the expense of local stability and supports communities with a wide range of interspecific interactions (Filotas et al., 2010b).

At many levels of life, from plants to human societies, cooperation thrives in conditions where the environment is most adverse and unpredictable (see chapter 15.2.1). Plants at lower temperatures and higher altitudes, where abiotic stress is high, compete less and cooperate more with their neighbors (Callaway et al., 2002); bacteria and amoebae socialize in adverse environments (see chapter 5); nematodes Caenorhabditis elegans aggregate in response to stressors (De Bono et al., 2002); animals form more cohesive or larger groups, with consequent greater mutualistic benefits under greater predation risk (Seghers, 1974; Farr, 1975; Dunbar, 1988; Spieler, 2003; Krams et al., 2010); mole-rats, a highly social species, delay dispersion more in arid than in mesic habitats (Spinks et al., 2000); human in-group solidarity is greatest when the group is under threat or in a harsh environment (Levine & Campbell, 1972; Goody, 1991; Hogg, 1992; Hewstone et al., 2002). Likewise, adverse, uncertain, and unpredictable environments favor cooperative behavior (Andras et al., 2007). In chapter 9, I cite a text passage in which Laura Betzig (2004) succinctly highlighted the interrelationship between limited dispersal and asymmetric conflict/dominance hierarchies underlying coerced cooperation in human societies. The environmental conditions that favor cooperativity discourage dispersal and promote philopatry, thus shaping the preferential kinship structure of cooperative communities. Thus, limited dispersal and cooperation are epiphenomena of their common “confounding factor” on which both depend: environmental heterogeneity and stochasticity. This shared causal factor results in a spurious relationship.

5. The social games of microbes: cooperation, altruism, cheating or exploitation?


From the elephant to butyric acid bacterium—it is all the same!
Albert Jan Kluyver, 1926

Anything found to be true of E. coli must also be true of elephants.
Jacques Monod, 1954

Summary
Understanding the ecological conditions that favor, and the genetic “fossil record” that regulates, social behavior in prokaryotes and “basal eukaryotes” holds the key to understand the evolutionary rationale of sociality. The natural lifestyle of microbes is characterized by inherent “feast and famine” cycles, limiting amounts of nutrients being rather the rule than the exception, long periods of nutritional deprivation being punctuated by short periods that allow fast growth. The feast-to-famine transition is not merely a response to a drop in nutrient availability; this transition also involves cooperative cell-to-cell signaling pathways and social aggregation, the results of which range from sporulation to fruiting body and complex pattern formation. Both the bacteria Myxococcus xanthus and amoeba Dictyostelium discoideum are single-celled under abundant resources but socialize under metabolic stress. Largely, stochastic processes route clonal cells into different cell fates. Cells destined to become metabolically dormant, resilient, spores coerce other cells into death by means of toxins. The resources that thereby become available fuel the metamorphosis to spores in an otherwise resource-depleted environment.
Cooperative assembly of unicellular organisms into multicellular aggregates, mounds and fruiting bodies has evolved many times in evolution and is common in nature. When exposed to harsh environmental conditions such as starvation, hard surfaces, heat, and hazardous chemicals, bacteria show cooperative behavior with the formation of complex colonies with different spatiotemporal patterns that represent efficient strategies for adaptation and survival. Biofilms are large, three-dimensional aggregates of bacteria with several properties that require differentiation and division of labor. Microbial cooperation is often associated with public goods that can increase the local carrying capacity, automatically creating a social trait-group structure. Bacteriocins have been found in all major lineages of Bacteria and Archaea and play a major role in communication and cooperation of cells. Almost all bacteriocins are synthesized during postlogarithmic growth when both food and space are exhausted. Under metabolic stress, Bacillus subtilisStreptococcus pneumonia and a host of other bacterial species commit siblicide/fratricide and cannibalism, using the remains of their kin as resource for sporulation, competence for genetic transformation or biofilm formation. Siblicide and the extracellularDNA it provides appear to have an important role in biofilm and matrix formation. The latter functions as a permeability barrier to limit both the diffusion of beneficial nutrients away from the biofilm and prevent or slow the diffusion of harmful substances. Biased by kin selection theory these siblicides, although clearly enforced, are interpreted as altruistic sacrifice.
In the “veil of ignorance” thought experiment, individuals choose between alternative social arrangements in a society. Ignorant of the position they will occupy in that forthcoming society individuals establish rules that they regard as fair from behind a veil of ignorance. As a measure of risk or inequity aversion, equal allocations of resources are chosen. Individuals consistently dislike unfair social outcomes but rather prefer playing risky but fair social lotteries. When it comes to distributing non-divisible, scarce goods in social competitions, a fair lottery excludes unjust influences. In evolutionary biology, the veil-of-ignorance concept has been used to explain the randomization of alleles in fair meiosis where each gene has a 50% chance to make it into any gamete.
In colonies of clonal cells that are metabolically stressed the competition for scarce resources is decided by a fair lottery. Behind the “veil of ignorance”, the competitors are not “aware” of their relative position in the competitive hierarchy. This hierarchy is determined stochastically through a variety of cellular processes with inherent noise that renders the cells heterogeneous and the lotteries fair.

According to Mayr (1961) evolutionary biology differs from functional biology because it also asks about the “why” of behavior: “It is obvious that the evolutionist has in mind the historical ‘how come?’ when he asks ‘why?’ Every organism, whether an individual or a species, is the product of a long history, a history that dates back more than 2000 million years.” Since the evolutionary past holds the key to understand the evolutionary present, this chapter is intended to trace the ecological conditions and genetic “fossil record” (Runnegar, 1986; Buss, 1987 p. 90; Heininger, 2012, 2013) of social behavior in prokaryotes and “basal eukaryotes”. Aggregation and the formation of stalks for dispersal of resistant and metabolically quiescent spores allow social bacteria and cellular slime molds to escape deteriorating local conditions (Bonner, 1982; Velicer et al., 2000). Importantly, experiments with bacteria (Travisano & Velicer, 2004) and social amoeba (Strassmann et al., 2000a) suggest that the ‘decision’ to cooperate or defect (in a general sense) is stochastic, and moreover that these ‘decisions’ are controlled by genetically-encoded probabilities that are evolvable (Khare et al., 2009). (When speaking of ‘decisions’, use of the term is in an evolutionary sense, not implying any conscious rationalization on the part of individual organisms.) Molecular biological processes underlying these ‘decisions’ are exemplarily illustrated with reference to the social bacteria Myxococcus xanthus and amoeba Dictyostelium discoideum.

The social interactions of Myxobacteria and D. discoideum highlight that
(i) Both cooperation and competition phenomena are shaped by ecological conditions and are context-dependent
(ii) Interactions are dynamic
(iii) Winners and losers of social interactions are largely determined by stochastic processes.  

5.1 Myxococcus xanthus

Myxobacterial cells are single-celled but social; they swarm by gliding on surfaces as they feed cooperatively. Upon nutrient deprivation, populations of the bacterium M. xanthus migrate towards high-density focal points and cluster into aggregates of approx. 100,000 rod-shaped individuals (Shimkets, 1999). These aggregates form a multicellular fruiting body in which a fraction of the cells develop into myxospores. During the process of aggregation, early mound and fruiting body formation, approximately 80% of the cells undergo cell lysis, while the remaining 20% are converted into myxospores. Depending on the particular conditions of development, however, the proportion of dying cells may also range from 20 to 90%. Spore development includes the differentiation from the rod-shaped vegetative cell to a spherical, nonmotile, environmentally resistant and metabolically quiescent myxospore. Both sporogenesis and programmed cell death in the cells building the fruiting body are effected by a toxin-antitoxin system (Nariya & Inouye, 2008). One pathway governs aggregation and sporulation of some cells in the starving population and requires the so-called C-signaling. The levels of the bifunctional transcription factor/antitoxin MrpC and its related proteolytic fragment MrpC2 are increased, inhibiting the cell death pathway in the spores via direct interaction of MrpC with the MazF toxin (Nariya & Inouye, 2008). Another pathway causes programmed cell death that requires the MazF toxin (Mittal & Kroos, 2009a, 2009b) whose endoribonuclease activity ultimately results in cell death (Nariya & Inouye, 2008; Mittal & Kroos, 2009a). In cells destined to undergo programmed cell death, binding of MrpC to the mazF promoter region would activate transcription, leading to increased MazF toxin levels. Accordingly, MrpC is a key determinant of cell fate (Mittal & Kroos, 2009b). Stochastic cell fate appears to be determined by developmental heterogeneity and feedback loop-based bistability (Russo-Marie et al., 1993; Lee, 2009; Bhardwaj, 2013, see chapter 5.4.1) involving the Spo0A transcriptional regulator (González-Pastor et al., 2003; Ellermeier et al., 2006). This bistable switch, cued initially by stochastic variations in gene expression at the single-cell level, divides a population of genetically identical cells into alternate cell fates (Søgaard-Andersen & Yang, 2008). Cannibalism is considered to provide essential nutrients for the conversion of rod-shaped cells to mature spores (Wireman & Dworkin, 1977; Dworkin, 1996). Deletion of the toxin results in elimination of the obligatory cell death during fruiting body development causing dramatic reduction in developmental progression and spore formation (Nariya & Inouye, 2008; Søgaard-Andersen & Yang, 2008) suggesting that cell death is a necessary component to complete sporogenesis that depends on energy and building blocks provided by the apoptotic cells. Interestingly, the toxin-antitoxin system is also linked to persisters (Moyed & Bertrand, 1983). Persisters are cells that neither grow nor die in the presence of bactericidal antibiotics (Lewis, 2010). These cells are not mutants but phenotypic variants of regular cells that form stochastically in microbial populations (Lewis, 2010), representing a type of microbial bet-hedging (Heininger, 2015).

There are two developmental strategies for myxobacteria. One is a complicated social behavior: fruiting body structures form from numerous cells, and the cellular development is dependent on cell density. A fruiting body ensures that a new life cycle is able to start in a terrestrial environment (Kraemer et al., 2010). The other is simple: the social behavior is an evolutionary burden in asocial habitats like seawater. In asocial habitats myxospores form directly from vegetative cells, and the cells grow independent of the cell density (Zhang et al., 2005). It seems that the adaptation of the halotolerant myxobacteria to a marine environment is a degenerative procedure (Zhang et al., 2005). There could be a shift between the two developmental patterns if the cells migrate between the two environments (Li et al., 2011).

5.2 Dictyostelium discoideum

The social amoeba or cellular slime mould Dictyostelium discoideum is at the transition of uni- and multicellularity and a valuable model organism for a variety of biomedical phenomena including social interactions (Annesley & Fisher, 2009; Heininger, 2012). D. discoideum is single-celled under conditions of nutritional abundance. When their bacterial food source is exhausted, as many as 100,000 cells aggregate to form a multicellular migrating slug, which moves toward a region suitable for culmination. This eventually becomes a fruiting body in which some amoebae in the group differentiate to form hardy spores and other microbes die to form a stalk structure. Most of the prestalk cells are located in the front one-fifth of the slug, whereas all of the prespore cells are located in the rear four-fifths. One apparent difference between the prespore and prestalk populations is in their degree of heterogeneity: there are several prestalk-cell subtypes, whereas, apart from an anteroposterior gradient in their utilization efficiency of certain prespore–promoter constructs (Haberstroh & Firtel, 1990), the prespore cells seem to be a homogeneous population (Williams, 2006). Importantly, the multicellular structure is able to pass through soil barriers that solitary amoeba cannot cross and lifts the spores above the soil where spores can be picked up for long-distance dispersal (Kuzdzal-Fick et al., 2007). Both stochastic and deterministic processes regulate migration and cell fate decisions (Miyanaga et al., 2007; Gregor et al., 2010; Prindle & Hasty, 2010; Chattwood & Thompson, 2011; Chattwood et al., 2013). A lottery system decides who becomes a spore and who does not (Gadagkar & Bonner, 1994). There is a manipulation of the aggregated cells so that the fertile spores become the reproductives and they literally suppress, by producing an inhibitor, some of the cells from following suit; they force those cells to become sterile stalk cells (Inouye, 1989). As in developmental programs of multicellular organisms (Heininger, 2001, 2013), prespore cells coerce prestalk cells into an apoptotic death program and the death of stalk cells fuels the energetically costly metamorphosis of the spore cells (Heininger, 2001, 2012). A closer look at the molecular biological events that regulate the social interactions elucidates the underlying exploitation in clonal populations. Prespore cells determine the fate of apoptosing prestalk cells by differentiation-inducing factors (DIFs), chlorinated alkylphenones, and other factors (Early & Williams, 1988; Maruo et al., 2004). D. discoideum amoebae are induced to differentiate into prestalk cells by the signalling molecule DIF-1 (Thompson & Kay, 2000a). The DIFs are synthesized in and released by prespore cells (Kay & Thompson, 2001) and are inactivated in prestalk cells (Brookman et al. 1987, Kay et al., 1999; Thompson & Kay, 2000a; b; Kay & Thompson, 2001), establishing a feedback loop that controls DIF-1 levels (Insall et al., 1992). The inactivation, however, is inhibited by cyclic AMP (cAMP) (Insall et al., 1992) that plays a pivotal role in chemotaxis, differentiation and response to stress (Saran et al., 2002). DIF-1 is rather a poison than a signal (Atzmony et al., 1997; Parkinson et al., 2011; Strassmann & Queller, 2011) and acts antagonistically to cAMP by repressing prespore differentiation and directing a proportion of the cell population to differentiate as prestalk cells (Kay & Jermyn, 1983; Early & Williams, 1988; Kawata et al., 1996; Strassmann & Queller, 2011b). At low levels DIFs may induce prespore differentiation (Oohata, 1996) but at higher doses are requisite for prestalk cell development (Kopachik et al., 1983; Sobolewski & Weeks, 1988; Maruo et al., 2004). Mitochondrial malate dehydrogenase (mMDH) may be one of the target molecules of DIF-1: DIF-1 inhibits the enzymatic activity of mMDH and cell energy production, probably leading to the inhibition of proliferation (Matsuda et al., 2010) and cell death (Luciani et al., 2009). Prespore-produced DIF-1 inhibits redifferentiation of prestalk cells into prespores, i.e. transition into the dispersing and perennial germline (Firtel, 1995; Hudson et al., 2002). Thus, prespore cells inhibit the conversion of prestalk to prespore cells (Inouye,1989, Shaulsky & Loomis, 1993; Kawata et al., 1996; Söderbom & Loomis, 1998; Maruo et al., 2004; Yamada et al., 2010) and ensure the supply of nutrients and building blocks by apoptotic stalk cells. The need to procure these resources in a resource-depleted environment regulates the dynamic and context-dependent relative proportion of prespore and prestalk cells. The balance is dynamic and responds to the loss of the respective other cell type. Cells that are initially destined to become prespore cells become apoptotic stalk cells when prestalk cells are removed from the equilibrium (Nadin et al., 2000; Ràfols et al., 2001) and prestalk cells become prespore cells when prespore cells are removed (Shaulsky & Loomis, 1993; Ràfols et al., 2001; Maruo et al., 2004). It can be inferred that survival and death of cells depends on environmental cues, the proportion of dying cells being regulated by the metabolic needs of the developing germ cells and the spore-dispersing role of fruiting bodies.

The conceptual framework of kin selection theory biased the interpretation of D. discoideum developmental skew. Despite the robust evidence that during clonal development prespore cells coerce prestalk cells into cell death, stalk cell death has consistently been regarded as being altruistic (Atzmony et al., 1997; Strassmann et al., 2000a; Hudson et al., 2002; Brännström & Dieckmann, 2005; Li & Purugganan, 2011). On the other hand, unequal representation of genotypes within the sporehead of D. discoideum chimera has been taken as evidence for social exploitation (Strassmann et al., 2000a; Fortunato et al., 2003). One interpretation of these results is that there are widespread fixed cheating strategies (Hudson et al., 2002). In fact, however, both D. discoideum clonal and chimeric development deploys a substantial element of cell competition (see chapter 4.1.1). Importantly, during clonal development all D. discoideum genotypes examined differ significantly from one another in their constitutional proportion of developmental spore and stalk cell formation (Buttery et al., 2009). Intriguingly, culture conditions may give rise to differences in DIF sensitivity of genetically identical amoebae (Leach et al., 1973; Thompson & Kay, 2000b; Castillo et al., 2011). Cells with lowered DIF sensitivity, resulting from growth with glucose (see chapter 10.3), “cheat” and are overrepresented in the sporeheads when mixed with microbes grown without glucose (Leach et al., 1973; Thompson & Kay, 2000b). Similarly, in chimeras between cells that were weakened by growing them in media lacking glucose, or more acid media than normal, and normally cultured cells that were genetically identical, weakened cells were outcompeted by healthy cells in becoming spores (Castillo et al., 2011). A gene, dimA, in D. discoideum has two contrasting effects. It is required to receive the signalling molecule DIF-1 that causes differentiation into prestalk cells. Ignoring DIF-1 and not becoming prestalk should allow cells to “cheat” by avoiding the stalk. However, in aggregations containing the wild-type cells, lack of the dimA gene results in exclusion from spores (Foster et al., 2004). This pleiotropic linkage of stalk and spore formation limits the potential for cheating in D. discoideum because defecting on prestalk cell production results in an even greater reduction in spores.

In realistic starvation conditions, up to 15% of cells do not aggregate. In short-term starvation, non-aggregating cells have an advantage over cells in aggregates since they resume growth earlier upon arrival of new nutrients, but have a shorter lifespan under prolonged starvation. The partitioning of cells into aggregating and non-aggregating fractions is optimal in fluctuating environments with an unpredictable duration of starvation periods. D. discoideum thus constitutes a model system lying at the intersection of microbial cooperation and solitary bet-hedging (Dubravcic et al., 2014).

5.3 Microbial cooperation and competition

The natural lifestyle of microbes is characterized by inherent “feast and famine” cycles, limiting amounts of nutrients being rather the rule than the exception, long periods of nutritional deprivation being punctuated by short periods that allow fast growth (Kolter et al., 1993, Msadek, 1999; Navarro Llorens et al., 2010; Heininger, 2012). The feast-to-famine transition is not merely a 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 & Dworkin, 1997; Shimkets, 1999). These transitions are characterized by key events in shedding light on the dynamics and patterns of microbial sociobiology.

Cells alter their gene expression in response to each other and to the local microenvironment (Thiery & Chopin, 1999; Eichenberger et al., 2004; Waters & Bassler, 2005; Battesti et al., 2011), and cell groups can evolve rapidly (Kreft, 2004a; West et al., 2006; Merlo et al., 2006; Nadell et al., 2009, 2013; Xavier, 2011, Le Gac et al., 2012). Arguably the most complex social behaviors occur in two spore-forming species: Bacillus subtilis and Myxococcus xanthus (Nadell et al., 2009, see chapter 5.1). Developmental processes display features of both cell cooperation and competition. Aggregation by these species comes closest to multicellular development: sporulation in both results from a series of temporally separated checkpoints that produce a predictable, directional sequence of differentiation events (Monds & O’Toole, 2008). B. subtilis provides a dramatic example of bacterial differentiation; populations of a single strain display bimodal distributions for multiple phenotypes, including competence, chaining (Losick & Desplan, 2008), motility (Kearns et al., 2004), production of extracellular polymeric substances (Vlamakis et al., 2008), and spore formation (Branda et al., 2001). A subpopulation of cells in wild B. subtilis biofilms produces spores in small fruiting body structures that develop at the tips of biofilm surface irregularities. Larger groups of spores are formed by M. xanthus, in which starving cells aggregate to form a fruiting body whose cell density approaches that of many biofilms (Julien et al., 2000). It is also striking that in both species spore-destined cells kill other members of the population during fruiting body formation (Wireman & Dworkin, 1977; González-Pastor et al., 2003; Nadell et al., 2009). In B. subtilis, cell death releases nutrients that increase the viability of the remaining population, allowing it to delay commitment to sporulation (González-Pastor et al., 2003).

5.3.1 Cooperation

Cooperative assembly of unicellular organisms into multicellular aggregates such as biofilms, mounds and fruiting bodies has evolved many times in evolution and is common in nature (Bonner, 2000; Claessen et al., 2014). The structural complexity and degree of organization of microbial multicellular structures vary from a simple single-layer biofilm and simple aggregates to complicated structures like the fruiting bodies of myxobacteria and slime molds, complex natural biofilms and the colonies of various microbes (Watnick & Kolter, 2000; Crespi, 2001; Palková & Váchová, 2006; Annesley & Fisher, 2009; Velicer & Vos, 2009; Cáp et al., 2012; Elias & Banin, 2012). O’Toole et al. (2000) describe biofilm development as a process of microbial development, not unlike that observed in fruiting-body formation by M. xanthus and sporulation in B. subtilis. Biofilms are large, three-dimensional aggregates of bacteria—usually comprised of several metabolically diverse species but sometimes hundreds of them—which adhere to surfaces in moist or watery environments. These include soils, teeth, living tissue, medical implants, air conditioning systems, pipelines, sewage treatment plants, and marine equipment—just about any sort of surface in an aqueous environment (Costerton et al., 1987; Kroes et al., 1999; Ram et al, 2005; Ley et al., 2006b; Lyon, 2007). When exposed to harsh environmental conditions such as starvation, hard surfaces, extreme heat, and hazardous chemicals, bacteria show cooperative behavior with the formation of complex colonies with different spatiotemporal patterns that represent efficient strategies for adaptation and survival (Shapiro, 1988; 1998; Budrene & Berg, 1991; Ben-Jacob E, et al., 1994; 2000; Harshey, 1994; Heininger, 2001; Ben-Jacob, 2003) and that are coordinated by cell-to-cell signaling (Kaiser & Losick, 1993; Kaiser, 1996; Wirth et al., 1996; Watnick & Kolter, 2000; Ben-Jacob et al., 2004; Bassler & Losick, 2006; Ben-Jacob, 2009). Cooperative biofilm formation tends to peak at intermediate frequencies of disturbance but the peak shifts toward progressively higher frequencies of disturbance as resource supply increases (Brockhurst et al., 2010). Saccharomyces cerevisiae is also able to agglomerate cells into organized structures such as flocs, biofilm, and pseudohyphae. These non-sexual clumps create protective shields for cells in inner parts to escape from abiotic/biotic attacks (Smukalla et al. 2008), and to respond to nutrient starvation (Gimeno et al., 1992; Cullen & Sprague, 2000). Biofilm maturation is controlled by the availability of nutrients and quorum sensing (i.e. the ability of cells to sense a high population density) (Stanley & Lazazzera, 2004). Biofilms can be thick, homogeneous mats of cells, or they can be complex structures composed of pillars with water channels that have been proposed to allow for nutrient influx, oxygen flow and waste efflux (Davey & O’Toole, 2000). Biofilm communities exhibit several properties that are thought to require differentiation and division of labor among cells to produce highly resilient biofilm structures (Stoodley et al., 2002). Mature biofilms of Pseudomonas aeruginosa were shown to have a radically different protein profile from planktonic bacteria grown in chemostats (Sauer et al., 2002). As much as 50% of the detectable proteome (over 800 proteins) was shown to have a sixfold or greater difference in expression. Of these, more than 300 proteins were detectable in mature biofilm samples that were undetectable in planktonic bacteria.

Microbial cooperation is often associated with public goods that can increase the local carrying capacity (West et al., 2007b; Platt & Bever, 2009). Public goods are those produced by an individual and shared with other group members (Driscoll & Pepper, 2010), automatically creating a social trait-group structure (Wilson 1975, 1980). They typically increase the fitness of all group members, but at a unilateral cost to the producer. Examples include the production of siderophores (West & Buckling, 2003), viral replication enzymes (Turner & Chao, 1999), specialized resources (Guyon et al., 1993; Gordon et al., 1996) and secreted exoenzymes (Hillesland et al., 2007; Koschwanez et al., 2011). The depth of the mature biofilm appears to be regulated to allow for maximal nutrient acquisition. Far from being a case of pure Darwinian competition, interactions among these species and with eukaryotic hosts may be mutually beneficial (Wingreen & Levin, 2006). A recent case in point is the discovery of a mutualistic interaction of four bacterial species with the tomato plant (Botta et al., 2013). Rather than competing, the four species coexist and strongly promote plant growth by fixing nitrogen, providing growth hormones, and preventing hostile bacterial species from growing.

Several stress conditions induce microcolony formation (Webb et al., 2003a; Jefferson, 2004). The most ubiquitous environmental trigger appears to be nutrient stress, suggesting that aggregation and formation of multicellular structures is an adaptive response that mediates survival under unfavorable conditions. As final common pathway of cellular stress responses (see chapter 9.3), oxidative stress triggers biofilm and fruiting body formation (Schembri et al., 2003; Murphy et al., 2005; Wen et al., 2005; Sampathkumar et al., 2006; Boles & Singh, 2008; Geier et al., 2008; Cáp et al., 2012). Escherichia coli and Salmonella typhimurium for example, when exposed to oxidative stress, attract neighboring cells that allows the cooperative degradation of toxic materials (Budrene & Berg, 1991; Blat & Eisenbach, 1995). The extracellular polymer matrix of biofilms provides protection from desiccation, toxins and antibiotics, and it might also serve to bind and hold nutrients and enhance physiological stability (Decho, 1994). Biofilm-associated cells are more resistant to many toxic substances such as antibiotics, chlorine, and detergents (Costerton, 1987; Costerton et al., 1999; Hogan & Kolter, 2002). There is evidence that decreased diffusion into the biofilm (De Beer et al., 1994; Suci et al., 1994), decreased bacterial growth rate in a biofilm (Evans et al., 1990), biofilm-specific substances such as exopolysaccharide (Yildiz & Schoolnik, 1999), and the quorum-sensing specific effects (Davies et al., 1998, Hassett et al., 1999) may be reasons for this resistance. In addition, a strong advantage to polymer production arises as an emergent property (see chapter 16). Secretion of extracellular polymers by a cell allows it to push descendents into a more oxygen-rich environment. At the same time, it provides a strong competitive advantage at the scale of the cell lineage by suffocating neighboring nonproducers (Xavier & Foster, 2007). This pleiotropic linkage of competitive advantage and polymer production limits the potential for cheating in biofilms. The effect of polymer production, therefore, has a strong analogy in plants competing for light, where vertical growth and increased foliage area selfishly increase access to light at the expense of competitors (Goodnight et al., 1992).

The human oral cavity is one of the most densely populated sites of the human body, consisting of as many as 600 to 800 bacterial species (Paster et al., 2006; Dewhirst et al., 2010). Tooth biofilms have been shown to consist of stable consortia of hundreds of distinct species, and bacterial mats are believed to consist of even larger numbers of species, in dynamic equilibrium among themselves, and with multiple bacterial viruses (Wingreen & Levin, 2006). These microbial inhabitants have co-evolved not only with their host, but also with each other, leading to extensive intercellular communications across species. There is evidence that oral microbes participate in intercellular communications with co-inhabitants through cell contact-dependent physical interactions, metabolic interdependencies, as well as coordinative signaling systems to establish and maintain balanced microbial communities (Bamford et al., 2009; Guo et al., 2014; Sztajer et al., 2014). Cooperative metabolic interactions either via cross-feeding or through cooperatively metabolizing substrate maximize co-residents’ potential to extract energy from limited substrates. In addition to synergistic interactions, oral bacterial species are also engaged in intense competition for limited space and nutritional resources using compounds such as bacteriocin and H2O2, which plays a crucial role in defining the structure and activity of oral microbial communities (Guo et al., 2014).

5.3.2 Competition and siblicide

In search of food and space to multiply, bacteria secrete ribosomally synthesized polypeptides that cause the selective killing of competing microorganisms. The antimicrobial compounds are called bacteriocins, and usually have a narrow spectrum of activity as they kill only closely related bacteria, which compete for the same resources (Heininger, 2001; González-Pastor et al., 2003; Guiral et al., 2005; Ellermeier et al., 2006; Håvarstein et al., 2006; Claverys and Håvarstein, 2007; Be’er et al., 2009, 2010). Also bacteria competing with unrelated or distantly related strains for limited resources in the same niche cooperate to secrete antibacterial compounds as “chemical weapons” to attack the competing strains (Czárán et al., 2002; Eijsink et al., 2002). Bacteriocins have been found in all major lineages of Bacteria and Archaea (Torreblanca et al., 1994; Riley & Wertz, 2002). Within a species tens or even hundreds of different kinds of bacteriocins are produced (James et al., 1991). According to Klaenhammer (1988), 99% of all bacteria make at least one bacteriocin and the only reason more have not been found is that few researchers have looked for them. Several studies have shown the involvement of bacteriocins and bacteriocin-like peptides in, for example, quorum sensing (Diep et al., 1995; Kuipers et al., 1995; Kleerebezem et al., 1999; Kleerebezem & Quadri, 2001; Eijsink et al., 2002; Kleerebezem, 2004; Gobbetti et al., 2007) and remodelling of sugar metabolism in lactic acid bacteria (Opsata et al., 2010), suggesting that these peptides also play a role in communication and cooperation of cells. Bacteriocins may also act as communication signals in bacterial consortia, e.g., biofilms (Gillor, 2007). Almost all bacteriocins are synthesized during postlogarithmic growth when both food and space for bacterial multiplication are exhausted (Baba & Schneewind, 1998) or during stress (Gillor et al., 2008). Bacteriocins and antibiotics are frequently upregulated by stress responses to nutrient limitation and cell damage but very rarely upregulated by stress responses to heat or osmotic stress, which typically are not competition related (Cornforth & Foster, 2013). Further growth can then only occur at the expense of bacterial competitors, and the fight for survival employs some exquisitely designed compounds that are toxic for specific microbial target cells. Stress-related genes were found to be upregulated in biofilm settings. For instance, genes regulated by the SOS system (involved in repairing DNA damage), such as recAdinI and sulA, were upregulated by twofold or more in E. coli biofilms compared with their respective expression in stationary-phase planktonic cultures (Beloin & Ghigo, 2005; Beloin et al., 2008). Antibiotic activities can be obtained from many different peptide sequences and structures, but bacteriocins generally share the fate of extensive posttranslational modification (Kolter & Moreno, 1992; Bierbaum et al., 1996). Such processing confers specific chemical properties and toxic activities that could not be otherwise achieved with a repertoire of 20 proteinogenic amino acids (Bierbaum et al., 1996). When combined with secretion into the extracellular medium, post-translational processing can also prevent the premature activation of the toxic properties of a bacteriocin, thereby protecting host cells from committing suicide (Garrido et al., 1988; Kolter & Moreno, 1992; Kupke & Götz, 1996; Peschei & Götz, 1996; Thumm & Götz, 1997). In E. coli, production of an antibacterial toxin (colicin) is controlled by the SOS system and stochastic factors resulting in colicin production by only a small fraction (0.5–3%) of their respective E. coli populations during stationary growth phase (Mulec et al., 2003; Cascales et al., 2007; Mrak et al., 2007; Kamenšek et al., 2010; Majeed et al., 2014). Once activated in the medium, bacteriocins have the remarkable property of finding and killing specific bacterial competitors while causing little or no harm to the host cell (Baba & Schneewind, 1996). In many strains, bacteriocin production is controlled in a cell-density dependent manner, using a secreted peptide-pheromone for quorum-sensing. The sensing of its own growth, which is likely to be comparable to that of related species, enables the producing organism to switch on bacteriocin production at times when competition for nutrients is likely to become more severe. Unlike other antimicrobials, the lethal activity of bacteriocins is often limited to members of the same species as the producer, suggesting a major role in competition with conspecifics (Riley et al., 2003; Be’er et al., 2009, 2010). Such ‘‘chemical weapons’’ are even used to attack sibling cells within the same colony (Heininger, 2001; González-Pastor et al., 2003; Ellermeier et al., 2006; Claverys & Håvarstein, 2007). In liquid cultures, the outcome of competition between E. coli that produce colicin and sensitive E. coli is frequency dependent (Chao & Levin, 1981); the colicinogenic bacteria are at an advantage only when fairly common (frequencies in excess of 2 x 10-2). However, in a soft agar matrix, a structured habitat, the colicinogenic bacteria have an advantage even when initially rare (frequencies as low as 10-6). In a liquid culture, bacteria exist as randomly distributed individuals and the killing of sensitive bacteria by the colicin augments the amount of resource available to the colicinogenic bacteria to an extent identical to that experienced by the surviving sensitive bacteria. On the other hand, the bacteria in a soft agar matrix exist as single-clone colonies. As the colicinogenic colonies release colicin, they kill neighboring sensitive bacteria and form an inhibition zone around themselves. By this action, they increase the concentration of resources around themselves and overcome their growth rate disadvantage. Based on these findings, Chao and Levin (1981) suggested that structured habitats are more favorable for the evolution of colicinogenic bacteria.

Another example of siblicide is ‘‘cannibalism’’ in Bacillus subtilis, where bacteria during the early stages of sporulation produce chemicals that kill some siblings, which become food for the surviving bacteria (González-Pastor et al., 2003; Ellermeier et al., 2006). Spore formation by Bacillus subtilis, which is triggered by nutrient limitation, is an elaborate developmental process that takes place over the course of 7 to 10 hours and involves the conversion of a growing cell into a dormant cell type (a spore) that can remain inert for many years. The decision to form a spore is a life-or-death one (Bassler & Losick, 2006). A B. subtilis cell could be at a considerable disadvantage if it commits to spore formation in response to what turns out to be a brief fluctuation in nutrient availability. To guard against this possibility, the bacterium deploys a system of cannibalism in which cells that have entered the sporulation pathway forestall the absolute commitment to spore formation (González-Pastor et al., 2003; González-Pastor, 2011). At the heart of the cannibalism system is the master regulator for sporulation, Spo0A, which is activated by phosphorylation via a phosphorelay that is subject to three positive feedback loops. In response to nutrient limitation, about half of the B. subtilis cells activate Spo0A and enter the pathway leading stochastically to competence or sporulation (Mirouze et al., 2011), and the other (Spo0A-inactive) cells do not. The ultimate decision to sporulate is, however, stochastic in that only a portion of the population sporulates even under optimal conditions (Chastanet et al., 2010; Chastanet & Losick, 2011; Narula et al., 2012). Cells that have activated Spo0A produce and export a killing factor and a protein toxin that together kill nonsporulating siblings. Their deaths result in the release of nutrients that, in turn, delay or reverse progression into sporulation by the cells that have activated Spo0A. When no siblings remain to be cannibalized and no other sources of nutrients become available, development progresses to the point that spore formation becomes irreversible. The response to the cannibalism toxin involves an intercellular chemical signaling system of unusual simplicity (Ellermeier et al., 2006). To avoid suicide, toxin-producing cells (i.e., Spo0A-expressing cells) simultaneously produce a membrane bound immunity protein that neutralizes the toxin in the membrane. It is likely that cells become cannibalistic after the commitment to sporulation. If cells, however, become cannibalistic for lower levels of phosphorylated Spo0A (that corresponds to the competence window), they can decide between competence and cannibalism (Schultz et al., 2009).

Another example of fratricide is the allolysis behavior of the pathogen Streptococcus pneumoniae (during the transition to competence) (Steinmoen et al., 2002, 2003; Guiral et al., 2005; Håvarstein et al., 2006; Claverys & Håvarstein, 2007; Eldholm et al., 2009; Trappetti et al., 2011). S. pneumoniae is known for its ability to enter into a state of genetic competence under conditions of high cell population density in response to a secreted signaling peptide. Analogous to the case of B. subtilis sporulation, only a fraction of the S. pneumonia cells in the population become competent in response to the peptide autoinducer. Those that do so elaborate a bacteriocin that causes the lysis of noncompetent cells in the population. Claverys and coworkers (Guiral et al., 2005; Håvarstein et al., 2006) report that the lysed cells release not only transforming DNA and nutrients but also pneumolysin and other factors important for virulence. Thus, rather than relying on self for secretion of virulence factors, S. pneumoniae kills some its relatives for this purpose which facilitates invasion of its host.

After the pioneering work in B. subtilis and S. pneumonia, siblicide has been demonstrated in various other bacterial species (Sedgley et al., 2009; Thomas et al., 2009; Hwang et al., 2011), suggesting that this may be a frequent phenomenon. Exopolymeric substances (EPS) form the extracellular matrix of biofilms (Lawrence et al., 1991) and comprise a wide variety of polysaccharides, proteins, glycoproteins, glycolipids and, in some cases, large amounts of extracellular DNA (eDNA). eDNA was first shown to be present in the extracellular matrix of biofilms formed by Pseudomonas aeruginosa and necessary for biofilm formation (Whitchurch et al., 2002), and is now widely recognized as a major constituent of the matrix (Flemming et al., 2007). The matrix functions as a permeability barrier to limit both the diffusion of beneficial nutrients away from the biofilm and prevent or slow the diffusion of harmful substances such as antibiotics and predatory cells of the immune system from accessing matrix-embedded cells (Costerton et al., 1999). Fratricide and the eDNA it provides may also have an important role in biofilm formation (Webb et al., 2003b; Thomas et al., 2009; Trappetti et al., 2011; Berg et al., 2012; Wei & Håvarstein, 2012) and is essential for efficient gene transfer between Pneumococci in biofilms (Wei & Håvarstein, 2012). 

5.4 Excursion: "Veil of ignorance" and fair lotteries

John Harsanyi (1953, 1955) and John Rawls (1971) both used the veil of ignorance thought experiment to study the problem of choosing between alternative social arrangements and the problem of social justice (Okasha, 2012). Parties to an original agreement to establish a society and laws to govern are assumed to be completely ignorant of the position they will occupy in that forthcoming society. The purpose of this ignorance is to ensure that any decisions how to distribute resources across different positions are not motivated by a desire to tailor society to benefit one's own specific circumstances. A just society establishes rules that individuals regard as fair from behind a veil of ignorance about their position within society (Frank, 2013).

When it comes to distributing non-divisible, scarce goods in social competitions, three distributive mechanisms seem particularly prominent: (i) Selection, i.e. allocation according to some relevant criterion; (ii) Auction, i.e. bestowing the good on the highest bidder; or (iii) Lottery, i.e. a random allocation (Saunders, 2008). Given appropriate selection is impossible when parties have equal claims (e.g. in clonal populations), a lottery is preferable to an auction because it excludes unjust influences. The fairness of lotteries (Sher, 1980) gives each individual an equal chance of obtaining the good in question, as a surrogate for their equal claim to the good (Saunders, 2008). Of course, not all goods can be literally cut in half, as illustrated by the Biblical story of Solomon – half a baby is no use to anyone, and the real mother would rather surrender her claim than have her child cut in half. Thus, there are indivisible goods that cannot be shared in any sense and for which any allocation is necessarily ‘winner takes all’ (Saunders, 2008).

Individuals consistently have a strong betrayal aversion (Bohnet et al., 2008) and strongly dislike unfair social outcomes (Saito, 2012; Gaudeul, 2013), but rather prefer playing risky but fair social lotteries (Gaudeul, 2013; López-Vargas, 2014). A biased outcome is more readily accepted when chosen by an unbiased random draw than by one that is biased (Bolton et al., 2005). As a measure of risk or inequity aversion, equal allocations are chosen either for insurance purposes or are due to impartial social preferences that value equality per se (Carlsson et al., 2005; Frignani & Ponti, 2008; Schildberg-Hörisch, 2010). In the inequity aversion model, individuals enforce social norms of fairness, thereby stabilizing cooperation and social cohesion (Fehr & Schmidt, 1999; Fehr & Fischbacher, 2004). Animals are risk-averse (Heininger, 2015) and individuals from cooperative species appear also to be inequity-averse (de Waal, 1996; Brosnan & de Waal, 2003, 2014; de Waal & Davis, 2003; Range et al., 2009; Brosnan et al., 2010a; Massen et al., 2012; Wascher & Bugnyar, 2013; but see Henrich, 2004; Bräuer et al., 2009; Silberberg et al., 2009; Horowitz, 2012). Brosnan (2011) argued that inequity aversion is a mechanism to promote successful long-term cooperative relationships amongst nonkin, a hypothesis that is supported by model simulations (Ahmed & Karlapalem, 2014).

The veil-of-ignorance concept has also surfaced in biology, in the context of meiosis, the cell division process by which sexually reproducing organisms halve their chromosome number: only one of each chromosome pair is passed to each gamete. Most of the time meiosis is ‘fair’, so that any particular gene has a 50% chance of making it into any gamete – a fact known as Mendel’s law of segregation. When meiosis is fair, randomization puts each allele behind a veil of ignorance with regard to its direct transmission (interests) in each progeny. Behind the veil, each part of the genome can increase its own success only by enhancing the total number of progeny and thus increasing the success of the group (Frank, 2013). Randomization of position levels individual opportunity and promotes group cohesion. Given randomization of individual success within the group, an individual increases success only by increasing the success of the group as a whole (Frank, 2013). Probably any fair process can be corrupted by selfish individuals that try to “corriger la fortune”. For example in many species, rogue genes can cheat Mendel’s law and get into more than their fair share of gametes. This is known as ‘meiotic drive’ or ‘segregation distortion’; the genes in question are called segregation-distorters (Zimmering et al., 1970; Okasha, 2012).

In addition to meiosis, various fair lottery processes determine the random allocation of the reproductive position in the forthcoming society of a multicellular organism or in a termite colony. Mammalian germline determination is a stochastic process. Cell populations in the embryo are not comprised of a single cellular entity, but instead display significant heterogeneity at the molecular level, heterogeneity that is associated with an apparent probabilistic element of fate determination (Enver et al., 2009; Hough et al., 2009). Phenotypic fluctuations may be a general feature of any non-terminally differentiated cell (Stockholm et al., 2007, 2010; Chang et al., 2008; Hayashi et al., 2008; Kalmar et al., 2009). The cells fluctuate slowly but continuously between different phenotypic states that leads to a dynamic equilibrium with relatively constant proportions of various phenotypic variants in the population. The cellular microenvironment created by the cells themselves contributes actively and continuously to the generation of fluctuations depending on their phenotype. As a result, the cell phenotype is determined by the joint action of the cell-intrinsic fluctuations and by collective cell-to-cell interactions (Stockholm et al., 2010). Pluripotency is the capacity of a single cell to generate in a flexible manner all cell lineages of the developing and adult organism. The master pluripotent regulator Nanog has a key role in safeguarding stem cell pluripotency against differentiation and mediates germline development (Chambers et al., 2007; Silva et al., 2009; Zhang & Wolynes, 2014). Nanog is expressed in pluripotent embryo cells, derivative embryonic stem cells, and the developing germline of mammals and birds (Chambers et al., 2003; Mitsui et al., 2003; Yamaguchi et al., 2005; Lavial et al., 2007). Murine embryonic stem cells display heterogeneity characterized by fluctuations between two clearly different phenotypic states of embryonic stem cells; one is stable (‘‘high Nanog’’; HN) and the other is unstable (‘‘low Nanog’’; LN). The transition between the HN to LN phenotype is stochastic and rare, whereas those from LN to HN are frequent. The observations are consistent with a model with excitable dynamics where the first change is rapid and noise-triggered followed by slow relaxation to the initial state (Kalmar et al., 2009). Likewise, in mouse embryonic stem cell cultures, a subset of cells is positive for both the pluripotency marker gene Oct-4, the undifferentiated state marker of embryonic stem cells Rex1, and a definitive marker of the germ cell lineage Stella, and these cell types and the marker-negative cells can interconvert (Hayashi et al., 2008; Toyooka et al., 2008).

Among the workers of a termite colony, only individuals in the sensitive period (a short period during the moulting interval when the developmental fate of an organism at the next moult is determined) are able to respond to orphaning of the colony and become neotenic replacement reproductives. As all workers pass through this period, they all have a fair chance of becoming neotenic, while at the same time the number of actually competing individuals (that fight to death to take over the colony) is reduced. Thus, the sensitive period, besides honest signalling (via cuticular hydrocarbon profiles and trophallaxis), functions as a ‘fair lottery’ mechanism (Hoffmann, 2011; Hoffmann & Korb, 2011).

Evolution has no foresight. The “veil of ignorance” thought experiment also applies to a reproducing organism concerning the evolutionary scenarios that its offspring will experience. In uncertain, unpredictable environments bet-hedging is the evolutionarily stable strategy (Heininger, 2015). Risk spreading in unpredictable environments confers equality of opportunity and thus ex-ante fairness of lottery (Saito, 2012). It is an intruiging aspect of lotteries that individuals that can gain more from lotteries (low-income people or cells/animals under vital stress) are more willing to participate in lotteries (Spiro, 1974; Brinner & Clotfelter, 1975; Suits, 1977; Clotfelter & Cook, 1987, 1989; Livernois, 1987; Hansen, 1995; Clotfelter et al., 1999; Hansen et al., 2000; Blalock et al., 2007; Haisley et al., 2008; Beckert & Lutter, 2013; Leach, 2013). In fact, at the transition of unicellularity to multicellularity, vitally endangered, starving, microbes are ready to try their luck in life-death lotteries. If they would not cooperate and gamble, the chances that all of them die are substantial. The curtain of ignorance appears to affect the behavior from vitally endangered cells to social decisions in humans. In the Prisoner’s Dilemma game, a paradigmatic example for studying the dilemmas between individual interests and collective welfare, many subjects compete when they know that the opponent has competed and when they know that the opponent has cooperated, but behind the curtain of ignorance they cooperate significantly more often (Shafir & Tversky, 1992). In the disjunctive case, when the other player’s strategy is not known, 65% of the subjects exhibited cooperation on at least one of the six Prisoner’s Dilemma triads that they played (Shafir & Tversky, 1992). Unpredictable environments shaped a multitude of evolutionary processes that engage fair randomization processes and downstream selection (either by a selective environment or internal milieu) such as RNA quasispecies (Eigen, 1996), somatic generation of antibody diversity (Jerne, 1955; Burnet, 1957; Tonegawa, 1976, 1983), sexual reproduction, cell differentiation, and cancerogenesis (Heininger, 2001, 2013, chapter 4.1.1).

5.4.1 Apoptosis as fair lottery with unequal outcome

If a number of persons engage in a series of fair bets, the distribution of cash after the last bet is fair, or at least not unfair, whatever this distribution is.
John Rawls (1971, p. 75 of revised edition 1999)

As noted in chapter 5.4, in the case of indivisible goods, the winner of a fair lottery “takes it all”. When resources are too scarce to ensure the survival of all members of a clonal society (particularly if the survivors have to undergo the energetically demanding metamorphosis from vegetative cells into metabolically dormant, resilient spores), the outcome of the competition for resources is decided by a fair lottery. Behind the “veil of ignorance”, cells competing for the indivisible good “life” are not “aware” of their relative position in the “commonwealth” of cells (Leigh, 1977).

The phenomenon that clonal cells display stochastic bistability has been attributed to cellular noise (Korobkova et al., 2004; Maamar et al., 2007; Leisner et al. 2008; Veening et al., 2008; Lopez et al., 2009). Connections between gene expression noise and cell state transitions have been demonstrated in many biological processes (Eldar & Elowitz, 2010). Bistability and the binary decision making it imparts have been widely observed and hypothesized as one of the possible mechanisms for cell fate determination (Biggar & Crabtree, 2001; Ferrell, 2002; Xiong & Ferrell, 2003; Wu et al., 2013). For example, competence/fratricide bistability is the result of stochastic cell fate determination (Leisner et al., 2008; Veening et al., 2008; Johnston et al., 2014, see chapter 5.3.2). In a gene network model, different levels of noise, designed to mimic degrees of ‘‘noisy’’ transcriptional activity in cellular systems, were found to either promote or disrupt state transitions, with some transitions requiring layovers at one or more intermediate states (Faucon et al., 2014). Cell-to-cell variation in genetically identical cells of multicellular organisms is often regulated by active non-genetic mechanisms (Kimble & Hirsh, 1979; Kimble, 1981; Doe & Goodman, 1985; Sternberg & Horvitz, 1986; Priess & Thomson, 1987; Jan & Jan, 1995; Karp & Greenwald, 2003; Hoang, 2004; Colman-Lerner et al., 2005). The two pathways resulting in death or survival may be mechanistically independent, and cell fate is determined by a stochastic kinetic competition between them that results in cell-to-cell variation (Huang et al., 2010). The specific molecular interactions and/or chemical conversions depicted as links in the conventional diagrams of cellular signal transduction and metabolic pathways are inherently probabilistic, ambiguous, and context-dependent (Kurakin, 2007). Regulatory systems or decisions, in which the outcome of a cellular event is at least partially the result of intrinsic noise, are said to be stochastic (Theise & Harris, 2006; Losick & Desplan, 2008; Eldar & Elowitz, 2010). Cell fate decisions are often controlled by both stochastic and deterministic features (Losick & Desplan, 2008; MacArthur et al., 2009; Snijder & Pelkmans, 2011). For instance, bacteria determine their fate by “playing dice with controlled odds” (Ben-Jacob & Schultz, 2010). Constrained randomness, intermediate between rigid determinism and complete disorder is what is usually seen (Theise & Harris, 2006). Specific environmental or genetic cues may bias the process, causing certain cellular fates to be more frequently chosen (as when tossing identically biased coins). Still, the outcome of cellular decision making for individual cells is a priori unknown (Balázsi et al., 2011).

Stochastic models of apoptosis make it possible to represent reactions as processes that are discrete and random, rather than continuous and deterministic. Stochastic models are advantageous when the number of individual reactants of any type is small (typically fewer than ~100) or reaction rates very slow (Zheng & Ross, 1991). In these cases, a Monte Carlo procedure is used to represent the probabilistic nature of collisions and reactions among individual molecules (Gillespie, 1977). Modeling and experimentation in bacteria, yeast, and mammalian cells, have provided a mechanistic framework for understanding stochastic variation (‘‘noise’’) in rates of transcription and translation (Raj & van Oudenaarden, 2008). The number of transcriptional initiation complexes on any single gene is small (potentially as small as 1–2), and the probability that a transcript will be created in any time interval is therefore highly stochastic. Fluctuations in mRNA levels result in fluctuating rates of protein synthesis. With short-lived or low-copy-number proteins, this can cause large fluctuations in protein levels, whereas with relatively abundant proteins, such as those controlling apoptosis, the most significant effect is that different cells contain different concentrations of each protein, and thus unique proteomes. Model-based simulation suggested that natural variation in the levels of apoptotic regulators is responsible for variability in the time and probability of cell death (Spencer et al., 2009; Spencer & Sorger, 2011). Stochastic behavior-dependent bistability results in different individual cells responding at somewhat different concentrations and time. When observed at a population level, the response is thus graded. Such contrast between population level and single cell level has been illustrated experimentally in a number of systems, including Xenopus levis oocytes (Bagowski et al., 2001, 2003; Pomerening, 2008; Ferrell et al., 2009), and oxidative stress induced apoptosis mediated by ERK pathway and p53 pathway (Nair et al., 2004).

Individual cells differ widely in their responses to apoptotic stimuli (Spencer & Sorger, 2011). Correlation in the probability of death and death time among sister cells has been observed in a variety of cell types following exposure to a variety of apoptosis-inducing agents. In contrast, randomly selected cells were found to be uncorrelated, and no obvious correlation with cell-cycle phase or with position in the dish could be detected (Bhola & Simon, 2009; Spencer et al., 2009). Importantly, the degree of similarity between sisters fell as the time since cell division increased so that within one to two generations, sisters were no more correlated than randomly chosen pairs of cells. This transient heritability in timing of death argues against a genetic or epigenetic explanation for cell-to-cell variability in apoptosis, as genetic and epigenetic differences tend to be stable over much longer timescales (Spencer & Sorger, 2011).

5.4.2 Are "self-destructive" acts the phenotype of fair lotteries?

Within the kin selection paradigm, apoptosis of unicellular and within multicellular organisms is often interpreted as altruistic suicide or self-sacrifice (Kondo, 1988; Alison & Sarraf, 1992; Allsopp & Fazakerley, 2000; Gardner & Kümmerli, 2008; Ackermann et al., 2008; Nedelcu et al., 2011). In D. discoideum cells killed by nonkin are called victims (Ho et al., 2013), but cells killed by kin are considered altruists that sacrifice themselves (Strassmann et al., 2000a). Ackermann et al. (2008) provided several examples of behavior of unicellular pathogens that was interpreted as self-destructive cooperation (Avery, 2006). Some bacterial toxins that are instrumental in pathogenesis can only be released if the cell producing the toxin lyses (Paton, 1996; Wagner et al., 2002; Voth & Ballard, 2005). Some of these toxins induce inflammation in the host, and there is now growing evidence that pathogens can decrease competition by co-inhabitants of the same niche through manipulation of the host’s immune system (Lysenko et al., 2005; Raberg et al., 2006; Brown SP et al., 2007; Stecher & Hardt, 2008). One example is pneumolysin from Streptococcus pneumoniae. This toxin is released by “self-destructive” bursting which promotes the invasion of the lung by cells that refrain from this extreme behavior (Paton, 1996; Ogunniyi et al., 2007). In the S. typhimurium enterocolitis model, most of the bacteria that invade the gut tissue and thereby contribute to the public good (inflammation as a proxy for the public good) seem to be killed by the intestinal innate immune defenses. Thus, “cooperation through invasion of the gut tissue is a largely self-destructive act” (Ackermann et al., 2008). Another example is TcdA, a key virulence factor of Clostridium difficile (Voth & Ballard, 2005). TcdA lacks a standard secretion signal and is released by bacterial lysis. Purified TcdA toxin alone can trigger gut inflammation (Lima et al., 1988), and gut inflammation enhances intestinal C. difficile colonization (Rodemann et al., 2007). In this case, TcdA released by “self-destructive” acts seems to provide the pathogen with a competitive advantage, presumably by decreasing competition from commensal bacteria. It has been noticed that from a genetic-determinism point of view, such self-destructive cooperation is very puzzling (Gardner & Kümmerli, 2008): a gene that leads all its carriers to make the ultimate sacrifice should become extinct in a single generation. However, these processes are the result of gene expression noise (Fraser & Kærn, 2009) that may lead to random cell fates (see chapter 5.4.1 and Heininger, 2015). Randomization of fate levels individual opportunity and promotes group cohesion. Given randomization of individual success within the group, an individual increases success only by increasing the success of the group as a whole (Frank, 2013). These random fates have been termed “coin-flipping altruism” by Cooper and Kaplan (1982). According to this logic every participant of a lottery, by buying a lottery ticket, would commit an act of altruism towards the eventual winner(s) of the lottery.

6. The ecological context-dependency of social interactions


The key remaining questions of evolutionary biology are more ecological than genetic in nature.
Edward O. Wilson, 1987

Summary
Ecological conditions play a major role in the evolution of sociality.Harsh or unpredictable environments, intense predator pressure, constraints on independent breeding, strong intra- or inter-specific competition, or resources that are difficult to acquire except as a group have been suggested to select for social living. Cooperative breeding is chiefly viewed as a ‘best-of-a-bad-job’ strategy, undertaken when constraints prevent independent reproduction, forcing some individuals to pursue the alternative strategy of helping others. Current theory suggests that mutualisms are best viewed as reciprocal exploitations that nonetheless provide net benefits to each partner This view stresses the disruptive potential of conflicts of interests among the erstwhile partners resulting in a dynamic and flexible balance between cooperation and competition. Cooperation is never here to stay.

Ecological conditions play a major role in the evolution of sociality (Rubenstein & Wrangham, 1986; Slobodchikoff, 1988; Krebs & Davies, 1993; Arnold & Owens, 1999; Foster & Xavier, 2007; Korb & Heinze, 2008; Westneat & Fox, 2010; Davies et al., 2012; Gordon, 2014; McAuliffe & Thornton, 2015). For instance, several ecological hypotheses, not mutually exclusive, have been proposed to explain the evolution of cooperative breeding. The most influential hypotheses can be classified according to the following scheme (Pen & Weissing, 2000):
(i) Ecological constraints hypothesis (Brown, 1974; Emlen, 1982a, 1997; Koenig et al., 1992).
(ii) Life-history hypothesis (Russell, 1989; Cockburn, 1996; Arnold & Owens, 1998).
(iii) Benefits-of-philopatry hypothesis (Stacey & Ligon, 1987, 1991).
The ecological constraints and life-history hypotheses both stress that the direct fitness benefits of seeking independent breeding opportunities are too small to outweigh the indirect inclusive fitness benefits of helping relatives. In contrast, the benefits-of-philopatry hypothesis emphasizes the long-term direct benefits of staying near the natal nest: short-term losses incurred by not dispersing are compensated by greater long-term direct benefits, in the form of inheritance of the natal territory (Pen & Weissing, 2000).

The life history strategies of cooperative breeders appear to be biased towards the K-selected end of the r-K continuum, being characterized by delayed maturity, high adult survival, and low reproductive and dispersal rates (Brown, 1974, 1987; Gaston, 1978; Russell, 1989; Rowley & Russell, 1990; Poiani & Jermiin, 1994; Arnold & Owens, 1998, 1999). Cooperative breeding seems to have a secondary role for the maintenance of delayed dispersal, although 96% of bird species where the offspring remain with their parents into adulthood to form family groups also breed cooperatively (Emlen, 1995). Cooperative breeding can be seen as an independent decision, and as such it is a consequence rather than a cause of delayed dispersal (Brown, 1987; Stacey & Ligon, 1987; Koenig et al., 1992; Emlen, 1994; Hatchwell & Komdeur, 2000), which is consistent with the observation that dispersal can be delayed without the retained offspring engaging in reproduction (Ekman et al., 2001a). Even if some of the retained offspring in a species participate in cooperatively breeding units, there are usually a substantial fraction of them that do not engage in help-at-the-nest (Ekman et al., 2001a). It has been hypothesized that several species that delay dispersal and associate in family groups but do not practise alloparenting, exercise prolonged brood care to raise offspring fitness (Ekman et al., 1994; Ekman, 2006). Phylogenetic analyses revealed two important characteristics of family living (Covas & Griesser, 2007). First, family living has a strong phylogenetic component, being unevenly distributed between families (or between genera in families that present both cooperative and pair breeders), and is the ancestral state in several lineages (Cockburn, 1996, 2003; Arnold & Owens, 1998). Second, family living occurs more frequently among long-lived bird species (Arnold & Owens, 1998; Covas & Griesser, 2007; Griesser & Barnaby, 2010). Life-history theory predicts that long-lived species should benefit from a delayed onset of reproduction (Goodman, 1974; Stearns, 1992; Charlesworth, 1994), and this has been supported by studies that demonstrated a positive effect of delayed onset of reproduction on lifetime reproductive success in long-lived bird species (Stacey & Ligon, 1987; Ekman et al., 1999; Krüger, 2005). Furthermore, longevity not only gives the option to offspring of delaying the onset of reproduction, but it also reduces the cost to parents of a prolonged investment in offspring (Ekman & Rosander, 1992). Extended parental investment is an important factor that has been suggested to facilitate family formation (Brown, 1987; Ekman et al., 2001a; Ekman, 2006). Clearly, extended brood care does increase the reproductive success of parents and survival of offspring, both direct fitness benefits. Australian brown thornbill (Acanthiza pusilla) juveniles that delayed dispersal were four times more likely to recruit than juveniles that dispersed early (Green & Cockburn, 2001). Philopatric Siberian jay (Perisoreus infaustus) offspring have an odds ratio of being killed by predators 62% lower than offspring that dispersed promptly after independence to join groups of unrelated individuals (20.6% versus 33.3% winter mortality). The higher survival rate among philopatric offspring was associated with parents providing nepotistic predator protection that was withheld from unrelated group members (Griesser et al., 2006). Ekman (2006) argued that a variety of family groups maintained in the absence of alloparental care underlines the capacity of general group living enhancing survival as a primary agent selecting for family cohesion. These seasonal constraints on fitness components selecting for family cohesion may contribute to the large scale geographical pattern with a relative paucity of family cohesions among bird species in the northern hemisphere. On the basis of an extensive literature survey, Russell (2000) and Russell et al. (2004) found that juveniles of tropical and southern temperate species stay on natal territories significantly longer than juveniles of northern temperate species. Juveniles in an estimated 40% of tropical–southern temperate species (excluding flocking and cooperatively breeding species) remain on natal territories for 3–10 months after fledging which means that in these species a significant proportion of the offspring’s first year of life is spent with its parents on natal territories (Gill & Stutchbury, 2010). Green and Cockburn (2001) concluded that delayed dispersal is not a strategy confined to species that breed cooperatively and suggested that prolonged philopatry in cooperative species in the Corvida are more likely to be driven by direct fitness benefits to offspring rather than indirect benefits accrued by raising non-descendent kin. It has been suggested that rather than regarding life history traits as predisposing and ecological factors as facilitating cooperation, they are more likely to act in concert (Hatchwell & Komdeur, 2000).

The fact that cooperative breeding should not be essential for delayed dispersal is consistent with the view that the behavior of remaining in the natal territory is maintained as a product of ecological constraints on dispersal options (Ekman et al., 2001a). A variety of conditions have been suggested to select for social-living (Avilés, 1999), including high risks involved with dispersal (Emlen, 1982a), lack of mates (Rowley, 1981, Pruett-Jones & Lewis, 1990), constraints on independent breeding (Emlen, 1991), intense predator pressure (Alexander, 1974; Wilson EO, 1975; Caraco & Pulliam, 1984; Stern & Foster, 1996), strong intra- or inter-specific competition (Wilson EO, 1975; Buss, 1981; Hogendoorn & Velthuis, 1993), harsh or unpredictable environments (e.g. Reyer, 1980; Emlen & Wrege, 1991; Jarvis et al., 1994; Covas et al., 2008), or resources that are difficult to acquire except as a group (Slobodchikoff, 1984; Raffa & Berryman, 1987; Wyatt, 2003; Whitehouse & Lubin, 2005; Platt et al., 2012). In theoretical models, harsher environments led to higher long-term frequencies of cooperators (Smaldino et al., 2013a, b), lending support to Kropotkin’s (1902) proposal that harsh environments should select for cooperation. In a comparative study of reproduction among southern sea lions (Otaria byronia) during a single breeding season, it was documented that only one of 143 pups born to gregarious group-living females died before the end of the season, compared to a 60 percent mortality rate among solitary mating pairs. The main reasons were that pups in colonies were protected from harassment and infanticide by subordinate males and were far less likely to become separated from their mothers and die of starvation (Campagna et al., 1992). Diverse selective pressures have contributed to the evolution of the varied social groups of carnivores: the benefits of strength of numbers for defense of kills and territory, and in the hunting and killing of large prey; the ability to intimidate predators and to be vigilant against their approaches; the potential for information transfer and social learning (Dukas, 1998), and a suite of alloparental behavior patterns (Macdonald, 1983). The various conditions, which have been collectively referred to as the ‘ecological constraints hypothesis’ (Brown, 1974; Emlen, 1982a, 1991; 1997; Brockmann, 1997) would result in deaths outpacing births and, thus, in unsustainable rates of growth for populations of solitary individuals attempting to colonize such environments.

Cooperative breeding is chiefly viewed as a ‘best-of-a-bad-job’ strategy, undertaken when constraints prevent independent reproduction, forcing some individuals to pursue the alternative strategy of helping others (Arnold & Owens, 1999; Dickinson & Hatchwell, 2004; Russell, 2004). The fitness benefits of helping are most apparent in harsh conditions (Reyer, 1980; Covas et al., 2008). Numerous studies have demonstrated the importance of territory quality, access to breeding sites, resource availability, and other ecological factors in influencing reproductive and dispersal decisions in cooperatively breeding species (e.g., Pruett-Jones & Lewis, 1990; Walters, 1990; Komdeur, 1992; Walters et al., 1992). In the cooperatively breeding Seychelles warbler, Acrocephalus sechellensis, transfers of warblers to unoccupied islands showed that both habitat saturation and variation in territory quality dramatically affected the frequency of delayed dispersal. At first there was no cooperative breeding, but as all high-quality areas became occupied, young birds hatched on them began to stay as helpers, rather than occupy breeding vacancies on lower quality territories. However, as the number of helpers on high-quality territories increased, it paid some helpers to leave, even to poor territories. Thereafter, young reared on poor territories did better to leave to breed on poor territories, rather than stay at home (Komdeur, 1992; Komdeur et al., 1995). Among 20 species of sponge-dwelling shrimp (Synalpheus), eusocial species, consistent with hypotheses that cooperative groups enjoy an advantage in challenging habitats, are more abundant, occupy more sponges and have broader host ranges than nonsocial sister species (Duffy & Macdonald, 2009). In addition to ecological and demographic constraints, the energetic costs of reproduction, specifically costly gestation and costly postnatal investment in litter growth limit the benefits of attempting to breed as a subordinate in communally breeding carnivores (Creel & Creel, 1991). In mole-rats, the degree of sociality increases as resource abundance decreases and variability in rainfall increases (Faulkes et al., 1997; O’Riain & Faulkes, 2008).

Environmental uncertainty and harshness play a key role in explaining the incidence and distribution of avian cooperative breeding behavior (Emlen, 1982a, 1991; du Plessis et al., 1995; Hatchwell & Komdeur, 2000; Hatchwell, 2007; Rubenstein & Lovette, 2007; Cockburn & Russell, 2011; Jetz & Rubenstein, 2011). Experimental studies in which constraints on independent breeding were relaxed resulted in helpers moving to adopt the vacant breeding opportunities (Pruett-Jones & Lewis, 1990; Komdeur, 1992; Walters et al., 1992). Seychelles warblers became independent breeders once they had been released into an unoccupied island containing many good breeding territories. However, as the new population increased and all the high-quality habitat became occupied, the young again began to remain at home where they acted as helpers, as they usually do in their native habitat in which breeding vacancies are rare (Komdeur, 1992). The highly eusocial allodapine bee species Exoneurella tridentata, appears to have evolved sociality in very harsh, xeric conditions (Dew et al., 2012), and in years with harsh weather conditions colonies of primitively eusocial sweat bee, Halictus ligatus, showed an increase in sociality, i.e. higher levels of queen–worker dimorphism and decreased worker cheating (Richards & Packer, 1996). In a similar vein, communality among the bees may also be associated with more arid environments (Australia, Southwest USA) possibly because of restricted opportunities for fossorial nesting (Wcislo & Tierney, 2009). In some ant, termite, and social spider species, rainfall is positively correlated with sociality (Riechert et al., 1986; Murphy & Breed, 2007; Picker et al., 2007; Purcell, 2011).

Several authors argue that humans can be considered cooperative breeders because, although parenting behavior is highly culturally variable, in no culture do mothers raise their children without help from others (Hrdy, 1999, 2009; Mace, 2000; Sear & Mace, 2008; Hill & Hurtado, 2009; Smaldino et al., 2013b). Hill and Hurtado (2009), in their studies of contemporary South American hunter-gatherer societies, found not only that cooperative breeding behavior was ubiquitous, but also, crucially, that husband-wife pairs were physically incapable of procuring sufficient food for their offspring and themselves without help from others (Smaldino et al., 2013b). Moreover, they found that meat acquisition of Ache hunters over a given 90-day period was often highly variable for any given individual, as a result of illness, injury, or luck. Sharing food resources between nuclear families was therefore necessary to ensure the survival of young children (Smaldino et al., 2013b). It is likely that hominins have been raising their children cooperatively for some time (Hrdy, 2009). The large brains of humans inevitably make our offspring costly to raise because the growth and development of brain tissue requires high levels of energy and nutrients (Charnov & Berrigan, 1993). Isler and van Schaik (2009) have argued that the increased encephalization of the hominin line would not have been possible unless females were receiving help provisioning their young, particularly during the heightened habitat instability in the East African Rift System at the beginning of the Pleistocene (Trauth et al., 2005; Bobe & Leakey, 2009; Potts, 2013). Cooperative breeding has been suggested as a potentially crucial factor in the evolution of human prosociality and our tremendous cognitive advantage over our nearest relatives, the great apes (Burkhart et al., 2009; Hrdy, 2009).

6.1 The dynamics of cooperation and competition

As self-organized systems (see chapter 16), social units are dynamic. The production of structures as well as their persistence requires permanent interactions between the members of the colony and with their environment. These interactions promote the positive feedbacks that create the collective structures and act for their subsistence against negative feedbacks that tend to eliminate them (Garnier et al., 2007). Ongoing interactions in social systems may be a real-life manifestation of cooperative coaction that in models evolves more readily than reciprocal cooperation (van Doorn et al., 2014). Current theory suggests that mutualisms are best viewed as reciprocal exploitations that nonetheless provide net benefits to each partner (Nowak et al., 1994; Leigh & Rowell, 1995; Maynard Smith & Szathmáry, 1995; Herre & West, 1997; Doebeli & Knowlton, 1998; Herre et al., 1999; Sachs et al., 2004; Foster & Wenseleers, 2006; Sachs, 2006, 2013; Leigh, 2010; Jones et al., 2012). This view stresses the disruptive potential of conflicts of interests among the erstwhile partners. There is a dynamic and flexible balance between cooperation and competition (Nonacs & Reeve, 1995). In models, the unstable balance results in the oscillatory nature of cooperation (Nowak & Sigmund, 1989, 1993a; Imhof et al., 2005, van Doorn et al., 2014). Cooperation is never here to stay. Instead, there are endless cycles between all-out defectors, harsh retaliators, careful forgivers and unconditional cooperators (Imhof et al., 2005; Imhof & Nowak, 2010). Even eusociality can be lost or suppressed under appropriate ecological circumstances (Wcislo & Danforth, 1997; Danforth et al., 2003). Competition for resources is a common feature of mutualisms (Holland et al., 2005; Holland & 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 & DeAngelis, 2009, 2010). Across a dozen genera, queens able to found a colony alone often join unrelated queens, thereby enhancing worker production and colony survivorship. The benefits of joining other queens vary with group size and ecological conditions. However, after the first workers mature, the queens fight until only one survives (Bernasconi & Strassmann, 1999).

Phylogenies reveal that parasites as well as autonomous (non-mutualist) taxa are nested within ancestrally mutualistic clades. Close scrutiny of mutualistic interactions reveals that heterogeneous selection pressures (e.g. adaptation to local habitats, drift and low gene flow) can alter the strength and net fitness effect, causing these interactions to become antagonistic or parasitic under certain conditions (Johnson et al., 1997; Hochberg et al., 2000; Hochberg & van Baalen, 2000; Bever, 2002; Thompson & Cunningham, 2002; Sachs & Wilcox, 2006; Thrall et al., 2006; Palmer et al., 2008). Although models have focused on the propensity of mutualism to become parasitic, such shifts appear relatively rarely (Sachs & Simms, 2006; Sachs et al., 2011a). By contrast, diverse systems exhibit reversions to autonomy, and this might be a common endpoint to mutualism (Nishiguchi & Nair, 2003; O’Brien et al., 2005; Sachs & Simms, 2006; Mueller et al., 2010; Sachs et al., 2010, 2011b; Kikuchi et al., 2011; Sachs, 2013). Mutualisms and commensalisms dissolve when one party ceases to benefit from, or becomes less dependent on, the other. 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 & Jansen, 2001; Jones et al., 2012). Changing resource supply rates influence mutualistic interactions (Schwartz & Hoeksema, 1998). The energy flow through an ecosystem (or productivity) varies across landscapes (Rosenzweig, 1995) and is believed to play a major role in determining coevolutionary dynamics (Rosenzweig, 1995; Hochberg & van Baalen, 1998, 2000; Thompson, 2005; Lopez-Pascua & Buckling, 2008). The benefits derived from resource-trading depend strongly on the nature of the trade-off between the acquisition of one resource and the acquisition of another, described by the shape (linear, convex or concave) of the resource acquisition constraints of the individuals involved. The benefit derived from resource exchange depends on three factors: (i) relative differences between the partners in their resource acquisition abilities; (ii) relative differences between the partners in their resource requirements; and (iii) variation in the shape of resource acquisition trade-offs (Hoeksema & Schwartz, 2003). These models provide a suite of predictions about whether or not resource exchange is beneficial for two heterospecific individuals relative to a strategy of non-interaction (Schwartz & Hoeksema, 1998; Hoeksema & Schwartz, 2003). Long-term mutualisms dissolve if one party ceases to need, or to benefit from, the other’s services (Leigh, 2010). Corals expel zooxanthellae when they can no longer provide carbohydrates (Trench, 1997; Baker, 2001). The saxifrage herb Lithophragma preferentially aborts flowers full of eggs of the flower-parasitizing pollinator moth Greya when cheaper pollinators are available (Thompson & Cunningham, 2002). Interactions between African acacias and their ant defenders become more aggressively antagonistic if large herbivores, whom the ants deter, are removed (Palmer et al., 2008). Live-in nematodes compromise the fitness of their fig-wasp host if the fig fruit their host enters is likely to harbour other fig wasps whose young can be colonized by these nematodes’ young (Herre, 1993).

7. Cooperation and competition: threshold traits on a continuum of ecological variables


Summary
In complex nonlinear relationships changes in effects can be disproportionate to the changes in the causal element(s). 
Many traits are phenotypically discrete but may be polygenically or polyetiologically determined. Such traits can be understood using the threshold model of quantitative genetics that posits a continuously distributed underlying trait, called the liability, and a threshold of response: individuals above the threshold display one morph and individuals below the threshold display the alternate morph.
Theoretical work found that kin competition can strongly antagonize the benefits of kin cooperation and inhibit the evolution of cooperation in viscous populations.
Conflicts may have both beneficial and disastrous consequences. Conflict is likely to be a major driver of evolutionary change within and between species. Cheater’s payoff is frequency-dependent. Cheating or defecting strategies may do very well as long as they are rare in a population of cooperative individuals but fare worse or even perish when surrounded only by other cheaters. The anthropocentric and moralizing concepts of kin selection theory may bias the interpretation of behavior deviant to expectation. Phenotypes of dominance hierarchies, division of labor or bet-hedging strategies may be mistaken as “cheating”.

Typically, animals first respond behaviorally to changes and challenges in their environment (Tregenza, 1995; Kappeler et al., 2013), whereas adaptations of their morphology, physiology and life history take much longer (Releya, 2002). In fact, maximal behavioral flexibility might be generally advantageous in various natural situations (de Witt et al., 1998; Wright et al., 2010), including adaptations to climate change (Parmesan, 2006). Thus, evolution places a premium on behavioral plasticity or flexibility (Thorpe, 1974; Wilson, 1978; Kappeler et al., 2013). Variation in social systems occurs in many species, either between or within populations (Lott, 1984, 1991; Schradin, 2013). Interactions between organisms exist along a continuum from mutualism and commensalism to parasitism (Starr, 1975; Lewis, 1985; Ewald, 1987; Hochberg et al., 2000; Moran & Wernegreen, 2000; Thompson & Cunningham, 2002; Neuhauser & Fargione, 2004; McCreadie et al., 2005; Moran, 2007; Leung & Poulin, 2008; Pérez-Brocal et al., 2013; Heath & Stinchcombe, 2014). Variation in environmental conditions or the nature of density-dependent interactions along environmental gradients may be important in generating dynamic patterns of competition and cooperation in many natural populations (Buss, 1981; Branch & Barkai, 1988; Bertness, 1989; Bertness & Yeh, 1994; Holmgren et al., 1997; Tielbörger & Kadmon, 2000; Amar et al., 2008; Hudson & Trillmich, 2008; Roulin & Dreiss, 2012; Wong et al., 2013, 2014). Many traits are phenotypically discrete but may be polygenically or polyetiologically determined. Such traits can be understood using the threshold model of quantitative genetics that posits a continuously distributed underlying trait, called the liability, and a threshold of response: individuals above the threshold display one morph and individuals below the threshold display the alternate morph. In complex nonlinear relationships (see chapter 16) changes in effects can be disproportionate to the changes in the causal element(s). This may be a threshold effect so that change in the effect is proportionate to change in causal element(s) until a particular point is reached when the change becomes disproportionate (Byrne & Callaghan, 2014). The exploration of such effects is the domain of catastrophe theory. Semelparity and iteroparity (Roff, 1996, 1998; Lesica & Young, 2005; Heininger, 2012), and sexual and asexual reproduction (Heininger, 2013) are threshold traits. The discontinuous variation depends on both the genotype and the environment. The plasticity of morphs suggests that they are opposite ends of a continuum of variation rather than representing a simple dichotomy. Likewise, empirical data suggest that “altruism” and selfishness are rather plastic phenotypic expressions of a single genotype (Yakubu, 2012, 2013).

Intra- and interspecific competition has the inherent potential to reduce diversity via exclusion of inferior competitors (Grime, 1973; Armstrong & McGehee, 1980; Passarge et al., 2006). Full symmetry of competition may be evolutionarily unstable in populations of related individuals as it may increase the probability of extinction due to demographic stochasticity (Aikio & Pakkasmaa, 2003). The theory of heterogeneous advantage predicts that competition intensifies when genetic diversity is low and, therefore, diametrically opposes the predictions of kin association (Griffiths & Armstrong, 2001). Several mechanisms may lead to heterogeneous advantage (Griffiths & Armstrong, 2001): (i) If different genotypes have different ecological needs they may use a homogeneous resource in different ways (Young, 1981). (ii) Mixtures may exploit a spatially heterogeneous environment more fully than homogeneous groups (Bell, 1985). (iii) In temporally heterogeneous habitats, mixtures may be more likely than homogeneous groups to produce genotypes that are better suited to the environment (Williams, 1975). (iv) Mixtures may be more resistant to pathogens because there should be a greater chance of there being a resistant genotype present (Wolfe, 1985). Evidence for heterogeneous advantage has come from laboratory studies in animals and plants (Kearsey, 1965; Caligari, 1980; Pérez-Tomé & Toro, 1982; Ellstrand & Antonovics, 1985; Fowler & Partridge, 1986; Martin et al., 1988; Jasienski et al., 1988; Kelley, 1989; López-Suárez et al., 1993). Heterogeneous advantage outweighed the benefits of kin association for juvenile salmon in a natural habitat both at an individual level, as shown by the differences in the condition indices, and at a population level as shown by the differences in density (Griffiths & Armstrong, 2001). Studies of the effects of kinship on larval amphibian growth found no significant differences between sibling and mixed treatments with respect to mass or variation in mass, suggesting that the effects of kin selection and genetic similarity between competitors may in some cases cancel each other out (Twomey et al., 2008). On the other hand, when siblings are likely to interact, genetic variation among individuals can decrease competition for resources and generate substantial fitness benefits within a single generation (Aguirre & Marshall, 2012). In ecological communities and theoretical models, mutualism/facilitation mediates competition and increases biodiversity and stability in ecosystems (Hacker & Gaines, 1997; Pachepsky et al., 2002; Schmitt & Holbrook, 2003; Bascompte et al., 2006; Guimarães et al., 2007; Bastolla et al., 2009; Thébault & Fontaine, 2010; Cain et al., 2011; Mittelbach, 2012; McIntire & Fajardo, 2014).

By helping kin/neighbors to produce more offspring, the intensity of competition experienced by the focal individual’s offspring and that of its neighbors is increased. Helping relatives/neighbors thus leads to local crowding and an increase in local competition, here understood as the extent to which an actor and a recipient (or their offspring) are more likely to compete against each other for the same resources than are two adult individuals (or offspring) sampled at random from the population (Grafen, 1984; Wade, 1985; Kelly, 1992, 1994a; Taylor, 1992; Wilson et al., 1992; Queller, 1994a; West et al., 2002a; El Mouden & Gardner, 2008; Grafen & Archetti, 2008; Lehmann & Rousset, 2010; Van Dyken, 2010). Evidence that resource-based helping creates negative ecological feedback already exists for survival or fecundity restraint (Wilson et al., 1992; van Baalen & Rand, 1998; Mitteldorf & Wilson, 2000; Le Galliard et al., 2003; Werfel & Bar-Yam, 2004; Hauert et al., 2006; Alizon & Taylor, 2008; Lion & Gandon, 2009, 2010; Van Dyken & Wade, 2012). For example, the level of fighting between males of fig wasp taxa shows no correlation with the estimated relatedness of interacting males, but is negatively correlated with the number of females (future mating opportunities) (West et al., 2001). Thus, the benefits of kin-cooperation can potentially be negated by kin-competition (Grafen, 1984; Murray & Gerrard, 1984; Wilson et al., 1992; Taylor, 1992a, b; Queller, 1992, 1994; Kelly, 1994; Frank, 1998; West et al., 2001, 2002a; Gardner et al., 2004; Queller, 2004; El Mouden & Gardner, 2008; Grafen & Archetti, 2008; Platt & Bever, 2009; Lehmann & Rousset, 2010; Van Dyken & Wade, 2012). This tends to inhibit the evolution of helping. The cooperation-negating threat of kin competition has heightened the sense that true “altruism” in nature poses a serious conceptual dilemma. In particular, a number of models have shown that increased kinship “altruism” is exactly balanced by increased kin competition, making it impossible for true “altruism” to evolve (Charlesworth, 1979; Taylor, 1992; Wilson et al., 1992; Gardner & West, 2006). As stated by Kümmerli et al. (2009a), “relatedness and the scale of competition … will not usually be independent”.

Competition among relatives often emerges because kin are in close spatial proximity and depend upon the same limited resources (Stockley & Bro-Jorgensen, 2011). This problem was pointed out by Alexander (1974) and West-Eberhard (1975) many years ago, suggesting that an individual’s closest relatives, and by extension his/her closest associates and/or social allies, are often also his/her closest competitors (Smith, 2014). Competition among kin can reduce, or even negate, the kin-selected indirect benefits of altruism directed towards relatives. Interestingly, in such contexts, the direct benefits gained from outcompeting relatives through forces such as sibling rivalry and parent conflict generally appear to overwhelm the indirect benefits of social tolerance among kin (Trivers, 1974; Mock & Forbes, 1992; Johnstone, 2000; Cant, 2006). In some species, rates of conflict actually increase with levels of genetic relatedness. This is the case for rhesus macaques (Bernstein & Ehardt, 1986), ringtailed lemurs (Kappeler, 1993), African elephants, Loxodonta africana (Archie et al., 2006a), and yellow-bellied marmots, Marmota flaviventris (Smith et al., 2013). For example, as predicted by kin selection, rates of affiliation are positively correlated with genetic relatedness in yellow-bellied marmots as pups, yearlings and adults, but rates of conflict are also highest among the closest relatives at all three stages (Smith et al., 2013).

In an apparently wide class of models the cooperation-enhancing effect of limited dispersal is balanced by the competition-enhancing effect of limited dispersal (Taylor, 1992a, b; Wilson et al., 1992; Queller, 1994a; West et al., 2002a; El Mouden & Gardner, 2008; Grafen & Archetti, 2008). Whereas cooperative individuals are more likely to benefit kin in viscous populations, they also compete for limiting resources with these same kin (Platt & Bever, 2009). Early theoretical work found that such kin competition can strongly antagonize the benefits of kin cooperation and inhibit the evolution of cooperation in viscous populations (Grafen, 1984; Queller, 1992a, 1994a; Taylor, 1992a, b; Wilson et al., 1992). Consistent with this, empirical studies have failed to find a relationship between relatedness and aggressiveness in colonies of the multiple-queen wasp, Parachartergus colobopterus (Strassmann et al., 1997), fig wasps (West et al., 2001), and bruchid beetle larvae (Smallegange & Tregenza, 2008), suggesting that the effects of kin competition might negate any kin-selected benefits associated with being less aggressive toward kin.

In plants, competitive interactions may occur among the siblings that germinate from the seeds produced by a maternal parent (Willson et al., 1987; Cheplick, 1992, 1993a, 1993b; Kelly, 1996; Cheplick & Kane, 2004). Whenever the dispersal system results in relatives that are spatially aggregated, local competition becomes kin (sibling) competition (Lambin et al., 2001). Evolutionary consequences of such neighbor interactions will depend on whether there is variation among genetically related groups (i.e., families) as well as the relatedness of competing individuals (Nakamura, 1980; Wilson, 1987; Donohue, 2003). Sibling competition has been proposed as a selective force that could account for the evolutionary advantages of sexual reproduction—the hypothesis is that the genetically variable offspring produced will experience less severe competition than genetically similar or identical offspring (Maynard Smith, 1978; Bulmer, 1980; Barton & Post, 1986; Cheplick, 1992). This is because more diverse offspring are predicted to show greater ability to partition limiting resources (Young, 1981; Argyres & Schmitt, 1992). By analogous reasoning, outcrossing breeding systems may be selected for as a way to minimize the negative fitness consequences of sibling competition (Schmitt & Ehrhardt, 1987; McCall et al., 1989). In species with self-fertilizing or cleistogamous breeding systems, individuals are especially likely to be competing with close relatives (Cheplick, 1993b, 1996; Stevens et al., 1995; Cheplick & Kane, 2004). Data exist for agronomically important species and the cleistogamous summer annual Triplasis purpurea that reveal greater overall yield whenever different (as opposed to identical) genotypes are grown in competition (Allard & Adams, 1969; Harper, 1977; Price & Waser, 1982; Turkington, 1996; Cheplick & Kane, 2004). On the other hand, there have been a few reports showing greater growth or reproduction of individuals when competing with genetic relatives (compared with unrelated plants) in some species (Willson et al., 1987; Tonsor, 1989; Andalo et al., 2001; Donohue, 2003). The balance between negative and positive interactions (see chapter 10.4) has been shown to shift along environmental gradients (Tielbörger & Kadmon, 2000), with competition prevailing under environmentally benign conditions and positive interactions dominating under more harsh conditions (Bertness & Callaway, 1994; Bertness & Hacker, 1994; Bertness & Leonard, 1997; Callaway & Walker, 1997; Brooker & Callaghan, 1998; Callaway et al., 2002).

7.1 Conflict as a driver of evolutionary innovation

More than 40 years ago, Van Valen’s Red Queen hypothesis (1973) emphasized the primacy of biotic conflict over abiotic forces in driving selection. According to the Red Queen hypothesis, each adaptation by a species is matched by counteracting adaptations in another interacting species, such that perpetual evolutionary change is required for existence. Despite continued evolution, average relative fitness remains constant: evolution is a zero-sum game (Brockhurst et al., 2014). Thus, for a vast number of biological situations, the salient aspects of the selective environment are biotic conflicts (Venditti et al., 2010; Ezard et al., 2011; Liow et al., 2011; Brockhurst et al., 2014). Conflicts may have both beneficial and disastrous consequences. Conflicts can have a variety of outcomes including co-existence, cooperation, specialization, diversification, niche shifts, speciation, and extinction (MacArthur & Levins, 1964; Tilman, 1982; Helling et al., 1987; Schluter, 1994, 2000, 2001, 2010; Grover, 1997; Rainey & Travisano, 1998; Dieckmann & Doebeli, 1999; Bolnick, 2001, 2004; Pfennig & Pfennig, 2005; Hall & Colegrave, 2007; Svanbäck & Bolnick, 2007; M’Gonigle et al., 2009; Heininger, 2012). The fitness-boosting effects of conflict in coevolutionary systems has been demonstrated at various levels of biological organization (Spitze, 1991; Spitze et al., 1991; Clarke et al., 1994; Lynch & Spitze, 1994; Reznick et al., 2004; Fisk et al., 2007; Pal et al., 2007; Paterson et al., 2010). A growing body of work directly identified parasites and other natural enemies as key contributing factors in driving host/prey diversification (Bohannan & Lenski, 2000; Schluter, 2000; Buckling & Rainey, 2002; Brockhurst et al., 2004, 2005; Morgan & Buckling, 2004; Vamosi, 2005; Nosil & Crespi, 2006; Meyer & Kassen, 2007; Benmayor et al., 2008). Conflicts of interest between cooperators and cheaters maximize population fitness under co-existence (MacLean et al., 2010a). Likewise, the conflict between more or less cooperative partners is central to the maintenance of partner choice and cooperation itself in mutualisms (Foster & Kokko, 2006).

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. There have been several studies suggesting that intraspecific competition (Pacala & Roughgarden, 1985; Taper & Case, 1992; Schluter, 1996, 2000, 2010), predation (Naisbit et al., 2001; Vamosi & Schluter, 2002; Nosil, 2004; Nosil & Crespi, 2006), or infectious agents (Buckling & Rainey, 2002; MacColl, 2009; MacColl & Chapman, 2010; Schluter, 2010; Karvonen & Seehausen, 2012; Loker, 2012) lead to divergent selection. Mathematical theory has also indicated that, given the right conditions, divergence is expected to occur readily between competing species by a coevolutionary sequence of reciprocal changes in the traits used to consume resources (Taper & Case, 1985; Abrams, 1986; Doebeli, 1996).

Mutualism is characterized by competition for resources between the partners (Holland et al., 2005; Holland & DeAngelis, 2010; Jones et al., 2012) and increases biodiversity and stability in ecosystems (Pachepsky et al., 2002; Bastolla et al., 2009). Mutualism promotes coexistences of two species. Moreover, mutualism often can increase the carrying capacities of both species, and then promotes their competitive abilities (Zhang, 2003). An inferior competitor, if cooperative to a superior competitor, is also able to survive (Zhang, 2003). On the other hand, parasitic exploitation can generate substantial genetic and phenotypic polymorphism within species and may also be an important factor causing reproductive isolation and promoting speciation (Summers et al., 2003). When the bacterium Pseudomonas fluorescens and its viral parasite, phage Φ2 coevolved with each other, the rate of molecular evolution in the phage was much 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., 2004). 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 & Rowe, 1995, 2005; Chapman & Partridge, 1996; Rice, 1996, 1998a, b; Alexander et al., 1997; Parker & Partridge, 1998; Holland & Rice, 1999; Arnqvist et al., 2000; Gavrilets, 2000, 2004; Martin & Hosken, 2003; Chapman, 2006) and, indeed, sexual conflict seems to be a key ‘‘engine of speciation’’ (Arnqvist et al., 2000; Martin & Hosken, 2003; Gavrilets, 2004). On the other hand, when selection differs between the e.g. 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 & Partridge, 1998; Arnqvist et al., 2000; Kisdi et al., 2001; Johansson, 2008; Heininger, 2012, 2013).

7.2 The ecology of "cheaters"

Sociobiology seeks to explain social behavior as a product of natural selection (Wilson EO, 1975). Natural selection acts on both cooperative and competitive phenotypes, across multiple scales of biological organization. In the Darwinian tradition it has not been challenging to explain behaviors that are competitive, but it is regarded as a conundrum how cooperation can be stable in the face of selfishness and cheating (Wilson EO, 1975; Keller, 1999; Foster, 2004; Travisano & Velicer, 2004; West et al., 2006; Foster et al., 2007). “Cheaters” may be selected for (Porter & Simms, 2014). By its nature, cooperation can be exploited by selfish individuals, meaning, firstly, that selfish individuals derive an advantage from exploitation which is greater than the average advantage that accrues to unselfish individuals. Secondly, exploitation has no intrinsic fitness value except in the presence of the “cooperative behavior” (Koeslag & Terblanche, 2003). Exploiting the behavioral efforts of others is particularly widespread. It is known as ‘‘freeloading’’ in economics, “tolerated theft,” and “food-sharing” in anthropology and under various names such as ‘‘joining,’’ ‘‘kleptoparasitism,’’ and ‘‘scrounging’’ (Barnard & Sibly, 1981; Brockmann & Barnard, 1979; Giraldeau & Caraco, 2000) in behavioral ecology.

Exploiters save the time and energy that mutualists spend on reciprocating. “Cheaters” are widespread in unicellular and facultatively multicellular organisms (Hilson et al., 1994; Dao et al., 2000; Pál & Papp, 2000; Strassmann et al., 2000a; Velicer et al., 2000, 2002; Ennis et al., 2003; Fiegna & Velicer, 2003; Rainey & Rainey, 2003; Castillo et al., 2005; Rankin et al., 2007; Kuzdzal-Fick et al., 2011). “Cheating” is also rampant in most mutualisms (Poulin & Grutter, 1996; Johnson et al., 1997; Foster & Delay, 1998; Irwin & Brody, 1998; Addicott & Bao, 1999; Currie et al., 1999; Ferriere et al., 2002); in some cases, “cheaters” have been associated with mutualisms over long spans of evolutionary time (Addicott, 1985, 1996; Machado et al., 1996; Pellmyr et al., 1996; Després & Jaeger, 1999; Pellmyr & Leebens-Mack, 1999; Bronstein, 2001). Exploiters can evolve from former mutualists that cease service provisioning and then represent “cheaters”, or they can invade the mutualism starting from an originally independent lifestyle and then represent parasites of the mutualism (Bronstein, 2001; Segraves et al., 2005, Kautz et al., 2009; Orona-Tamayo & Heil, 2013). Exploitation has been reported for a wide range of mutualisms, including nectar robbing (e.g., bees and birds) (Roubik, 1982; Maloof & Inouye, 2000), domatia exploiters (Gaume et al., 2006, Shenoy & Borges, 2008), mycorrhizal fungi that uptake the plant carbon but transfer no nutrients to the plant (Smith et al., 1996), and strains of Rhizobium and Bradyrhizobium that transfer less or no nitrogen to the host than mutualistic strains (Wilkinson et al., 1996). On the other hand, non-photosynthetic orchids may be “cheaters” because they obtain nutrients from mycorrhizal fungi without providing carbon in return (Taylor & Bruns, 1997). Within the conceptual framework of a continuous gradation between mutualism, commensalism, and parasitism, host resistance and immunological responses against parasitic exploiters can be viewed as analogous to the “policing” of mutualists/symbionts (Leung & Poulin, 2008).

Public goods are those produced by an individual and shared with other group members (Driscoll & Pepper, 2010; Powers et al., 2011). They typically increase the fitness of all group members, but at a unilateral cost to the producer. Examples of such public goods production are widespread in nature and include the production of extra-cellular substances by microbes (Griffin et al., 2004; Gore et al., 2009), the sharing of information by an individual with the rest of its group, as occurs during predator inspection by guppies (Dugatkin, 1990), and alarm calls in birds and mammals (Charnov & Krebs, 1975; Hollén & Radford, 2009). The production of such goods is a type of cooperative behavior (West et al., 2007c) that is vulnerable to exploitation by “cheating” nonproducers, which reap the benefits of the public goods without contributing to them. The exploitation of hosts by opportunistic pathogenic bacteria, such as Bacillus thuringiensis, involves sharing the exploits of toxin production from multiple individuals as single individuals are incapable of overcoming host defenses. The exploitation of toxin producers by cheats (non-toxin producing strains) has consequences for pathogen virulence (Raymond et al., 2007, 2009), and host-pathogen epidemiology (Bonsall, 2010). This can lead to a “Tragedy of the Commons” (Hardin, 1968; Rankin et al., 2007), in which cheating nonproducers increase in frequency, even though this leads to a decline in mean fitness (Powers et al., 2011).

Despite exploitation of common goods by naturally arising cheaters, however, cooperation persists (Naumov et al., 1996; Strassmann et al., 2000a; Schaber et al., 2004; Dobata & Tsuji, 2009; Vos & Velicer, 2009; Wilder et al., 2009). Cooperators have evolved various processes to curb cheaters (Waite & Shou, 2012): direct and indirect reciprocity, sanctions, partner choice and fidelity (Weyl et al., 2010; Archetti et al., 2011; Nowak & Highfield, 2011; Frederickson, 2013), pleiotropy, privatization of common goods, diminishing returns, positive assortment, adaptive race. In the following, I adhere to the categorization of anti-cheater mechanisms presented by Waite and Shou (2012):
(i) Genes required for cooperation can have pleiotropic effects, such that a cell defective in paying the cost of cooperation is also incapable of enjoying the cooperative benefit (Foster et al., 2004; Banin et al., 2005; Xavier & Foster, 2007; Harrison & Buckling, 2009; Dandekar et al., 2012). For example, in the social amoeba Dictyostelium discodium, a gene encoding the receptor necessary for differentiation into stalk cells is also necessary for proper spore formation; thus, cheaters trying to avoid the stalk fate cannot become spores (Foster et al., 2004).
(ii) The budding yeast Saccharomyces cerevisiae secretes invertase to hydrolyze the disaccharide sucrose into glucose and fructose, which can be metabolized more efficiently. These monosaccharides were initially thought to be strictly common goods (Greig & Travisano, 2004), although it was later found that ∼1% are retained by the producing cell (Gore et al., 2009). Even such a seemingly insignificant level of privatization can allow cooperators to invade a population of cheaters (Doebeli & Hauert, 2005; Gore et al., 2009). This could explain the coexistence of invertase-producing cooperative cells with nonproducing cheating cells in wild populations (Naumov et al., 1996).
(iii) When individuals interact through the production and/or consumption of inexpensive common goods in randomly formed groups that assemble and disassemble cyclically, as long as an increase in the availability of the common good leads to a less than proportional increase in the fitness of its consumers, a stable equilibrium between cooperators and cheaters is expected (Foster, 2004; Archetti & Scheuring, 2011). This “diminishing return” of the common good (Foster, 2004; Gore et al., 2009) can account for the surprising observation that in the yeast invertase system, maximum group size is attained with a mixture of cheaters and cooperators: Cooperators produce more invertase than they can use, and cheaters convert this excess benefit into additional biomass (MacLean et al., 2010a). The cost-to-benefit ratio can be kept low if the common good is produced facultatively (i.e., only when needed), which is the case for most organisms, or if the durability of common goods is high, as found in siderophore production in Pseudomonas aeruginosa (Kümmerli et al., 2009c; Kümmerli & Brown, 2010).
(iv) For cooperation based on scarce common goods, mechanisms of “positive assortment” that increase the frequency of interactions between cooperators (Fletcher & Doebeli, 2009) can facilitate the persistence of cooperation. Positive assortment can involve specifically directing benefits to other cooperators and excluding or punishing cheaters based on recognition or previous experience (Hamilton, 1964; Trivers, 1971; Axelrod & Hamilton, 1981). This can occur even in organisms lacking nervous systems. For instance, microbes can achieve “recognition” through cell adhesion and chemical communication (Strassmann et al., 2011b), and legumes “reward” and “punish” beneficial and cheating rhizobia, respectively (Kiers et al., 2003; Heath & Tiffin, 2009). Another mechanism of positive assortment is “population viscosity,” brought about by limited dispersal in spatially structured environments, which keeps cooperators clustered with their relatives in homotypic cooperation (Hamilton, 1964; Maynard Smith, 1964; Chao & Levin, 1981; Nowak & May, 1992), or with their partners in heterotypic cooperation (Harcombe, 2010). Thus, natural cooperative systems, whether homotypic or heterotypic, use many different mechanisms to mitigate the tragedy of the commons, allowing cooperation via common goods to be a successful evolutionary strategy (Waite & Shou, 2012).
(v) The “adaptive race” model: If during adaptation to an environment, the fitness gain of cooperators exceeds that of cheaters by at least the fitness cost of cooperation, the tragedy of the commons can be averted. Although cooperators and cheaters sample from the same pool of adaptive mutations, this symmetry is soon broken: The best cooperators purge cheaters and continue to grow, whereas the best cheaters cause rapid self-extinction (Waite & Shou, 2012).
(vi) Direct and indirect reciprocity, sanctions, partner choice and fidelity are discussed in chapter 15.3.

Taking into account that cooperation may take place at a multitude of levels of communal life, “cheater” identification may not be trivial. Cheating, as noted by Lekberg and Koide (2014), must be defined in relation to an expectation. If the expectation is cooperation/mutualism, then anything less is parasitism. In that case, parasitism and cheating are synonymous, as suggested by Smith and Smith (2012). Obviously if the expectation were parasitism, not reciprocating could hardly be considered cheating. What in sociobiological parlance is rated as “cheater” phenotype may often only reflect dominance hierarchies (Fiegna & Velicer, 2005; Fiegna et al., 2006; Santorelli et al., 2008; Buttery et al., 2009; Vos & Velicer, 2009) that may be fixed or plastic depending on the ecological conditions in which the competitions take place. Such dominance hierarchies already take place within clonal communities where cell fate decisions (i.e. spore or stalk cell) are the result of stochastic and deterministic factors (see chapter 5). Intriguingly, within the conceptual framework of the kin selection theory the interpretation of these coercion-characterized dominance hierarchies are biased: when they occur within clonal populations they are interpreted as “altruism” (e.g. Strassmann et al., 2000), but when they occur between less related strains as “cheating”. Phenotypically, queens and dominants of eusocial colonies are social parasites, at least from the perspective of suppressed workers and helpers (see chapter 9), but I am not aware of any publication in which this has been noted explicitly.

Studies of helping in birds in particular tend to focus on a single aspect of helping behavior, usually nestling provisioning. Chick provisioning alone may not, however, be a representative measure of an individual’s overall contribution to the group’s reproductive success (Arnold et al., 2005). For example, a helper may contribute little in the form of chick feeding behavior, but by participating in sentinel or mobbing behaviors, that helper may reduce the risk of a group losing young to predators (Austad & Rabenold, 1985; Schaub et al., 1992; Hailman et al., 1994; Maklakov, 2002). Such a scenario could also explain why seemingly non-cooperative individuals (Heinsohn & Packer, 1995; Boland et al., 1997; Pusey & Packer, 1997) are tolerated in some communally breeding species. Heinsohn & Packer (1995), for example, found that female lions showed persistent individual differences in the degree to which they contributed to inter-group conflicts over territory. Intriguingly, other group members clearly recognised ‘cowardly lions’ but failed to punish them suggesting that cooperation was not maintained by simple reciprocity. Instead poor defenders might be too costly to eject from the group or they may be good at other activities. Within large complex social groups, it is possible that individuals might appear to defect only with respect to one form of helping, but contribute to other forms of communal behavior.

Lekberg and Koide (2014) further outlined the multiple levels of mutualistic relationships: “If cheating is simply defined as being less beneficial than others, mycorrhizal fungi that cheat plants almost certainly exist. Indeed, results from controlled greenhouse experiments show drastic differences in the effects on plant growth among fungal taxa (e.g., Bever, 2002; Munkvold et al., 2004; Smith et al., 2004; Pringle & Bever, 2008) and in the ratio of phosphate delivered to plant and carbohydrate transferred to fungus (Pearson & Jakobsen, 1993; Lendenmann et al., 2011). Because experiments are frequently performed under conditions that differ markedly from those in nature, labeling an arbuscular mycorrhizal fungus a cheat should be done with caution. In the greenhouse, for example, pathogen pressure may be low. But in the field, pathogens may be very important and the fungus that is poor at transferring phosphate may be superior in providing pathogen protection (Newsham et al., 1995). Also, while short-term greenhouse experiments may suggest the potential for cheating under very specific conditions, the real laboratory is the natural community in which time and space are very much expanded and in which plant populations struggle for survival over multiple generations during which environmental conditions are changing constantly. The importance of adopting a lifetime fitness view of symbiotic functioning was recently shown in an ant–Acacia symbiosis in which symbionts that appeared to be cheats at one life-stage were mutualistic at a different life-stage and contributed to increased lifetime fitness (Palmer et al., 2010). The same issue continues to be discussed and researched in regards to orchid mycorrhiza, which provide carbohydrate to germinating seedlings and may (Cameron et al., 2006) or may not (Rasmussen & Rasmussen, 2009) be repaid by adult plants.”

7.2.1 Frequency-dependence of cheater's payoff

Conflicts of interest exist between cooperators and “cheaters”. After competition between ‘‘cheat’’ and ‘‘co-operator’’ strains of yeast, population fitness is maximized under co-existence. The conditions necessary to recover the anti-intuitive result are 3-fold: (i) that resources are used inefficiently when they are abundant, (ii) that the amount of cooperation needed cannot be accurately assessed, and (iii) the population is structured, such that cooperators receive more of the resource than the “cheats” (MacLean et al., 2010a). Both experimental and theoretical work has shown that cooperative systems are highly tolerant to “cheater” exploitation (Marco et al., 2009). Frequency-dependent effects can even promote the maintenance of cooperative behavior in microbes and animal societies. Frequency-dependent selection on “cheating” phenotypes has recently been demonstrated in a number of studies (e.g., Velicer et al., 2000; Dugatkin et al., 2003, 2005; Fiegna & Velicer, 2003; Harrison et al., 2006; MacLean & Gudelj, 2006; Ross-Gillespie et al., 2007; Sandoz et al., 2007). Cheating or defecting strategies may do very well as long as they are rare in a population of cooperative individuals but fare worse or even perish when surrounded only by other cheaters (Nowak et al., 1994; Turner & Chao, 1999; Hauert et al., 2002; Sigmund, 2002; Le Galliard et al., 2003; MacLean & Gudelj, 2006; Diggle et al., 2007b; Ross-Gillespie et al., 2007; smith et al., 2010). Dugatkin et al. (2005) demonstrated that negative frequency-dependent mechanisms can act in Escherichia coli colonies: antibiotic-sensitive “cheaters” can persist at low concentrations (5–12%) on antibiotic-laced medium when associated with cooperative colonies of antibiotic-resistant individuals. “Cheaters” then enjoyed the benefit of antibiotic resistance without the cost of producing the resisting agent. However, “cheater” E. coli were prevented from reaching higher concentrations because they become more sensitive to the antibiotic substrate with increasing frequency. This is a case of negative frequency-dependent selection, where the fitness of a phenotype (e.g., “cheaters”) increases as it becomes rarer. As “cheaters” become more abundant either (i) they lose the benefit of associating with the group as “cheaters” reach numerical prevalence or (ii) colonies that harbor the highest concentrations of “cheaters” perform more poorly in interdemic (i.e., group-level) competition (Avilés 1986, 1993, 1997, 2002; Wilson & Wilson, 2007). In stationary phase cultures of E. coli mutants do not enter, or exit early, the nondividing stationary-phase state, cooperatively maintained by the wild type. Thus they end up overrepresented as compared to their initial frequency at the onset of the stationary phase, and subsequently they increase disproportionately their contribution in terms of progeny to the succeeding generation in the next growth cycle, which is a case of evolutionary “cheating”. However, the survival advantage of the mutant at any given time during a takeover is inversely dependent on its frequency in the population, its growth adversely affects the survival of the wild type, and its ability to survive in stationary phase at fixation is lower than that of its ancestor (Vulic & Kolter, 2001).

Simulation of the evolution of cooperation among nonrelatives revealed that cheaters increase in frequency when rare, but are selected against when common due to the reduced productivity of the groups they overburden with their presence. Freeloader frequencies thus periodically rise and fall around a dynamic equilibrium (Avilés, 2002). The spider Anelosimus studiosus shows a within-population social polymorphism with social (cooperative) and asocial (cheater) phenotypes. Asocial individuals experience negative frequency-dependent foraging success in staged foraging contests. Asocial individuals outperform social individuals when their representation is low, but lose this competitive advantage as their relative numbers increase (Pruitt & Riechert, 2009). In groups of foraging animals, individuals often have the choice between two behavioral roles: actively search for hidden food sources (‘producer’) or exploit food sources discovered by others (‘scrounger’) (Barnard & Sibly, 1981; Barnard, 1984; Beauchamp, 2013). In some bird species individuals readily switch between producer and scrounger strategies (Coolen et al., 2001). The effectiveness of scrounging decreases linearly as the frequency of scroungers increases, whereas the effectiveness of the producer strategy remains relatively constant (Mottley & Giraldeau, 2000; Morand-Ferron et al., 2007). Morand-Ferron et al. (2007) showed that the payoffs to kleptoparasites were frequency dependent; the more scroungers that were present in a group, the lower their payoffs (estimated in latency to a successful kleptoparasitic act). When they experimentally manipulated the costs of scrounging, they found that the frequency of scrounging was high when its costs were low, and the frequency was low when its costs were high. Experimental manipulation of the costs of producing (circuitous vs direct route) had the expected effect on the frequency of scrounging: when producing was cheap there were fewer scroungers, but the reverse was true when producing was costly. It has been suggested that scrounging could affect demographic parameters and predator–prey population dynamics (Coolen, 2002). By comparing the dynamics of groups where scrounging was either absent or present, Coolen et al. (2007) found that the presence of scrounging contributes to the regulation of both predator and prey populations. Moreover, their model predicted that prey persist at higher densities when scrounging is present because the exploitation pressure is then reduced. Predators therefore can exploit a higher density of resources when scrounging is present and consequently they can also maintain a higher abundance under this condition. Thus, although scrounging negatively affects the growth rate of a population, the presence of nonsearching individuals would also contribute to maintaining higher numbers of both prey and predators (Giraldeau & Dubois, 2008).

Producer-scrounger games have been found to apply to a potentially wide range of resource exploitation problems (Barnard, 1984). Scroungers can benefit from stealing not only food but also mating opportunities or parental care. Obligate and conditional reproductive “cheater” strategies, each with its own distinct genetic basis, do exist collaterally in numerous animal species as evolutionarily stable strategies (Alcock, 2005). The costs of courting and mating include mainly a decreased foraging efficiency and an increased mortality rate due to predation. Females’ preferences for exaggerated traits favor the use of alternative tactics, in which males steal fertilizations and hence parasitize the efforts of others while avoiding the cost of courting (Giraldeau & Dubois, 2008). Scrounging payoffs are negatively frequency-dependent and scrounger does worse than producer when scrounger is common but better than producer when scrounger is rare. There is empirical support that the proportion of males adopting alternative tactics may depend on social and ecological conditions, and that males are capable of adjusting their behavior in response to changes in the benefits associated with each tactic. For instance, the proportion of pupfish (Cyprinodon pecosensis) males adopting territorial, sneaker, and satellite breeding tactics depends on population density: as population density decreases, so does the benefit of maintaining a territory, leading to an increase in the proportion of sneakers and satellites (Kodric-Brown, 1986). The cost of calling, however, especially in terms of increased predation danger, will also play an important role. As the cost increases, it should erode the caller’s advantage and hence increase the proportion of satellite males. Tungara frogs, for instance, curtail conspicuous activities and reduce the intensity of their mating calls in response to simulated attacks by model bats (Ryan, 1985). Experimental manipulation of the risk of predation resulted in the expected change in tactics use in guppies (Godin, 1995): males performed a lower proportion of displays and increased their sneaky mating attempts when the risk of predation was increased.

A frequency-dependent behavioral evolutionary game concerns the parental investment decision of the great golden digger wasp (Sphex ichneumoneus) (Brockmann et al., 1979). The females of this species lay their egg in a burrow. They can either dig the burrow themselves; a strategy called ‘‘digging,’’ or use the burrow of another female, called ‘‘entering.’’ Digging is for all intents and purposes the producer strategy while entering corresponds to scrounger. Research with these insects has established that the frequency- dependence of payoffs contributes to maintaining these alternative parental strategies (Brockmann et al., 1979). Conspecific brood parasitism is a taxonomically widespread alternative reproductive tactic, particularly in birds, in which a female lays eggs in the nest or egg group of a conspecific that provides all subsequent parental care (de Valpine & Eadie, 2008). Because conspecifics provide the only hosts for brood parasites, obligate parasitism cannot become fixed in a population. Further, the advantages of parasitic laying are likely to be greatest when the frequency of parasitism is low and many host nests are available containing few parasitic eggs; the advantages will decrease as frequency of parasitism increases and more host nests contain many parasitic eggs (de Valpine & Eadie, 2008). In the treehopper insect (Publilia concava), Zink (2003) found that the time to find hosts increased as the frequency of the parasitic tactic increases. The hatching success of such parasitic eggs may be low when hosts can detect them and adopt countermeasures. In American coots (Fulica americana), for instance, parasitic eggs suffer a high mortality rate, mainly because of egg rejection by their hosts (Lyon, 1993).

Cheater resistance may be an important mechanism of social parasite/cheater control and might provoke coevolutionary arms races (Foitzik et al., 2003; Travisano & Velicer, 2004; Khare et al., 2009; Martin et al., 2010) similar to those seen in victim-exploiter/host–pathogen interactions (Gavrilets, 1997; Marques & Carthew, 2007). Possible examples of such coevolution include: (i) the conflict between brood parasitic cuckoos and their hosts, where the hosts are selected to reject cuckoo eggs, and the cuckoos are selected to circumvent this (Davies, 2000; Spottiswoode & Stevens, 2010; Langmore et al., 2011; Stoddard & Stevens, 2011); (ii) in the obligate mutualism between yuccas and yucca moths, selective abortion of flowers with heavy egg loads selects against moths that lay many eggs per flower or provide low-quality pollinations (Pellmyr & Huth, 1994); (iii) both “cheats” and cooperators performing better against their opponents from the past, but less well against future phenotypes, considering a laboratory selection experiment on biofilm production in Pseudomonas fluorescens (Zhang et al., 2009). Asymmetrical competition for commodities can explain the long-term persistence of mutualistic partnerships in spite of the evolution or incorporation of cheaters (Ferriere et al., 2002). Paradoxically, the presence of cheaters/parasites and cheater/parasite species in many mutualisms may be central to the maintenance of cooperation itself (Foster & Kokko, 2006; Holland et al., 2013). Consequently, mutualism–parasitism food web modules are common in nature (Ferriere et al., 2002; Genini et al. 2010; Holland et al., 2013). A mixture of “cheaters” and cooperators may even enable maximal group benefit (MacLean et al., 2010a). S. cerevisiae secrete invertase as a “public good” when the medium involves sucrose (and no glucose), which is digested by this enzyme. The resulting glucose is then taken up by the yeast cells. The growth rate is actually a concave function of the available glucose because there are diminishing returns to increased glucose. Thus, when the local glucose concentration is low, the increased glucose due to invertase production outweighs the cost of cooperation, but as there is more glucose in the environment from other cooperators, the additional growth rate per glucose molecule reduces below the cost of invertase production. Since production and secretion of invertase are costly, a “cheating” strategy is to take up glucose without making a contribution to the common pool of invertase. Defectors have a higher growth rate when the density of secreting cells is high but are less fit when the density of secreting cells is low (Greig & Travisano, 2004; Gore et al., 2009; MacLean et al., 2010a; Damore & Gore, 2012).

Like frequency dependence in classical population genetics, nonlinear social interactions can lead to coexistence of cooperators and defectors (Doebeli et al., 2004; smith et al., 2010; Damore & Gore, 2012). Archetti and Scheuring (2013) have shown that if benefits are nonlinear, mutualism can be maintained in well-mixed population without punishing the free riders. Examples of nonlinear benefit functions have been reported for cooperative hunting (Bednarz, 1988; Packer et al., 1990; Stander, 1991; Creel, 1997; Yip et al., 2008) and cooperative nesting (Rabenold, 1984) in vertebrates and for the production of diffusible enzymes in microbes, for example adhesive polymers (Rainey & Rainey, 2003) and antibiotic resistance in bacteria (Lee HH et al., 2010). This is likely to be a common feature of public goods in microbes, as the effect of enzyme production is generally a saturating function of its concentration (Hemker & Hemker, 1969), more specifically, a sigmoid function (Ricard & Noat, 1986). In certain cases, if a metabolic pathway is controlled by signal metabolites through an enzyme cascade, the final product follows a highly nonlinear switch-on switch-off behavior that resembles a step function (Mendes, 1997; Eungdamrong & Iyengar, 2004). Saturating benefits, and in particular sigmoid benefits, therefore are probably the rule in nature (Archetti & Scheuring, 2013). Nonadditivity makes cooperative sporulation remarkably resistant to exploitation by “cheater” strains showing how nonlinear interactions among cells insulate bacteria against cheaters (smith et al., 2010). Selection is driven by higher-order moments of population structure, not relatedness.

Under adverse environmental conditions defectors have an elevated mortality (Harms, 2001). The mere presence of a mortality risk gradient can create a kind of stochastic segregation between cooperators and defectors. Because the distribution of cooperators is patchy in the boundary region, eventually defectors run out of cooperators to free-ride on, and perish, making the distribution of cooperators even more uneven increasing the effective barriers against the spread of future invasions (Harms, 2001). Generally, coexistence is favored in an oscillating environment (such as feast and famine cycles) and periodicity increases the number of niches in the sense of the principle of competitive exclusion (Cushing, 1986; MacLean & Gudelj, 2006). Clonal amoebae and bacterial colonies with an asocial “cheater” phenotype (and no cooperators to parasitize) have a high risk to go extinct under conditions of feast and famine cycles (Hilson et al., 1994; Pál & Papp, 2000; Ennis et al., 2003; Fiegna & Velicer, 2003; Rainey & Rainey, 2003; Castillo et al., 2005; Kuzdzal-Fick et al., 2011; Waite & Shou, 2012; Waite, 2013). In chimera of wild-type and cheater mutants of D. discoideum, Gilbert et al. (2007) have calculated that, when the proportion of cheater cells exceeds 25%, the advantage of the cheater relative to wild type is lost. Due to abnormal stalk morphogenesis, some social cheaters would be doomed if their spores were dispersed to grow by themselves. Gilbert et al. (2007) have also sampled 95 independent fruiting bodies isolated in the wild and tested 3,316 spores without finding a colony with defective morphology, which indicates that, at least in a limited sample, morphologically recognizable D. discoideum cheaters are not common (Shaulsky & Kessin, 2007).

7.2.2 Social parasitism as bet-hedging strategy?

The biological default setting of individuals is neither selfish nor cooperative but ecologically context-dependent and dynamic. Competition and cooperation are threshold traits on a continuum of ecological variables. The traditional concepts of group and individual selection appear to be two extremes of a continuum, with systems in nature operating in the interval in between (Wilson DS, 1975). What appears to be cheating could actually be a strategy to associate with many partners, even those that may temporarily be poor mutualists, to maximize lifetime fitness in uncertain and stochastic environments. This has been referred to as bet-hedging (Lekberg & Koide, 2014, Heininger, 2015). Cues during child-hood (such as childhood exposure to family neglect, conflict and violence, and to neighborhood crime) that life will be harsh, unstable and short–these cues may probabilistically indicate that it is in one's fitness interests to exploit co-operators and to retaliate quickly against defectors (McCullough et al., 2012). Individual differences in mutual benefit cooperation are not merely due to genetic noise, random developmental variation or the operation of domain-general cultural learning mechanisms, but rather, might reflect the adaptive calibration of social strategies to local social-ecological conditions and their dynamics. Manipulation of neurohormonal signaling pathways of stress responses is able to affect the balance between cooperation and cheating in the coral reef cleaner-client fish mutualism (Bshary et al., 2011; Soares et al., 2012, 2014; Cardoso et al., 2015), arguing for the dynamic context-dependency of selfish and cooperative behaviors.

Conspecific brood parasitism (CBP) is a taxonomically widespread alternative reproductive tactic in which a female lays eggs in the nest or egg group of a conspecific that provides all subsequent parental care (de Valpine & Eadie, 2008). Spreading eggs among nests, parasites can increase the likelihood that at least some offspring will escape predation and survive to independence, also known as the “risk spreading” hypothesis (e.g., Rubenstein, 1982; Pöysä & Pesonen, 2007). In a well-studied model species of CBP, the common goldeneye (Bucephala clangula) Pöysä and coworkers (Pöysä 1999, 2003, 2006; Pöysä et al., 2001, 2014) found that nests are not predated at random and that parasites use public information in locating and selecting nests that have high prospects of success and preferentially lay in safe nests. By taking these findings into account, model simulations revealed that the selective advantage of parasitic egg laying related to nest predation is much higher than previously thought (Pöysä & Pesonen, 2007).

Division of labor is a routine behavior in cooperative communities (Arnold et al., 2005; see chapter 16.2), even in microbial biofilms (Crespi, 2001; Kearns, 2008; Nadell et al., 2009). Inter-individual heterogeneity in producing or responding to microbial quorum sensing signals may be interpreted as a sort of division of labor that confers group level benefits (e.g. Anetzberger et al., 2009, 2012). On the other hand, it may be considered as cheating or coercion, whereby individuals chemically manipulate quorum sensing-based collective decisions to their own advantage, and benefit at the expense of the group (Keller & Surette, 2006; Diggle et al., 2007a; Stacy et al., 2012). In fact, signal-blind quorum sensing mutants are known to arise in natural settings (e.g. the cystic fibrosis lung; Köhler et al., 2010), and in vitro experiments suggest that these mutants may be acting as cheats (Diggle et al., 2007b). Both scenarios, although plausible (Ross-Gillespie & Kümmerli, 2014), should exclude each other as division of labor should benefit the group while cheating should harm its fitness. An alternative interpretation could be that in some cases cheating may be part of a bet-hedging strategy: behavior that is harmful under some conditions may be beneficial under other ones.

In nature, soil moisture, nutrient availability, pathogen and herbivore pressure, local plant density, the amount of light available to the plant, etc., are in constant flux. Different fungal species and families differ in their growth patterns (Hart & Reader, 2002), which may result in different abilities to provide benefits. For example, the extensive extraradical mycelium of many members of the Gigasporaceae may be advantageous for phosphate acquisition, whereas the greater intraradical colonization by members of the Glomeraceae may help provide pathogen protection (Maherali & Klironomos, 2007; Sikes et al., 2009). From the fungal perspective, the phenology of root production varies with plant species, as do a number of other traits that may influence their quality as hosts. Given this variation, and because the degree of benefit both partners experience is context-dependent (Koide, 1991; Hoeksema et al., 2010), bet-hedging may be advantageous (Lekberg & Koide, 2014). For the arbuscular mycorrhizal symbiosis, this may mean that plants and fungi support multiple partners despite some of the partners being significantly less beneficial than others under the current circumstances (Lekberg & Koide, 2014). Lekberg et al. (2010) found that while the fungus allocated more phosphate to the host that provided more carbohydrate, arbuscular mycorrhizal colonization did not differ between the good and poor hosts (although the abundance of arbuscules was higher in the good host). One interpretation of these results is that selection has favored fungi that optimize their likelihood of acquiring carbohydrate across space and time by colonizing multiple plants. Also, perhaps the biggest surprise in the research of Kiers et al. (2011b) was not that the better phosphate-transferring fungus received more carbohydrate than the worse phosphate-transferring fungus, but that the latter received as much carbohydrate as it did. From a bet-hedging perspective, it is possible that selection has favored plants that associate with fungi that range in the benefits they provide (Lekberg & Koide, 2014). For example, the worse phosphate-transferring fungus may provide other benefits that become more important under different conditions (as carefully pointed out by Kiers et al., 2011b). For example, the behavior of the fungus they termed a “hoarder” may serve to store phosphate during winter months when plants are inactive (Merryweather & Fitter, 1998), thus preventing loss of important resources from the system (Lekberg & Koide, 2014).

The carbon that rhizobia in root nodules receive from their host fuels both reproduction and the synthesis of the storage polyester poly-3-hydroxybutyrate (PHB), as well as N2 fixation, which mainly benefits the host. Rhizobia escaping nodules can use stored PHB to survive starvation and reproduce up to 3-fold, but PHB synthesis is energetically expensive and trades-off with N2 fixation (Anderson & Dawes, 1990; Ratcliff et al., 2008). Some rhizobia have evolved sophisticated mechanisms to increase PHB accumulation, such as the production of rhizobitoxine, a chemical inhibitor of legume ethylene synthesis (Sugawara et al., 2006). Rhizobitoxine reduces host growth, decreasing rhizobia per nodule for all strains on a plant, but substantially increases PHB accumulation for rhizobitoxine-producing rhizobia (Ratcliff & Denison, 2009) and can be considered a tool for cheating (Ratcliff, 2010). In addition to enhancing reproduction, PHB has a role in bet-hedging: when starved, free-living high-PHB rhizobia divide asymmetrically, forming dormant, high-PHB "persisters" that survive long-term starvation and antibiotic treatment, and low-PHB "growers" that are sensitive to these stresses (Ratcliff, 2010). Sinorhizobium meliloti integrates bet hedging and phenotypic plasticity, forming fewer high-PHB persister cells when low competitor density predicts shorter-term starvation (Ratcliff, 2010). Since rhizobitoxine and PHB production is a marker of both cheating propensity and bet-hedging strategy, a synopsis of these phenomena identifies rhizobial cheating as potential bet-hedging strategy.

Pyoverdine defective mutants which have been isolated in natural populations, are potential cheats (De Vos et al., 2001; Visca et al., 2007). However, the status of a cheater is often determined unidimensionally, e.g. whether organisms contribute to a single commodity, e.g. in Pseudomonas aeruginosa siderophore production in iron-limited conditions (Griffin et al., 2004), ignoring the possibility that the organisms may contribute to another commodity of communal life (Kiers et al., 2011b). Iron is essential for, yet toxic to, bacteria. For most pathogens, including P. aeruginosa, there is intense competition for iron with the host (Braun, 1997; Ratledge & Dover, 2000; Skaar, 2010). Pyoverdine-deficient mutants have a greatly reduced ability to cause infection (Meyer et al., 1996; Takase et al., 2000). Therefore, it is surprising that in a detailed study of pyoverdine synthesized by P. aeruginosa obtained from cystic fibrosis patients, over one-quarter of the isolates were unable to synthesize pyoverdine (De Vos et al., 2001). By sequestering iron, subgrowth inhibitory concentrations of the mammalian iron chelator lactoferrin block the ability of P. aeruginosa biofilms to mature from thin layers of cells attached to a surface into large multicellular biofilm structures (Singh et al., 2002). On the other hand, pyoverdine defective mutants can form biofilms with the wild-type mushroom-like structures under provision of chelated forms of environmental iron. In fact, biofilms of the pvdA mutant grown in the presence of 1 μM ferric dicitrate show the mushroom-like appearance of the parent strain whereas the parent forms only the expected flat, thin biofilms (Banin et al., 2005). Intriguingly, mutants with a “cheater” phenotype in iron-limited conditions may achieve the maturation of biofilms when iron is only available in chelated forms. Thus, the “defectors” in iron-limited conditions turn out to provide another commodity for communal life and may be yet another example of bacterial bet-hedging. Likewise, bet-hedging strategies with regard to social competence in microbial populations resulting in superior competitors under conditions favoring antisocial behavior may be misidentified as cheating in more prosocial conditions (Brock et al., 2013; Kraemer & Velicer, 2014).

8. Eusociality


..ecological, demographic and physiological factors can be more important in promoting the evolution of eusociality.
Gadagkar, 2001

Summary
Eusociality is the Holy Grail of kin selection/inclusive fitness theory. Though eusociality is taxonomically rare, eusocial organisms frequently meet great ecological success and high persistence through geological time, with eusocial species representing approximately 50% of the world’s insect biomass.Each colony can be viewed as a superorganism, with a reproductive queen and non-reproductive workers as her extended phenotype. Natural selection in eusocial insects acts at the level of the colony; what benefits the queen will become heritable.
T
he haplodiploidy hypothesis of eusociality suggested that the high relatedness between colony members, resulting in indirect fitness gains of helpers, favored the evolution of eusociality and its associated reproductive skew. However, neither haplodiploidy nor split sex ratios in haplodiploid species appear to have played an important role in facilitating the evolution of eusociality. Currently, the monogamy hypothesis of eusociality is the most favored hypothesis by the advocates of kin selection theory. The phylogenetic evidence, however, is weak. Social systems of low relatedness have often evolved (e.g., becoming polygynous or polyandrous) from high-relatedness ancestral species.Groups of low relatedness are likely to have correlated benefits arising from increased group-level genetic diversity, and it appears these benefits often exceed the evolutionary advantages of high within-group relatedness.Eusociality is an accommodated/assimilated phenotype in 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 trigger.
An alternative hypothesis for the evolution of eusociality in insects is put forward that includes both genetic, physiological, behavioral and ecological factors that jointly contributed to the phenomenon. Body size/body mass predicts several fitness components, suggesting that it is a reliable proxy for fitness in many species. The respiratory system of insects is relatively simple: gas exchange takes place through a system of tubes, the tracheae that lead directly to the cells.Diffusion of oxygen into the organism limits the maximum possible size.Due to physiological and ecological constraints, body size is limited in insects, barring them from exploiting one of the most important means to increase fitness. However, another fitness-boosting strategy, cooperation with division of labor, stood at their disposal and, in fact, was exploited to create advanced integrative units, “superorganisms”.Brood care and, particularly, maternal care is the constitutional feature of any complex sociobiological behavior and is associated with neurobiological agents that mediate social behavior ranging from mother-offspring bonding to prosociality.

8.1 Ecological success

Eusociality (from Greek eu- meaning ‘well’ or ‘good’), is regarded the apex of animal social organization (Wilson, 1971; Queller & Strassmann, 2003). Eusociality has arisen independently at least 11 times in the insects (Cameron & Mardulyn, 2001; Brady et al., 2006; Hines et al., 2007; Cardinal et al., 2010), and eusocial insects have all converged on the following three characteristics: reproductive division of labor, cooperative brood care, and overlapping generations (Michener, 1969, 1974; Wilson, 1971; Fischman et al., 2011). Though eusociality is taxonomically rare (ca. 2% of the nearly 1 million known insect species are eusocial, Wilson, 1990; Hölldobler & Wilson, 2009), organisms that have achieved this higher level of organization frequently meet great ecological success, with eusocial species representing approximately 50% of the world’s insect biomass (Wilson, 1971, 1987; Strassmann & Queller, 2007). In the Amazonian rain forest near Manaus, for example, the eusocial ants, termites, bees and wasps constitute an astonishing 75% of the entire insect biomass (Fittkau & Klinge, 1973). Ants alone have more than 12,000 species and are found in every terrestrial habitat, using every resource. Their biomass exceeds that of all terrestrial non-human vertebrates combined (Hölldobler & Wilson, 1990). Humans, which can be loosely characterized as eusocial (Foster & Ratnieks, 2005), are dominant among the land vertebrates. Sherman et al. (1995) argued that cooperative breeding and eusociality are not discrete phenomena but rather form a continuum of fundamentally similar social systems and proposed to array invertebrate and vertebrate cooperative breeders along a common axis representing a standardized measure of reproductive variance.

Taking persistence of a phyletic line through geological time as the key measure of ecological success, eusocial ants have a much higher long-term success (and fitness, sensu Thoday, 1953; Cooper, 1984) than their extinct solitary counterparts (Wilson, 1987). Although there are exceptions (Danforth, 1990), low fecundity is a common feature in solitary aculeate Hymenopterans (Michener & Rettenmeyer, 1956; Danks, 1971; Else et al., 1978; Minckley et al., 1994; O’Neil, 2000), where females often produce a total of only 6–12 eggs in their lifetime. For example, females of the solitary specialist bee Dieunomia triangulifera produce a total of 2–6 eggs during their lifetime (Minckley et al., 1994), and those of Andrena erythronii produce a total of 8 eggs (Michener & Rettenmeyer, 1956). Similarly, in a survey based on 12 species of solitary nest-provisioning wasps, females produced an average of 9.8 eggs during their lifetime (O’Neil, 2000). Usually high preadult mortality rates (up to ~50–60%; Danks, 1971), common entire nest failure (Richards & Packer, 1995), and potentially high adult predation rates (Dukas, 2001) further combine to reduce net fecundities of solitary aculeate Hymenopterans to levels where the extinction vortex can be initiated (Zayed & Packer, 2005). In Ropalidia marginata, a solitary foundress produces on the average no more than one or two offspring whereas a queen of a multi-female colony produces on the average 76 offspring (Gadagkar, 1990a).

Most Hymenoptera are not social, although all non-social bee and wasp females have the same probability as the honeybees of being more closely related to sibs than to progeny (Slobodchikoff & Shields, 1988). For example, the vast majority of the ~18,000 described species of bees are solitary or cleptoparasitic, with only approximately 6% displaying eusociality (Danforth, 2007; Michener, 2007). Eusociality occurs in three halictid taxa: AugochloriniHalictus, and Lasioglossum. These groups have experienced several evolutionary reversals back to a solitary state, as many as 4–6 times both in Halictus and Lasioglossum (Wcislo & Danforth, 1997; Danforth, 2002; Danforth et al., 2003; Rehan et al., 2012). At least in these taxa evidence indicates extremely high barriers to the evolution of eusociality. Its origins are likely to have required very unusual life-history and ecological circumstances, rather than the amount of time that selection can operate on more simple forms of sociality (Rehan et al., 2012).

8.2 Superorganism

In eusocial insects, each colony can be viewed as a superorganism, with a reproductive queen and non-reproductive workers as her extended phenotype (Seeley, 1989; Wilson & Sober, 1989; Queller & Strassmann, 2002; Hölldobler & Wilson, 2009; Wilson & Nowak, 2014). Already Weismann (1893) was struck by the similarity between germ-soma specialization in multicellular organisms and the existence of reproductive and worker castes in ants: “The whole colony behaves as a single animal; the state is selected, not the single individual; and the various forms behave exactly like the parts of one individual in the course of ordinary selection” (p. 309). Weismann argued that the colony—not the individual ant—is the unit of selection by analogy to his view that it is germ-soma specialization in animals that defines them as individual units of selection. The term “superorganism” was first coined by Wheeler (1911), who focused on functional similarities such as metabolism, reproduction, and homeostasis, but who also stressed reproductive division of labor as analogous to germ-soma specialization: “One of the most general structural peculiarities of the [organism] is the duality of its composition, as expressed in the germ-plasm on the one hand and the soma on the other, and the same is true of the ant colony.” (Folse, 2011) Recently, an article in PNAS reported that Apoidea- and Apis-specific genes are enriched for signatures of positive selection, indicating that novel genes play a disproportionately large role in adaptive evolution of eusocial insects. Worker-biased proteins have higher signatures of adaptive evolution relative to queen-biased proteins, supporting the view that worker traits are key to adaptation. Moreover, genes regulating worker division of labor were found to be enriched for signs of positive selection (Harpur et al., 2014). In honeybees, in addition to the low viability of worker-laid eggs (Pirk et al., 2004), there are strong policing mechanisms which allow only fewer than 1 in 100 worker-laid eggs to develop (Ratnieks & Visscher, 1989) so that only about 1 in 1000 males reared to adulthood derives from a worker (Page & Erickson, 1988; Visscher, 1989). In addition, theory suggests that, all other factors being equal, indirect selection on workers will be effectively weaker than direct selection on queens (Linksvayer & Wade, 2009), especially when queens are polyandrous, as in A. mellifera (Hall & Goodisman, 2012). Yet, although direct worker contribution to the genes of the next generation is negligible the authors of the PNAS paper concluded that “it is reasonable to assume that indirect kin-selection is mostly responsible for the adaptive evolution of worker traits” (Harpur et al., 2014).

Reflecting Weismann’s argument, it has been stressed that natural selection in social insects acts at the level of the colony (Seeley, 1997; Korb & Heinze, 2004). What benefits the queen will become heritable. Darwin (1859) already suggested that worker traits could be favored by natural selection when they provide a benefit to reproductive family members within the colony. I would like to cite Nowak et al. (2010): “… descent is from queen to queen, with the worker force generated as an extension of the queen (or cooperating queens) in each generation. Selection acts on the traits of the queen and the extrasomatic projection of her personal genome and not the workers….The interaction between queen and workers is not a standard cooperative dilemma, because the latter are not independent agents. Their properties depend on the genotype of the queen and the sperm she has stored…. Inclusive fitness theory always claims to be a ‘gene-centred’ approach, but instead it is ‘worker-centred’: it puts the worker into the centre of attention and asks why does the worker behave altruistically and raise the offspring of another individual? The claim is that the answer to this question requires a theory that goes beyond the standard fitness concept of natural selection. But here we show that this is not the case.”

8.3 Kin selection and eusociality

Eusociality is the Holy Grail of kin selection/inclusive fitness theory (Queller & Strassmann, 1998; Hughes et al., 2008; Boomsma et al., 2010; Strassmann et al., 2010). Masterminded by the kin selection/inclusive fitness theory, various theories have been put forward that suggested that the high relatedness between colony members, resulting in indirect fitness gains of helpers, favored the evolution of eusociality and its associated reproductive skew.

8.3.1 Haplodiploidy hypothesis of eusociality

Hamilton suggested that due to the haplodiploid genetics of the Hymenoptera, females are relatively more related to their siblings than to their own offspring which promotes altruistic sib rearing and, consequently, the evolution of eusociality (Hamilton, 1972). Hamilton (1964, 1972) suggested that the inflated genetic relatedness of full sisters under haplodiploidy (life-for-life r = 3/4) explains why females, and not males, engage in altruistic sib-rearing in hymenopteran societies. This haplodiploidy hypothesis has fallen out of favor with the realization that the average value of siblings of both sexes is exactly the same (r = 1/2) for females and males, irrespective of the sex ratio (Charlesworth, 1978; Charnov, 1978; Craig, 1979; Bourke & Franks, 1995; Ross et al., 2013). Moreover, the haplodiploid hypothesis of eusociality (Hamilton, 1964; Trivers & Hare, 1976; Andersson, 1984) lost favor (e.g. Smith SM et al., 2009; Nowak et al., 2010; Gardner et al., 2012) since it is not compatible with the discovery of eusociality in diploid animals such as termites (Thorne, 1997; Smith SM et al., 2009), aphids (Aoki, 1977; Stern & Foster, 1996), platypodid ambrosia beetles (Kent & Simpson, 1992), Synalpheus sponge-dwelling shrimp (Duffy, 1996), and naked mole rats (Alexander et al., 1991) that do not have skewed degrees of genetic relatedness within families. Moreover, diploid meerkats (Suricata suricatta), dwarf mongoose (Helogale parvula), African wild dogs (Lycaon pictus) and silver backed jackals (Canis mesomelas) are all examples of mammalian species characterized by a reproductive division of labor, an overlap of generations and the co-operative care of offspring born to the group, thus satisfying the original criteria of eusociality (Batra 1966; Michener 1969; Wilson 1971; O’Riain & Faulkes, 2008). On the other hand, several large taxonomic groups are haplodiploid but do not have eusocial members (e.g., some mites, scale insects, whiteflies, and beetles). The parasitic Hymenoptera constitute an extremely large number of species yet they are all non-social (Askew, 1971). Two major differences between the solitary parasitic Hymenoptera and the social Hymenoptera are the abundance and distribution of food resources available to the larvae of each group (Slobodchikoff & Schulz, 1988). The association between haplodiploidy and eusociality is below statistical significance (Nowak et al., 2010), all of which suggests that haplodiploidy alone is neither necessary nor sufficient for eusociality to emerge.

8.3.1.1 Split sex ratios

In eusocial species, the sex ratio of helpers varies from female-only, in taxa such as the social Hymenoptera (ants, bees, and wasps) (Wilson EO, 1975), to an unbiased mixture of males and females, as in most termites (Thorne, 1997). One possibility for rescuing the haplodiploidy hypothesis involves split sex ratios, whereby there is a greater female bias among a potential altruist’s siblings than in the population as a whole (Trivers & Hare, 1976; Seger, 1983; Grafen, 1986; Grafen et al., 2012). It is thought that this allows the potential altruist to enjoy the benefit of close genetic similarity to her siblings (because they are primarily sisters) without the concomitant reduction in the reproductive value of females (because the population sex ratio need not be as biased; Boomsma & Grafen, 1990, 1991; Boomsma, 1991). However, it has been argued that Hamilton’s hypothesis does not work (Trivers & Hare, 1976; Charlesworth, 1978; Charnov, 1978; Craig R, 1982; Pamilo, 1991a; Bourke & Franks, 1995; Davies & Gardner, 2014) and that the sex of helpers could instead be explained by variation in the ecological factors that favor eusociality (Queller & Strassmann, 1998).

While most articles have examined whether the haplodiploidy hypothesis can be made to work with split sex ratios (e.g., Trivers & Hare, 1976; Seger, 1983; Grafen, 1986; Godfray & Grafen, 1988), Gardner et al. (2012) focused on those scenarios that are biologically most plausible and found that the extent to which haplodiploidy favors eusociality will be either:
(i) small (unmated females); in an empirically plausible range of unmatedness rates (0%–6%, with a mode of 0%), the potential for helping is boosted by only 0%–3% when considering the origin of helping and only 0%–7% when considering the subsequent elaboration of helping; or 
(ii) small to medium but not widespread (queen replacement). Under the empirically supported range of probabilities of queen survival (60%–80%), haplodiploidy always promotes helping, with the potential for facultative helping boosted by up to 50% and the potential for obligate helping boosted to up to 7%. The overall importance of this mechanism will depend on how frequently queen replacement leads to split sex ratios: to date, it has been found only in some cooperative bees, suggesting it is not a general factor on the route to eusociality (Boomsma, 1991; Mueller, 1991; Packer & Owen, 1994).

In the absence of split sex ratios, haplodiploidy neither promotes nor inhibits the origin of facultative helping, irrespective of the population sex ratio, because the increased relatedness to siblings is exactly balanced by the decreased reproductive value of daughters (Craig, 1979; Gardner et al., 2012). When parameterized with empirical data, results of two recent studies suggest that split sex ratios in haplodiploid species have not played an important role in facilitating the evolution of eusociality (Gardner et al., 2012; Alpedrinha et al., 2013). Ross et al. (2013) have provided clear support for the hypothesis that the sex of the helping caste in eusocial species is driven by ecology and not the genetics of sex determination. It appears that the sex bias in "altruistic" sib-rearing within this group of insects owes to a sex bias in the provision of parental care among their solitary ancestors, a pattern which holds across other eusocial species in which the original function of workers was nursing (Lin & Michener, 1972; Alexander, 1974; West-Eberhard, 1975; Evans, 1977; Charlesworth, 1978; Eickwort, 1981; Craig R, 1982; Andersson, 1984; Starr, 1985; Bourke & Franks, 1995; Queller & Strassmann, 1998; Ross et al., 2013; Davies & Gardner, 2014). Gardner et al. (2012) concluded “that: (1) many of the proposed consequences of haplodipoidy are unlikely to have been important for the evolution of eusociality, because they rely on biological assumptions that the comparative data suggest did not occur en route to eusociality, such as multiple mating or associations between same-generation breeders (the “semisocial route”); (2) the most plausible route by which the haplodiploidy hypothesis could work is with split sex ratios, building on Trivers and Hare (1976); (3) although split sex ratios can be favored for many reasons, there are only two mechanisms that have both been observed empirically and are consistent with the biology of primitively social hymenopterans—virginity and queen replacement; and (4) while these two mechanisms can lead to haplodiploidy favoring eusociality, the overall effect is likely to be small and can even be negative”.

8.3.2 Monogamy hypothesis of eusociality

The monogamy hypothesis is the currently most favored hypothesis for the evolution of eusociality (Boomsma, 2007, 2009, 2013; Hughes et al., 2008b; Cornwallis et al., 2010; Davies et al., 2012; Gardner et al., 2012; Wilson & Hölldobler, 2014). All currently available evidence indicates that obligatory sterile eusocial castes arose only via the lifetime association with monogamous mothers (Wilson & Nowak, 2014). Wilson and Hölldobler (2014) noted: “At least in ants and other social Hymenoptera, the reason for the monogamy window principle is open to several alternative explanations that are relevant to the origin of eusociality. The first explanation is based on direct observation of natural history. A single mating, with the sperm stored in the mother’s spermatheca, provides the same amount of genetic variation as matings by individual solitary (noneusocial) species. Because the earliest eusocial colonies consist of a relatively small number of individuals, the number of sperm from a single mating (paid out through the spermathecal valve) is adequate to last for the founding female’s usually brief lifetime as queen. Another selection pressure favoring this explanation of monogamy was seldom invoked by previous authors but also is confirmed by direct observation: The mortality of the eusocial foundresses is very high from the moment they leave the mother nest and mate until they finish constructing a nest. Time is of the essence in the interval between leaving the relative safety of the mother nest and entering the relative safety of the new one. […] …there is little relative advantage to a foundress to mate more than once. The advantage of acquiring greater genetic diversity in the small cohort of first offspring is outweighed by the protection afforded by a constructed nest.” The monogamy hypothesis was challenged (Nowak et al., 2010; Nonacs, 2011). Two recent papers have modeled the effects of the number of matings for the evolution of cooperation, and have found that monogamy often is not intrinsically advantageous. Indeed, polygamy often spreads cooperation faster through populations (Nonacs, 2011; Leggett et al., 2012). In contrast to the predictions of the kin selection theory, worker egg-laying and killing the queen is a much more common event in Bombus terrestris colonies (van Honk et al., 1981) than in genetically more diverse Apis mellifera colonies where workers are related by only 0.3 (Strassmann, 2001). Nowak et al. (2010) consider the phylogenetic evidence as weak because the ancestral state of the majority of solitary Hymenoptera species is likely to be monogamy. Because there are hundreds of extant related solitary species and only six known hymenopteran eusocial lines, the kinship-based monogamy hypothesis of Hughes et al. (2008) does not explain why only a few of the monogamous lines evolved eusociality (Wilson & Hölldobler, 2014). Hence, monogamy as an evolutionary preadaptation is indistinguishable from ecological factors favoring cooperation, such as defending nests or sequentially provisioning offspring.

8.3.3 Low relatedness due to polygyny and polyandry

Due to the haplodiploidy of Hymenoptera, mothers and daughters have a relatedness of 0.5 and relatedness for two sisters is always 0.75. Thus, a female would be able to pursue her genetic interests and increase her inclusive fitness more effectively by raising her mother’s daughters rather than her own. Hamilton (1964, 1972) argued that haplodiploidy makes eusociality easier to evolve, even in the absence of efficiency benefits to cooperation. In reality, however, genetic relatedness between workers in social insect colonies and the reproductive brood they rear is far below 0.75, the value expected for full sisters, often below 0.5 the value expected between mother and daughter and, not uncommonly, approaching zero. Such values are on account of queen turnover, multiple mating by queens or polygyny (Gadagkar, 1985).

An at least equally interesting message in the phylogenetic data may be how often social systems of low relatedness have evolved (e.g., becoming polygynous or polyandrous) from high-relatedness ancestral species (Nonacs, 2011). Multiple mating is prominent in eight genera, the honeybee (Apis: Palmer & Oldroyd, 2000), the yellow jacket wasps (Vespula: Goodisman et al., 2002), the seed harvester ants (Pogonomyrmex: Rheindt et al., 2004), the higher leaf-cutting ants (Atta and Acromyrmex: Boomsma et al., 1999; Sumner et al., 2004), the army ants (Dorylus and Eciton: Kronauer et al., 2004, 2006) and the desert ant (Cataglyphis: Pearcy et al., 2004). This suggests that polyandry has benefits to queen fitness (Fournier et al., 2008). Groups of low relatedness are likely to have correlated benefits arising from increased group-level genetic diversity, and it appears these benefits often exceed the evolutionary advantages of high within-group relatedness (Baer & Schmid-Hempel, 1999; Cole & Wiernasz, 1999; Rosengaus & Traniello, 2001; Tarpy, 2003; Nonacs & Kapheim, 2007, 2008). For example, the Australian desert ant Rhytidopondera mayri has relatedness levels of 0.1 to 0.2 among workers (Crozier et al., 1984). Similarly, multiple queen colonies of the fire ant Solenopsis invicta have a relatedness near zero among both the workers and the queens (Ross & Fletcher, 1985). In the neotropical multiple-queen wasp, Parachartergus colobopterus, average genetic relatedness among colony members overall is low (0.32 to 0.06) (Strassmann et al., 1997). In the social wasps Polistes exclamans the relatedness between workers and the brood that they raise is quite small (Strassmann, 1985). In polyandrous honeybees that have been estimated to mate as often as 17.3 times (Adams et al., 1977) relatedness among workers is also greatly reduced (Page & Metcalf, 1982).

Multiple-queen (polygynous) colonies are common in social insects, especially among ants, in which polygyny may be the predominant social structure (Rissing & Pollock, 1988; Keller, 1993, 1995). Colonies containing many queens are generally characterized by low within-nest genetic relatedness. Workers thus raise brood to which they are only distantly related, presenting a potential challenge to kin selection theory (Nonacs, 1988; Pamilo, 1991b; Bourke & Franks, 1995; Helms Cahan & Helms, 2012). Helms Cahan and Helms (2012) used microsatellites to show that Messor pergandei ants display primary polygyny in Californian and Nevadan populations but secondary monogyny (queens become aggressive following worker emergence and reduce to a single queen) in Arizona, but that co-foundresses are unrelated in all these populations. Since co-founding with non-relatives cannot yield any indirect fitness benefits, primary polygyny must have evolved in California and Nevada but not in Arizona for unknown ecological reasons. It has been suggested that ecological constraints on independent colony founding favor polygyny in a number of ant species (Crozier, 1979; Pamilo & Rosengren, 1984; Herbers, 1986, 1993; Hölldobler & Wilson, 1990; Heinze, 1993a; Rosengren et al., 1993; Bourke & Franks, 1995). This argument is akin to ecological constraints on independent breeding favoring communal breeding in birds, vertebrates and insects (Emlen, 1982b, 1991, 1994; Vehrencamp, 1983; Pamilo & Rosengren, 1984; Reeve, 1991; Koenig et al., 1992; Reeve & Ratnieks, 1993; Keller & Reeve, 1994). In fact, the degree of polygyny and the frequency of empty nest sites are inversely correlated across populations of Leptothorax longispinosus (Herbers, 1986). Moreover, an experimental increase of nest sites resulted in a decrease in the mean number of queens per nest. In a comparative study of leptothoracine ants, Bourke and Heinze (1994) found that polygyny was associated with nest-site limitation, cold climate and habitat patchiness, all factors that increase dispersal costs. The differences in reproductive patterns, dispersal habits and reproductive phenotypes associated with variation in queen number support the hypothesis that ecological constraints on independent colony founding are a major factor selecting for polygyny (Keller, 1995).

Although nests of most social insects are established independently by a single female (haplometrosis), colony founding by multiple females (pleometrosis) occurs in a number of social insect groups including ants (Hagen et al., 1988; Strassmann, 1989; Herbers, 1993; Sasaki et al., 1996; Choe & Perlman, 1997; Cahan et al., 1998; Bernasconi & Strassmann, 1999; Johnson, 2004), bees (Kukuk & Sage, 1994; Schwarz et al., 1997; Schwarz et al., 1998), wasps (Reeve, 1991; Ito, 1993), termites (Shellman-Reeve, 1997; Hacker et al., 2005), mites (Saito, 1997), aphids (Miller, 1998a, b, 2004), and thrips (Morris et al., 2002; Crespi et al., 2004; Bono & Crespi, 2006, 2008). Geographic variation in method of colony foundation (pleometrosis versus haplometrosis) without respect to relatedness occurs in desert ants Pogonomyrmex californicusAcromyrmex versicolorMyrmecocystus mimicus and Messor pergandei (Cahan et al., 1998; Rissing et al., 2000; Overson et al., 2013). Cofounding has been studied most extensively in the ants, bees, and wasps, and the relative importance of genetic and ecological factors vary considerably for different species. Bartz and Hölldobler (1982) found that colonies of Myrmecocystus mimicus, founded by a single queen, could not persist beyond a few weeks of foundation, because they could not produce sufficient workers fast enough to engage in inter-colony raiding, and subsequently had their broods stolen. A similar relationship has also been reported in Lasius niger (Sommer & Hölldobler, 1995). Jointly founded societies grow faster than founding colonies with a single queen and are thus believed to have an advantage in territorial contests with neighboring colonies (e.g. Bartz & Hölldobler, 1982; Tschinkel & Howard, 1983; Rissing & Pollock, 1987; Sommer & Hölldobler, 1995). Cooperating foundresses are generally unrelated in ants (Hagen et al., 1988; Strassmann, 1989; Sasaki et al., 1996; Bernasconi & Strassmann, 1999) and some bees (Kukuk & Sage, 1994; Danforth et al., 1996; Paxton et al., 1996); such cofoundress associations are usually communal in that reproduction is shared more or less equally among females, with low reproductive skew (Kukuk & Sage, 1994; Crespi & Choe, 1997; Reeve & Keller, 2001). Thus, per capita brood production either increases or remains constant with increasing numbers of foundresses (Kukuk & Sage 1994; Danforth et al., 1996; Crespi & Choe, 1997), though there are some exceptions, particularly in ants (Bernasconi & Strassmann, 1999). In other colonies of unrelated cofounders, once the first workers have eclosed, the cofoundresses typically become intolerant of each other, begin to fight and only a single queen survives per colony (Hölldobler & Wilson, 1990; Heinze, 1993b; Choe & Perlman, 1997; Bernasconi & Strassmann, 1999). In contrast, cofoundresses are typically related in eusocial wasps and bees (McCorquodale, 1988; Reeve, 1991; Kukuk & Sage, 1994; Crespi & Choe, 1997; Schwarz et al., 1997; Reeve & Keller, 2001), though cumulating evidence suggests that the presence of unrelated females may be more common than previously assumed (Queller et al., 2000; Fanelli et al., 2005; Liebert & Starks, 2006; Nonacs et al., 2006). In either case, production of workers and reproductives is often dominated by a single female (high reproductive skew) (Reeve, 1991; Reeve & Keller, 2001; Liebert & Starks, 2006) and per capita brood production declines with increasing group size (Michener, 1964; Karsai & Wenzel, 1998; Clouse, 2001; Soucy et al., 2003; Liebert & Starks, 2006), though there are some exceptions to this general pattern (Schwarz et al., 1998; Bouwma et al., 2006; Smith AR et al., 2007).

8.4 Eusociality is an assimilated/accommodated phenotype

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 trigger (Waddington, 1942, 1953, 1956, 1957; Scharloo, 1991; Masel, 2004; Braendle & Flatt, 2006; Pigliucci et al., 2006). Genetic assimilation is a special case of a more general phenomenon, called genetic accommodation (West-Eberhard, 2003; Braendle & 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 & 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 & Flatt, 2006). Importantly, the ecological conditions that gave rise to the assimilated phenomenon in the first place can be traced through deep evolutionary time by the “fossil record” in the genome (Runnegar, 1986; Buss, 1987 p. 90), and in physiology and development (Heininger, 2001, 2012, 2013). There are demographic predispositions to the evolution of eusociality (Gadagkar, 1991b). In concert with assured fitness returns, delayed reproductive maturation and variation in age at reproductive maturity become more powerful in selecting for worker behavior, and mixed reproductive strategies become available to a wider range of individuals. These phenomena provide a consistently more powerful selective advantage for the worker strategy than do genetic asymmetries created by haplodiploidy.

The evolutionary road to assimilated eusociality is highlighted by each newly formed colony of primitively eusocial Hymenoptera where aggression is one of the most distinct behavioral attributes of the queen to establish dominance hierarchies in small colonies (Michener & Brothers, 1974; Breed & Gamboa, 1977; Michener, 1990; Pabalan et al., 2000). Aggression by the dominant reproductive individual limits both mating opportunities and reproduction of subordinates (Reeve, 1991; Röseler, 1991; Peeters, 1993; Premnath et al., 1996; Monnin & Peeters, 1998). However, after reproductive dominance is established, a reduction in aggression levels often is observed, which appears to be mediated by chemical cues, at least in Ropalidia marginataPolistes dominulus (Premnath et al., 1996; Sledge & Boscaro, 2001) and Bombus terrestris (Fletcher & Ross, 1985; Röseler & Van Honk, 1990; Sramkova et al., 2008; Amsalem & Hefetz, 2010). Following the process of genetic assimilation, conditionally induced phenotypic variation becomes constitutively produced (i.e. no longer requires the environmental signal for expression) (Pigliucci et al., 2006). Thus, like sexual reproduction (Heininger, 2013), eusociality is an accommodated/assimilated phenotype.

With regard to the dynamic and rather gradual than distinctive aspects of accommodation and assimilation of eusociality, the current definition of eusociality appears highly artificial and not helpful to understand its evolution (O’Riain & Faulkes, 2008). Various authors, e.g., Kukuk (1994), Gadagkar (1994a), Crespi and Yanega (1995), and Sherman et al. (1995) have argued for the redefinition of eusociality. Arguments (reviewed by Costa and Fitzgerald, 1996) range from the expanded view of Sherman et al. (1995), which attempts to classify all species with evidence of reproductive skew as eusocial, to the narrow view of Crespi and Yanega (1995) that argue for the restriction of eusociality to species characterized by irreversible behavioral or morphological castes. Gadagkar (1994a) suggested that, (i) the scope of eusociality is expanded to include semisocial species, primitively eusocial species, highly eusocial species as well as those cooperatively breeding birds and mammals where individuals give up substantial or all personal reproduction for aiding conspecifics, (ii) there should be no requirement of overlap of generations or of life-time sterility, and (iii) the distinction between primitively and highly eusocial should continue, based on the presence or absence of morphological caste differentiation.

8.5 A joint genetic-physiological-behavioral- ecological hypothesis of eusociality in insects

Here I put forward an alternative hypothesis for the evolution of eusociality in insects that includes both genetic, physiological, behavioral and ecological factors that jointly contributed to the phenomenon. The various factors that bring about eusocial behavior are ordered hierarchically from genetic to ecological levels. This ordering does not reflect a weighting of the relative importance of the various levels for the evolution of eusociality. I think that for each road to eusociality a specific blend of different factors effected the phenotype.

8.5.1 Genetic

Recombination directly affects the multilocus genotypic diversity among offspring and also influences population level genetic diversity through background selection or selective sweeps that depend on linkage disequilibria. Genetic recombination has been shown to increase colony genetic diversity (Sirviö et al., 2006; Oldroyd & Fewell, 2007). Generally, eusocial taxa have a high recombination rate (Beye et al., 2006; Sirviö et al., 2006, 2011; Wilfert et al., 2007). For instance, the polyandrous social honeybee has a genome-wide recombination rate approximately ten times higher than any other higher eukaryote studied so far (Hunt & Page, 1995; Beye et al., 2006; The Honeybee Genome Sequencing Consortium, 2006). Although not as high as the honeybee recombination rate, the phylogenetically very similar primitively eusocial and predominantly monogamous bumble bee Bombus terrestris also has a very high recombination rate (Stolle et al., 2011). The high genome-wide recombination rate of both bee species clearly exceeds the average recombination rate in insects or vertebrates (Wilfert et al., 2007; Lattorff & Moritz, 2008). This high recombination rate is probably not due to Red Queen host-parasite coevolution (Kidner & Moritz, 2013). In contrast, the genome-wide recombination rate for the solitary Nasonia parasitoid wasp (Niehuis et al., 2010) is less than one tenth of the rate reported for the honeybee. Thus, social Hymenoptera show a higher recombination rate (mean 10.27 cM/Mb, n = 4) than nonsocial parasitoid Hymenoptera (mean 3.99 cM/Mb, n = 4) (Wilfert et al., 2007; Lattorff & Moritz, 2008). Within species, recombination rates can also vary among individuals of the same sex (e.g. Broman et al., 1998; Kong et al., 2004), between the sexes (Lynn et al., 2004) and among different parts of the genome (Jensen-Seaman et al., 2004). These findings indicate that monogamy is the ancestral state in both solitary and eusocial Hymenoptera and thus hardly qualifies as the state facilitating the road to eusociality (see Nowak et al., 2010). Rather, an increased rate of recombination appears to differentiate already primitively eusocial from solitary Hymenoptera. Recombination rate is a genetically regulated trait that can be changed by selection without significant constraints (Otto & Lenormand, 2002), as natural populations of closely related species have been reported to differ in their crossover frequencies (True et al., 1996). Artificial selection has led to increased recombination rates in domesticated animals (Burt & Bell, 1987) and Otto & Lenormand (2002) estimated that the recombination frequency (number of chiasmata) increased by 24.6% over a median of 50 generations of artificial selection for traits unrelated to recombination. Given that recombination rate is readily selected for, it appears to have evolved in eusocial species as a secondary trait to increase genotypic diversity among offspring as a means to maintain colony fitness (see below) and eusociality.

High recombination rates (Wilfert et al., 2007), multiple mating (Boomsma & Ratnieks, 1996; Kronauer et al., 2007) and multiple queens (Hölldobler & Wilson, 1990), which have evolved many times in some social insect lineages (Hughes et al., 2008a, b; Nonacs, 2011), all lead to high levels of genetic variation among colony members. In fact, several species of social insect have among the highest reported degrees of polyandry and recombination rates of all animals (Boomsma & Ratnieks, 1996; Fuchs & Moritz, 1999; Kronauer et al., 2007; Wilfert et al., 2007; Smith et al., 2008). The resulting genetic diversity within colonies of these taxa generates low within-colony relatedness and more than 12 hypotheses have now been advanced to explain this transition (Palmer & Oldroyd, 2000). Of these, four are generally regarded as most plausible (Crozier & Fjerdingstad, 2001). They are: (i) bet-hedging against insufficient or deficient sperm (Cole, 1983); (ii) reducing the risk of producing diploid (non-functional) males, which arise as a byproduct of the sex determination system (Page, 1980); (iii) providing disease resistance via worker genetic diversity (Sherman et al., 1988); and (iv) generating a more stable and resilient system of division of labor (Crozier & Page, 1985). High genetic diversity of workers increases colony performance by enhancing disease/parasitism resistance (Shykoff & Paul Schmid-Hempel, 1991; Liersch & Schmid-Hempel, 1998; Baer & Schmid-Hempel, 1999; Brown & Schmid-Hempel, 2003; Tarpy, 2003; Tarpy & Seeley, 2006; Seeley & Tarpy, 2007; Mattila et al., 2012), division of labor (Crozier & Page, 1985; Gove et al., 2009), and a number of other potential fitness-enhancing mechanisms (Crozier & Page, 1985; Crozier & Fjerdingstad, 2001; Jones et al., 2004; Mattila & Seeley, 2007; Oldroyd & Fewell, 2007; Mattila et al., 2008, 2012; Wiernasz et al., 2008).

8.5.2 Physiological

Body size/body mass predicts several fitness components, suggesting that it is a reliable proxy for fitness in many species (Mittelbach, 1981; Peters, 1983; Post & Evans, 1989; van den Berghe & Gross, 1989; Merrett, 1994; Lundvall et al., 1999; Roitberg et al., 2001; Kingsolver & Pfennig, 2004; Neff & Cargnelli, 2004; Brown C et al., 2007; Pfennig et al., 2007; Schwagmeyer & Mock, 2008). For geometrically and physiologically simple aerobic organisms, diffusion of oxygen into the organism limits the maximum possible size (Raff & Raff, 1970; Alexander, 1971). It has been suggested that the atmospheric oxygen partial pressure (aPO2) provides a key biophysical limitation on the maximal size of some animal groups, and specifically that elevations in aPO2 increased oxygen supply to the tissues, allowing larger body sizes (Graham et al., 1995; Dudley, 1998). According to this oxygen-limitation hypothesis, high aPO2 values reaching a maximum of 27–35 kPa in the late Carboniferous and early Permian (< 300 million years ago) led to the evolution of giant organisms in many animal groups (Graham et al., 1995; Dudley, 1998; Harrison, et al., 2010). The subsequent aPO2 decrease to 13 kPa in the Triassic (Berner, 2006) resulted in lower gas transfer rates and in substantially smaller body sizes of the surviving fauna of these groups (Dudley, 1998; Klok & Harrison, 2009).

Insects were one of the major animal groups that experienced gigantism in the late Paleozoic (Shear & Kukalová-Peck, 1990; Grimaldi & Engel, 2005). Their respiratory system is relatively simple. Gas exchange takes place through a system of tubes, the tracheae that lead directly to the cells (Chapman, 1998). The circulatory system contributes little to oxygen transport, because the capacitance of hemolymph for oxygen is low. In contrast, most gill- and lung-breathing animals must operate respiratory and circulatory systems in series to deliver oxygen. A variety of recent empirical findings support a link between oxygen and insect size: (i) Cumulative evidence shows that oxygen levels influence body size in insects (Loudon, 1988; Dudley, 1998; Frazier et al., 2001; Harrison et al., 2006; Kaiser et al., 2007; Klok & Harrison, 2009; Harrison & Haddad, 2011; Heinrich et al., 2011). Hypoxia not only reduces body size (Harrison et al., 2006), but also constrains the evolution of increased body size by limiting the variation available to selection (Klok & Harrison, 2009). (ii) Insects developmentally and evolutionarily reduce their proportional investment in the tracheal system when living in higher aPO2, suggesting that there are significant costs associated with tracheal system structure and function (Locke, 1958; Loudon, 1989; Jarecki et al., 1999; Henry & Harrison, 2004; Arquier et al., 2006; Klok et al., 2010). (iii) Larger insects invest more of their body in the tracheal system, potentially leading to greater effects of aPO2 on larger insects. Studies that have investigated the scaling relationship of the tracheal system to date suggest that tracheal investment is hypermetric, with greater proportional investment in larger insects (Kaiser et al., 207; Greenlee et al., 2009). During ontogeny of the American locust, Schistocerca americana, tracheal investment increases in the leg muscle (Hartung et al., 2004) and at the whole-body level with tracheal volumes and ventilation scaling approximately with mass1.3 (Lease et al., 2006; Greenlee et al., 2009). Similarly, across four tenebrionid beetle species, tracheal volumes scale with mass1.29 (Kaiser et al., 2007). Such a trend appears to be general for insects: tiny stick insects have tracheal volumes of around 2 per cent (Schmitz & Perry, 1999), while giant scarabaeid beetles have tremendous air sacs (Miller, 1966). Theoretical calculations suggest that the observed hypermetry is consistent with a need to overcome reduced rates of diffusive gas exchange in longer, blind-ended tracheoles (Harrison et al., 2009). The relationship between atmospheric pO2 and maximum insect size is more complicated than implied by coincidence of late Paleozoic hyperoxia and insect gigantism. The overall correlation between pO2 and maximum wing length in each 10-Myr bin is highly significant (r = 0.54, P = 0.002), although the strength of the size-oxygen relationship is greatly diminished (P = 0.42) after controlling for collection paleolatitude (a proxy for temperature) via multiple linear regression or for autocorrelation in body-size data using generalized least-squared regression (P = 0.62, P = 0.80 including paleolatitude) (Clapham & Karr, 2012). These results, as well as instances of decoupling, such as an Early Cretaceous decrease in insect size during a substantial increase in atmospheric oxygen, imply that atmospheric oxygen did not control maximum body size over the entire evolutionary history of insects, or that it was not the only control raising the possibility that other factors, such as competitive or predatory interactions with flying vertebrates (birds, bats, and pterosaurs), may have contributed to or even been the dominant control on evolutionary size trends (Okajima, 2008; Butterfield, 2009; Chown, 2012; Clapham & Karr, 2012).

Putative termite nest fossils have been reported as early as the Triassic (Hasiotis, 2003; Bordy et al., 2009). Molecular evidence estimated Isoptera, all with social behavior, to have diverged in the mid Triassic/Early Jurassic (Ware et al., 2010). Although it has been questioned whether the imposing nests of Triassic and Early Jurassic insects have belonged to termites (Grimaldi & Engel, 2005; Vršanský & Aristov, 2014), the nests allow to date the evolutionary origins of insect eusociality to the Triassic. Intriguingly, the same signaling pathways, IIS and TOR that control body size in insects (Edgar, 2006), also regulate eusocial behavior including reproductive repression and division of labor (see chapter 11.2). The association of low aPO2, small insect body size and the evolutionary innovation of eusociality in the Triassic period may suggest a relationship between these factors. Due to physiological and ecological constraints, body size is limited in insects, barring them from exploiting one of the most important means to increase fitness. However, another fitness-boosting strategy, cooperation with division of labor, stood at their disposal and, in fact, was exploited to create advanced integrative units, “superorganisms”, a collection of single creatures that together possess the functional organization implicit in the formal definition of organism (Wilson & Sober, 1989).

An intriguing clue for the role of oxygen-body size limitation in the evolution of eusociality may come from the developmental interplay of oxygen and temperature (Frazier et al., 2001; Verberk & Bilton, 2011). Lower temperatures tend to cause insects to be larger, via both direct developmental effects and by evolutionary changes in mean size (Partridge & French, 1996; Kingsolver & Huey, 2008; Chown & Gaston, 2010). Although higher temperatures strongly stimulate ectothermic metabolic rates, they only slightly increase oxygen diffusion rates and decrease oxygen solubility. Consequently, insect gas exchange systems have more difficulty meeting tissue oxygen demands at higher temperatures. Higher temperatures reduce oxygen delivery capacity relative to tissue oxygen needs, which may partially explain why ectotherms are smaller when development occurs at higher temperatures (Frazier et al., 2001; Woods & Hill, 2004). Molecular divergence dating infers recent and simultaneous origins for halictid eusociality, ~20–25 Ma in each of the three groups AugochloriniHalictus, and Lasioglossum (Brady et al., 2006). This time period coincides with a global warming trend during the late Oligocene warming and mid-Miocene climatic optimum (Zachos et al., 2001). A potential correlation between eusocial evolution and climatic warming is strengthened by observations that climatic factors influence the manifestation of eusociality in some modern species that have solitary forms in colder areas and eusocial forms in warmer areas (Sakagami & Munakata, 1972; Yanega, 1988; Eickwort et al., 1996; Miyanaga et al., 1999; Richards, 2001; Soucy, 2002; Soucy & Danforth, 2002; Cronin & Hirata, 2003; Bradley et al., 2009; Soro et al., 2010). Intriguingly, polar gigantism of benthic amphipod crustaceans is dictated by oxygen availability (Chapelle & Peck, 1999).

8.5.3 Behavioral

Brood care and, particularly, maternal care is the constitutional feature of any complex sociobiological behavior. I am not aware of any species without brood care that was ever able to evolve a complex society. The mother-offspring affiliation can be considered the nucleus of cooperative societies. There is a continuum of neurobiological agents that mediate social behavior ranging from mother-offspring bonding to prosocial behavior in larger societies.

The formation of groups within a freely mixing population can occur in many ways (Wade, 1976; Swenson et al., 2000; Gadagkar, 2001; Pepper & Smuts, 2002; Thorne et al., 2003; Fletcher & Zwick, 2004; Hunt, 2007; Khila & Abouheif, 2010; Wade et al., 2010). A group can be pulled together when cooperation among unrelated members proves beneficial to them, whether by simple reciprocity or by mutualistic synergism or manipulation (Clutton-Brock, 2009a). A growing number of studies have documented the occasional merging of unrelated colonies in ants, wasps, and termites (Foitzik & Heinze, 1998, 2000; Fisher et al., 2004; Prezoto & Santos-Prezoto, 2005; Johns et al., 2009; Kellner et al., 2010; Kronauer et al., 2010; Kellner & Heinze, 2011). Grouping by family can hasten the spread of eusocial alleles, but it is not a causative agent. The causative agent is the advantage of a defensible nest, especially one both expensive to make and within reach of adequate food (Nowak et al., 2010).

Aggression has been long recognized as an important component of social systems (Hall, 1964). Aggression is often considered a category of anti-social interaction that weakens the positive social consequences of cooperation and other types of affiliative interactions (e.g. Kummer, 1971; Brown, 1976, pp. 75, 253). Slobodchikoff and Schulz (1988) modelled the relationship between resources and group size as a function of aggression. Generally, cooperation decreased as aggression increased. Aggression can provide a mechanism for controlling the size of the group, for defining the distribution of resources within the group, and for defining the reproductive relationships of the group. By varying the intensity of aggression only, one may switch from egalitarian to despotic virtual societies (Hemelrijk, 1999).

Aggression could be a barrier to cooperative sociality (Cahan et al., 1998; Overson et al., 2014). Aggression serves as perhaps the most important proximate barrier to pleometrotic cooperative group formation by normally haplometrotic queens. Although queens were not immediately aggressive, haplometrotic groups showed a much higher incidence of aggression over time with associated effects on queen mortality, than did pleometrotic associations (Overson et al., 2014). Less favorable conditions such as food shortage might lead to more frequent aggression and accentuated the reproductive dominance of one queen in pleometrotic ant colonies (Sommeijer & Van Veen, 1990; Kolmer & Heinze, 2000). A dominance hierarchy may even support the beneficial competitive nature of social living as a positive selective force (West, 1967; Moosa & Ud-Dean, 2011).

8.5.4 Ecological

The non-social parasitic Hymenoptera (Askew, 1971) differ from the social Hymenoptera with regard to the abundance and distribution of food resources available to the larvae of each group (Slobodchikoff & Schulz, 1988). The highly eusocial allodapine bee species Exoneurella tridentata, appears to have evolved sociality in very harsh, xeric conditions (Dew et al., 2012), and in years with harsh weather conditions colonies of primitively eusocial sweat bee, Halictus ligatus, showed an increase in sociality, i.e. higher levels of queen–worker dimorphism and decreased worker cheating (Richards & Packer, 1996). In a similar vein, communality among the bees may also be associated with more arid environments (Australia, Southwest USA) possibly because of restricted opportunities for fossorial nesting (Wcislo & Tierney, 2009). Within families and within species of arthropods, many organisms show an increasing degree of sociality at lower latitudes and altitudes. In a smaller number of cases, organisms form larger groups or found nests cooperatively at higher latitudes and altitudes (Purcell, 2011). In some ant, termite, and social spider species, rainfall is positively correlated with sociality (Riechert et al., 1986; Murphy & Breed, 2007; Picker et al., 2007; Purcell, 2011). Harsh environmental conditions may lead to a behavioral syndrome (Hemelrijk, 1999) promoting the evolution of a despotic regime (Sommeijer & Van Veen, 1990; Kolmer & Heinze, 2000).

It has been suggested that ecological constraints on independent colony founding favor polygyny in a number of ant species (Crozier, 1979; Pamilo & Rosengren, 1984; Herbers, 1986, 1993; Hölldobler & Wilson, 1990; Heinze, 1993a; Rosengren et al., 1993; Bourke & Franks, 1995). This argument is akin to ecological constraints on independent breeding favoring communal breeding in birds, vertebrates and insects (Emlen, 1982b, 1991, 1994; Vehrencamp, 1983; Pamilo & Rosengren, 1984; Reeve, 1991; Koenig et al., 1992; Reeve & Ratnieks, 1993; Keller & Reeve, 1994). In fact, the degree of polygyny and the frequency of empty nest sites are inversely correlated across populations of Leptothorax longispinosus (Herbers, 1986). Moreover, an experimental increase of nest sites resulted in a decrease in the mean number of queens per nest. In a comparative study of leptothoracine ants, Bourke and Heinze (1994) found that polygyny was associated with nest-site limitation, cold climate and habitat patchiness, all factors that increase dispersal costs. The differences in reproductive patterns, dispersal habits and reproductive phenotypes associated with variation in queen number support the hypothesis that ecological constraints on independent colony founding are a major factor selecting for polygyny (Keller, 1995).

9. Reproductive skew


From this arises a debate: if it is better to be loved than to be feared, or the contrary. I reply that it would be nice to be both, but because they are difficult to combine together, if you cannot have both, it is much more secure to be feared than to be loved.
Niccoló Machiavelli (1532)

Summary
Dominance hierarchies determine access to resources. In fact, cooperative breeding coalitions are dominant-enforced social exploitations. Subordinates in cooperative societies commonly show a degree of reproductive restraint due to factors such as a lack of access to unrelated breeding partners, targeted aggression, temporary eviction or infanticide. Generally, harsh and stochastic environmental conditions can lead to a behavioral syndrome that promotes the evolution of a despotic regime. Suppression of subordinate reproduction by dominants and high levels of reproductive skew are often found in species that have evolved in harsh conditions, where competition for resources is likely to be intense and reproduction by subordinates particularly costly to dominants. There is an interrelationship between limited dispersal (due to limited options for independent breeding) and asymmetric conflict/dominance hierarchies that underlie coerced cooperation. Beehives, wasp and ant nests are police states in which selfish and despotic queens, by means of their pheromones suppress the reproductive activity of their daughters and enforce their reproductive monopoly by murder, torture and imprisonment.
Manipulative queen control leads to an evolutionary arms race between queen and workers reminiscent of parasite-host coevolutionary dynamics. When the dominant position becomes vacant previously “altruistic” subordinates may fight with their siblings until death to take over the dominant position: not exactly what can be expected from a mutually agreed-upon social contract based upon inclusive fitness gains.

Animals that rank highly in the social hierarchy have many advantages, such as better access to food, territories, mates and offspring (Huntingford & Turner, 1987; Jones & Mensch, 1991; Collias et al., 1994; Hall & Fegigan, 1997; Lahti et al., 1998). Several models point out the evolutionary route to despotic societies (Vehrencamp, 1983; Hemelrijk, 1999; Summers, 2005). Generally, the harsh and stochastic environmental conditions that select for cooperative behavior in the first place (see chapters 6 and 15) can lead to a behavioral syndrome (Hemelrijk, 1999) that promotes the evolution of a despotic regime (Sommeijer & Van Veen, 1990; Kolmer & Heinze, 2000). Experimental studies support the notion that resource limitation results in enhanced aggression and dominance hierarchies (Machmer & Ydenberg, 1998; Cook et al., 2000; West et al., 2001; Innocent et al., 2011). In habitats where resources may not be sufficient to sustain populations above the threshold of extinction, it may be a good strategy for dominants to “negotiate” with the subordinates to help them in the rearing of offspring. Moreover, a mutualism-competition model suggested that an inferior competitor, if cooperative to a superior competitor, is able to survive (Zhang, 2003). Given the manifold advantages of cooperation, evolution should have “found” ways to establish these cooperative alliances. However, there should have remained much conflict over the fair share of reproductive opportunities for subordinates. Models of reproductive skew were developed by Vehrencamp (1979, 1983) and Emlen (1982a). These models were inspired (Vehrencamp, 1984) by previous arguments concerning the relationship between subordinates’ options for dispersal and the ability of dominants to monopolize reproduction (Alexander, 1974). The main objective of the models was to determine the allocation of resources or fitness to dominants and subordinates that maximizes the fitness of the dominant. Skew was predicted to increase with increasing ecological constraints (a measure of the expected success of a subordinate that attempts to disperse and reproduce). The greater the difference in fitness for groups and solitary individuals, the greater the skew can be. Some reproductive skew models entailed a form of ‘social contract’ between dominants and subordinates, whereby the dominant should, in certain situations, yield reproductive concessions to the subordinate as an incentive for the subordinate to stay and help rather than leave or to dissuade them from engaging the dominant in a fight (Alexander, 1974, pp. 350-351; Reeve & Ratnieks, 1993; Reeve & Keller, 1997). For most social animals, however, reproductive concessions are unrealistic and most cooperative interactions are based instead on the direct adjustment of subordinates’ behavior by dominant individuals, i.e. manipulation. Crespi and Ragsdale (2000) assumed that dominants impose costs on subordinates which tip the behavioral decisions of subordinates towards staying and helping rather than leaving. Thus, rather than providing reproductive benefits (concessions) to subordinates in order to induce them to stay and help, dominants impose costs on subordinates which make staying and helping the subordinate’s best strategy. However, attempting to exert control over the distribution of reproduction is assumed to be itself costly, reducing the total reproductive output (Reeve et al., 1998; Johnstone, 2000; Reeve & Shen, 2006, 2013; Cant, 2012). Some forms of manipulation, such as reducing the feeding levels or rates of developing young, are cheap and easy for the dominant to perform, whereas other forms, such as harassment (e.g. Michener & Brothers, 1974; Emlen, 1982a; Reyer et al., 1986; Abbott, 1987; Clutton-Brock & Parker, 1995; Creel & MacDonald, 1995) or disruption of independent breeding attempts (e.g. Emlen & Wrege, 1992), will demand some energy or time investment by the dominant.

Strangely, the advocates of the kin selection theory are so prejudiced with the alleged “altruism” side of the coin that they widely ignore the factual selfishness flipside of the coin (but see Michener & Brothers, 1974). In fact, Hamilton’s rule is a zero-sum game: The “altruism” from one side is exactly balanced by the selfish exploitation from the other side. This oppressive system is driven by a dynamic balance of dominance and subordination (Chapuisat & Keller, 1999; Passera et al., 2001; Tóth et al., 2002; Beekman et al., 2003; Beekman & Ratnieks, 2003; Mehdiabadi et al., 2003; Rosset & Chapuisat, 2006; Helanterä & Ratnieks, 2009). The dynamics not only applies to the queen-worker conflict but extends to the individual workers: For example, when a breeder’s position becomes available in termite colonies, a subset of workers develops into neotenic replacement reproductives that will fight with their siblings until death to take over the queen-/kingless colony (Lenz, 1985; Thorne, 1997). The higher the reproductive skew the more despotic and oppressive is the social system. Again, this dynamic balance is the result of an arms race. Victim-exploiter/host–parasite interactions with Red Queen dynamics (e.g. Gavrilets, 1997; Summers et al., 2003; Marques & Carthew, 2007; Brockhurst et al., 2014) are characterized by arms races. In fact, most societies with reproductive skew are the result of parasitic exploitation. In this regard, the reproductive skew societies are the counterparts of unitary organisms with their soma-germline dichotomy and conflict (Heininger, 2002, 2012). As discussed earlier (chapter 3) a fundamental difficulty with the evaluation of acts of fitness transfer is that the outside observer can only observe the outcome but has no knowledge of the specific motives and mechanisms. Basically, a net transfer of commodities can take place due to voluntary intent, cheating, coercion, or be the result of a lottery. For example, altruistic acts involve the actor voluntarily donating fitness to beneficiaries. Parasitic acts, on the other hand, involve the actor extracting benefit from others at net cost to the donors (Doncaster et al., 2013). Both behaviors may have the same direct net-cost transferral of fitness from donor to beneficiary; the key difference between parasitism and altruism is thus who drives the interaction.

Taking the society of the Polynesian Ifaluk Atoll as example, Laura Betzig (2004) succinctly highlighted the interrelationship between limited dispersal and asymmetric conflict/dominance hierarchies underlying coerced cooperation in human societies: “Not so long before I first went to Ifaluk, in 1970, the American Museum of Natural History anthropologist Robert Carneiro published a paper in Science on the origin of the state (Carneiro, 1970). He looked at state formations in Mesopotamia, Egypt, and Peru; and he concluded that they were all “circumscribed” zones. Poor men paid Sumerian/Babylonian/Assyrian emperors, Egypt’s pharaohs, and Peru’s Incas in tribute and labor not because emperors/pharaohs/Incas were good to them in return, but because the costs of leaving were high. The rich land between the Tigris and Euphrates, or around the Nile Delta, or in the Andes valleys was surrounded by hostile deserts and mountains. Dissatisfied subjects had two options. They could pay overlords what they asked for, or they could vote with their feet and hope for the best.
“Skew” theories take those two options into account. Studies of animal societies have looked for “social contracts” – the equal return of social benefits for social benefits. And they’ve looked for “social controls” – the biased return of social benefits to better fighters. Evidence of the first is relatively equivocal (e.g., Emlen et al., 1998). Evidence of the second is relatively clear (e.g., Clutton-Brock, 1998). Better fighters do best where worse fighters are trapped: where the costs of running away to another good territory are high.
I think the whole of human history can be interpreted like that. The Sumerian word for “freedom,” ama.ar,gi, also means “freedom to move” (e.g., Lemche, 1979). Mobility makes equality; and inequality goes up where subjects can’t get away. Around 5,000 years ago, in fertile river valleys bordered by mountains and deserts, subjects started to pay overlords labor and taxes. They stopped, as soon as they found a way out. In the wide open spaces of Africa and Asia, people had voted with their feet for millions of years.
[…] sometimes they [i.e., foragers, KH] give social benefits away to stay on a good territory. They give up part of the hunt/catch/crop because the hunting/fishing/gathering isn’t as good anywhere else. It’s a little ironic, to me, that so many of the new evolutionists – the “Darwinian” psychologists and anthropologists – focus so much on cooperation and so little on competition. Where’s the Darwinism? Where’s the beef?
In Homo sapiens societies, as in any other animal societies, mobility is an aid to equality. Foragers are notoriously “egalitarian,” speaking relatively. But no society lacks unfairness completely. Strong egalitarian ethics, like “we refuse one who boasts, for someday his pride will make him kill somebody” (Lee, 1979), or “sell all you have and distribute to the poor” (Luke, 18:22), or “from each according to his ability, to each according to his needs!” (Marx, 1875/1970) aren’t repeated where cooperation is automatic. They get said, again and again, where individuals conflict, and where the winners take more than equal shares. Why do they get to take more? Sometimes, because the givers have nowhere better to go.” (Betzig, 2004)

9.1 Reproductive suppression in vertebrates

Social suppression of reproduction is a salient feature of cooperatively breeding species that is maintained by natural selection (Solomon & French, 1997; Koenig & Dickinson, 2004). Subordinates in cooperative societies commonly show a degree of reproductive restraint due to factors such as a lack of access to unrelated breeding partners, poor body condition, or underdeveloped foraging skills that reduce their expected payoff from attempted reproduction (Wasser & Barash, 1983; Heinsohn, 1991; Bennett et al., 1996; O’Riain et al., 2000). In other cooperatively breeding species dominants render their subordinates infertile by targeted aggression, temporary eviction or infanticide (Keverne et al., 1982; Reyer et al., 1986; Schoech et al., 1991; Wingfield et al., 1991; Solomon & French, 1997; O’Riain et al., 2000; Packard, 2003; Young et al., 2006; Cant et al., 2010). The way in which suppression is imposed at a proximate level can be direct, via interference, or harassment/aggression (e.g. Dunbar & Dunbar, 1977; Blaffer Hrdy, 1979; Silk et al., 1981a; Curry, 1988b; Rhine et al., 1988; Wasser & Starling, 1988; Faulkes & Abbott, 1997; Lacey & Sherman, 1997; Clutton-Brock et al., 1998; Clarke & Faulkes, 2001; Hackländer et al., 2003; Williams, 2004; Young et al., 2006; Stockley & Bro-Jørgensen, 2011), or indirect, via signals that communicate social dominance and can influence the reproductive and life-history decisions of young or subordinate females (e.g. Epple & Katz, 1984; Savage et al, 1988; Barrett et al., 1993; Saltzman et al., 2009). Such stimuli induce responses at different stages of the reproductive cycle from inhibition of mating (Stockley & Bro-Jørgensen, 2011) to delayed sexual maturity, disruption of ovulation, implantation, or spontaneous abortion (e.g. Bowman et al., 1978; Huck et al., 1983, 1988; Adams et al., 1985; Harcourt, 1987; Creel et al., 1992; Faulkes & Abbott, 1997; Solomon et al., 2001; Hackländer et al., 2003; Saltzman et al., 2006, 2009).

Suppression of subordinate reproduction by dominants and high levels of reproductive skew are often found in species that have evolved in arid or semiarid conditions, where competition for resources is likely to be intense and reproduction by subordinates is likely to be particularly costly to dominants (Clutton-Brock et al. 2006; Gilchrist, 2006). While infanticide by females has attracted less attention than infanticide by males, it is probably more widespread (Parmigiani & vom Saal, 1994; Rödel et al., 2008; Clutton-Brock & Huchard, 2013) and frequently represents a threat for group-living females (Digby, 2000). Infanticide of the offspring of other breeders has been recorded among a wide array of taxa, including many cooperative breeders (Blaffer Hrdy, 1979; Hausfater & Blaffer Hrdy, 1984; Hoogland, 1985; Pusey & Packer, 1987a; Brain, 1992; Creel et al., 1997; Clutton-Brock et al., 1998; Ebensperger, 1998; Young & Clutton-Brock, 2006; Clutton-Brock & Huchard, 2013). In cooperative breeders, like meerkats and dwarf mongooses, subordinate litters are often killed by the dominant female, and the success of subordinate litters is affected by the timing of parturition. Simultaneous breeding by more than one female reduces the ratio of helpers to pups and the growth of pups falls (Clutton-Brock et al., 2010) and evidence that infanticide is more likely in pregnant than non-pregnant females (Rasa, 1994; Creel & Waser, 1997; Clutton-Brock et al. 1998, 2001b) suggests that its function is partly to reduce resource competition for the killer’s offspring (Wolff & Cicirello, 1989; Tuomi et al., 1997; Clutton-Brock et al., 1998; Rödel et al., 2008). It may have additional benefits: victims of infanticide may subsequently contribute to suckling and rearing infants subsequently produced by infanticidal females as in marmosets (Digby, 1995) and meerkats (Clutton-Brock et al., 1998). Similarly, egg tossing or eating has been observed in communal laying bird species, in which the last female to start laying usually tosses the eggs already laid from the nest (Mumme et al., 1983; Vehrencamp et al., 1988; Zahavi, 1990; Koenig et al., 1995; Macedo & Bianchi, 1997; Macedo et al., 2001).

Dominant efforts at suppression are not inevitable, instead appearing to be sensitive to variation in the payoffs of interfering with subordinate breeding: attacks are targeted at subordinates who are most likely to breed (Clutton-Brock et al., 2010); are restricted to periods when resource competition peaks and the offspring of dominants may be at a competitive disadvantage (Young & Clutton-Brock, 2006; Clutton-Brock et al., 2010); or are avoided entirely when the subordinate retaliation is likely to be effective (Packer et al., 2001). In a substantial number of social mammals, competition between resident females leads to evictions or to groups splitting (Stephens et al., 2005; Cant et al., 2010; Clutton-Brock & Huchard, 2013). For example, in meerkats, dominant females evict (virtually) all female subordinates before they are 4 years old (Clutton-Brock et al., 2010). Increased levels of glucocorticoids are known to repress the reproductive axis (Wingfield & Sapolsky, 2003; Kalantaridou et al., 2004; Gore et al., 2006; Heininger, 2013). While evicted, subordinate females suffer chronic elevation of their glucocorticoid adrenal hormone levels, reproductive downregulation (reduced pituitary sensitivity to gonadotropin-releasing hormone), reduced conception rates, and increased abortion rates (Young et al., 2006). In red howler monkeys, high-ranking females frequently evict younger and lower ranking females from their groups (Pope, 2000) while, in banded mongooses, coalitions of older dominant females intermittently evict entire cohorts of younger females from their group (Gilchrist, 2006; Cant, 2010). When subordinates are experimentally prevented from breeding using contraceptive injections, dominants in groups of wild meerkats (Suricata suricatta) are less aggressive towards subordinates and evict them less often, leading to a higher ratio of helpers to dependent pups, and increased provisioning of the dominant’s pups by subordinate females (Bell et al., 2014). Dominants also showed improved foraging efficiency, gained more weight during pregnancy and produced heavier pups, which grew faster. These results confirmed the benefits of suppression to dominants, and help explain the evolution of singular breeding in vertebrate societies (Bell et al., 2014). The results also suggest why plural breeding is rare in cooperative vertebrates: dominants are only likely to tolerate subordinate reproduction when it has little effect on dominant reproductive success, which is only likely where social structure limits direct competition between offspring (Bell et al., 2014). Banded mongooses, a closely related social mongoose, are one of the few cooperative vertebrates where multiple females commonly breed together (Cant et al., 2010): direct competition between pups is limited because pups are cared for by a single helper who does not provision other pups (Bell, 2007).

Subordinates may try to seize the dominant position by fighting (Davies, 1992; Gust, 1995; Monnin & Peeters, 1999; Alberts et al., 2003; Cant et al., 2006a). The stability of the dominant–subordinate relationship therefore requires that subordinates prefer to queue peacefully rather than engage in escalated conflict over rank. To understand the factors influencing the subordinate’s decision of whether to enter into an escalated fight over rank, three features of the conflict between dominant and subordinate have been taken into account: (i) the contestants may be relatives; (ii) subordinates may eventually obtain the resource without fighting, i.e. they may inherit if they outlive the dominant; and (iii) subordinates may share the resource with the dominant without fighting, i.e. they may obtain a share of current reproduction. Skubic et al. (2004) showed that a male Neolamprologus pulcher helper cichlid should base its decision to parasitize primarily on an increase in expulsion risk resulting from reproductive parasitism (punishment), intra-group relatedness and the parasitism capacity. If expulsion risk is high then helpers should not parasitize reproduction at medium body size but should parasitize either when small or large. As far as can be juged from a limited data base, in paper wasps that do not distinguish degrees of relatedness among nestmates (Queller et al., 1990), genetic relatedness had no significant effects on the probability of escalation (Cant et al., 2006a). Since reproductively suppressed subordinates are more likely to fight for dominant status, attempts by dominants to reproductively suppress subordinates will increase the threat of escalated conflict. Overall, subordinates that stand to gain more from a reversal in the dominance roles were more likely to enter into the escalated conflict (Cant et al., 2006a). To avoid escalated conflicts, dominants may appease challengers by making reproductive concessions (Reeve & Ratnieks, 1993; Field & Cant, 2009).

In singular cooperative breeders, the death of the breeding female is often followed by intense fighting between her daughters and the death or eviction of unsuccessful competitors (Clutton-Brock et al., 2006; Sharp & Clutton-Brock, 2011). As soon as a subordinate animal rises to a dominant position in its cohort, the previous reproductive “altruist” (according to kin selection theory) turns into a selfish reproductive dictator (Faulkes & Abbott, 1997). Female subordinates in the group-living Lake Tanganyika cichlid N. pulcher are reproductively capable, but apparently suppressed with respect to egg laying. Nevertheless, some reproduction is tolerated, possibly to ensure continued alloparental care by subordinate females. Upon removal of the dominant female from the group, the medium females immediately seize the dominant breeding position and start to reproduce as frequently as control pairs (Heg, 2008). However, the non-reproductive worker and the reproductive queen are both products of the same genotype, and their phenotypic differences result from differences in environment rather than genotype. That previously “altruistic” subordinates become selfish dominants after the death of the breeding female can be regarded as due to pleiotropic plasticity (see chapter 9.3), the plastic phenotypic expressions of a single genotype (Yakubu, 2012, 2013). Likewise, field studies going back to the 1830s observed that the same honeybee egg could produce a selfish queen or an “altruistic” worker depending upon what diet the larva was fed (Prete, 1990).

9.2 Reproductive suppression in eusocial insects

Eusociality as the Holy Grail of inclusive fitness theory lost its gloriole when it became evident that the unselfish help of the workers is enforced (Ratnieks, 1988). (To remind you, the prediction of the inclusive fitness theory is that this help should be voluntary [given a sufficient benefit to cost ratio] since helping your kin should increase the representation of your genes in future generations and this should be a strong inclusive fitness benefit selecting for the reproductive “altruism”) In fact, beehives, wasp and ant nests are police states (Foster & Ratnieks, 2001; Monnin & Ratnieks, 2001; Cuvillier-Hot et al., 2004b) in which selfish and despotic queens, by means of their pheromones suppress the reproductive activity of their daughters and enforce their reproductive monopoly by murder, torture and imprisonment (Whitfield, 2002). Workers are enslaved, made docile and submissive by force and chemicals. In addition to queen pheromones, worker sterility is thought to be enforced by mutual policing. Worker policing occurs in two forms: (i) physical policing, in which workers with activated ovaries are attacked by nestmates (Hölldobler & Carlin, 1989; Liebig et al., 1999; Dietemann et al., 2003; Hartmann et al., 2003), and (ii) egg policing, in which workers detect and destroy worker-laid eggs (Ratnieks & Visscher, 1989; Monnin & Peeters, 1997; Foster & Ratnieks, 2000; Tsuchida et al., 2003; d’Ettorre et al., 2004; Endler et al., 2004; Helanterä & Sundström, 2005). The importance of policing in maintaining worker sterility has been demonstrated repeatedly (Foster & Ratnieks, 2001; Wenseleers et al., 2004a, b; Ratnieks & Wenseleers, 2005, 2008; Ratnieks & Helanterä, 2009). Long before Machiavelli, queens of eusocial societies “discovered” the principle of “divide and rule”. (No surprise in the light of Orgel’s second rule: “Evolution is cleverer than you are”.)

Inclusive fitness theory predicts that, other things equal, worker policing occurs when workers are more closely related to queen-produced males than to worker-produced males (Starr, 1984; Woyciechowski & Lomnicki, 1987; Ratnieks, 1988). Worker policing should be a classic kin-selected trait (Hamilton, 1964) that is favored when relatedness is low (Ratnieks, 1988; Frank, 2003; Wenseleers & Ratnieks, 2006a), due to relatedness differences arising from either polyandry (multiple mating by queens) or polygyny (multiple queens per colony). However, worker policing also occurs, according to comparative analyses, across the eusocial Hymenoptera as a whole (Wenseleers & Ratnieks, 2006a) and even in species where queen mating frequency is below 2: for example, in the ant Camponotus floridanus (Endler et al., 2004), the bumblebee Bombus terrestris (Zanette et al., 2012) and the hornet Vespa crabro (Foster et al., 2002), or in species that reproduce parthenogenetically, such as the ant Platythyrea punctata (Hartmann et al., 2003) and the Cape honeybee Apis mellifera capensis (Pirk et al., 2003). This has led to doubts concerning the importance of relatedness in selecting for worker policing (Foster et al., 2002; Hartmann et al., 2003; Pirk et al., 2003, 2004; Endler et al., 2004; Gadagkar, 2004a; Wilson, 2005). This doubt was supported by Hammond and Keller (2004), who used a comparative analysis of 50 species of ants, bees, and wasps to show that the proportion of males that are workers’ sons is not influenced by relatedness.

Punishment allows the evolution of cooperation (or anything else) in sizable groups (Boyd & Richerson, 1992; Clutton-Brock & Parker, 1995; Frank, 2003; Ratnieks & Wenseleers, 2008; Bourke, 2011; Shutters, 2012) and also enforced cooperation in an experimental setting (Fehr & Gächter, 2002). Who would call this state of coerced helping “altruism”? Advocates of Hamilton’s rule do it. George Orwell’s “1984” has made a home in the heads of the kin selection community. “Worker policing is a mechanism by which a society resolves its conflicts,” says Ratnieks. “I think it’s the best example of conflict resolution in nature” (Whitfield, 2002). Obviously this “conflict resolution” works well for Hymenopteran societies as is demonstrated by the anarchy that is inflicted upon these societies by workers that do no longer respond adequately to queen pheromones, establishing a type of social parasitism (Oldroyd et al., 1994; Barron & Oldroyd, 2001; Barron et al., 2001; Martin et al., 2002; Wossler, 2002; Lopez-Vaamonde et al., 2004; Hoover et al., 2005a, b; Beekman & Oldroyd, 2008; Dobata & Tsuji, 2009). But on the other hand, this forceful conflict resolution undermines the argumentative power of the kin selection theory that pretends to explain the occurrence of cooperation and “altruism” solely on the basis of inclusive fitness gains of “altruists”: a female would be able to pursue her genetic interests and increase her inclusive fitness more effectively by raising her mother’s daughters rather than her own. In a less dogmatic and biased scientific climate the findings of “enforced altruism” (chapter 20) should have refuted the kin selection/inclusive fitness theory long ago.

Overt aggression by the queen is very rare and appears restricted to small colonies (Franks & Scovell; 1983; Kikuta & Tsuji, 1999; Wenseleers et al., 2005; Brunner & Heinze, 2009). In larger colonies the queen is unable to control a larger number of workers by agonistic acts and reproduction appears to be controlled chemically by glandular or cuticular pheromones (Monnin, 2006; Heinze & d’Ettorre, 2009). Queen pheromones have been postulated to be either a manipulation that is detrimental to workers (‘queen control’) or a signal to which workers are selected to respond (‘queen signal’). It remains controversial whether these agents are manipulative, i.e., actively suppress worker reproduction, or honestly signal the fertility status of the queen to which workers react in their own interest by refraining from laying eggs (West Eberhard, 1981; Seeley 1985, p. 30; Woyciechowski & Lomnicki, 1987; Bourke, 1988a; Keller & Nonacs, 1993; Foster et al., 2000; Katzav-Gozansky et al., 2003, 2004; Alaux et al., 2004; Hefetz & Katzav-Gozansky, 2004; Tóth et al., 2004; Katzav-Gozansky, 2006; Conte & Hefetz, 2008; Heinze & d’Ettorre, 2009; Brunner et al., 2011; Tan et al., 2012). Manipulative queen control is thought to lead to an evolutionary arms race between queens and workers, resulting in complex queen bouquets that diverge strongly among different populations and species. In contrast, honest signals should evolve more slowly and might therefore differ less strongly within and among species. It has been argued that pheromonal manipulation of workers by the queen should not be evolutionarily stable since workers should be selected to ignore such signals (Keller & Nonacs, 1993). However, arms races would allow the queens to stay a step ahead of the workers, particularly if changing the chemical composition of pheromones is not too costly (West Eberhard, 1981; Foster et al., 2000; Hefetz & Katzav-Gozansky, 2004; Katzav-Gozansky, 2006). Katzav-Gozansky (2006) elaborated features of pheromone cocktails that argue for an arms race: (i) the evolution of multiple pheromonal sources, when theoretically one set should have sufficed; (ii) the evolution of a complex blend of pheromones in the first place (Hefetz & Katzav-Gozansky, 2004); (iii) escalation in the arms race whereby workers have become not only insensitive to the queen pheromones (Hoover et al., 2005b) but are also somehow able to camouflage their eggs in such a way that they are not recognized as worker eggs by their worker nestmates (Beekman & Oldroyd, 2003). According to Tóth et al. (2004), the high variability in worker reproduction in stingless bee species is consistent with the expectation of an evolutionary arms race in which worker and queen control change with time. Tan et al. (2012) also provided evidence for an ongoing evolutionary arms race for reproductive dominance in Apis cerana. Brunner et al. (2011) found that cuticular hydrocarbons associated with fecundity evolve slightly faster than worker-specific components in the blend of cuticular hydrocarbons which might reflect an arms race between queens and workers.

The manipulation model (Crespi & Ragsdale, 2000) assumes that subordinates cannot avoid dominant-imposed costs, when subordinates should often be under selection to escape manipulation by dominants. If manipulative acts pre-empt escape (e.g. feeding juveniles less), then such selection cannot be effective. In contrast, if manipulation involves ongoing behavioral interactions, such as frequent aggressive nudging or enforced displays of subordinate status, then subordinates might be expected to have the option of avoiding the interactions as best they can and perhaps thereby improving their dispersal prospects. In many social insect species, subordinates attempt to avoid the dominant while at the nest site (Buckle, 1982; Michener, 1990), a behavior that is compatible with the “queen control” but not with the “queen signal” concept. This behavior could be interpreted as avoidance of reproductive suppression, aggressive manipulation or both. Moreover, in some halictine bees, some first-brood offspring of the foundress avoid her by entering diapause directly and becoming next year’s foundresses rather than this year’s workers (Yanega, 1988). Has such early diapause evolved in part as a means of escaping maternal manipulation? In Halictus rubicundus, directly diapausing, first-brood offspring are larger than sibs which stay at home to help (Yanega, 1989), which is consistent with the idea that small worker size tilts behavioral choices towards helping. In other halictines, most first-brood offspring stay and help, while some leave and attempt to breed independently in the same year (Stockhammer, 1967; Sakagami & Hayashida, 1968; Sakagami, 1977); the presence of such species shows that the strategies envisioned by the manipulation model (Crespi & Ragsdale, 2000) are realistic and can be investigated in natural populations. There is a continuity of aggression-dependent and chemical cue-dependent reproductive suppression of workers in primitively eusocial species. The effects of queen pheromones on worker ovary degeneration (see chapter 11.2) should be clear-cut evidence for queen suppression of worker reproduction. Moreover, the trans-species effect of queen pheromones on Drosophila ovary development is not compatible with the argument that queen pheromones are a queen signal to which workers are selected to respond (obviously Drosophila cannot have been selected to respond). Finally, pheromones as queen-produced agents have their manipulative counterpart in the reproductive suppression of helpers enforced by dominants in vertebrate eusocial communities (see chapter 9.1).

In primitively eusocial insect species, suppression of ovarian development in workers is frequently a result of food restriction or aggressive behavior (i.e., nudging, butting) by the principal egg-layer (West, 1967; Free et al., 1969; Wheeler, 1986; Michener, 1990; Röseler & Van Honk, 1990). In primitively eusocial Hymenoptera, aggression is one of the most distinct behavioral attributes of the queen and may serve to establish dominance hierarchies in small colonies (Michener & Brothers, 1974; Breed & Gamboa, 1977; Michener, 1990; Pabalan et al., 2000). Workers may still escape dominance assertion by the queen by avoiding her and workers may lay eggs (Buckle, 1982). However, differential oophagy by the primary reproductive may prevent successful development of worker-laid eggs (Brothers & Michener, 1974; Packer & Owen, 1994). This egg-eating activity appears to bolster the queen’s reproductive dominance (Kukuk, 1992). In Polistes wasps and permanently queenless ants, aggression by the dominant reproductive individual limits both mating opportunities and reproduction of subordinates (Reeve, 1991; Röseler, 1991; Peeters, 1993; Premnath et al., 1996; Monnin & Peeters, 1998). However, after reproductive dominance is established, a reduction in aggression levels often is observed, which appears to be replaced by chemical cues as suppressive agents, at least in Ropalidia marginataPolistes dominulus (Premnath et al., 1996; Sledge & Boscaro, 2001) and Bombus terrestris (Fletcher & Ross, 1985; Röseler & Van Honk, 1990; Sramkova et al., 2008; Amsalem & Hefetz, 2010). Thus, chemical cues can be regarded as surrogate for agonistic behavior. Candidates for such cues are cuticular hydrocarbons that are altered by ovarian activity (Cuvillier-Hot et al., 2001; Liebig et al., 2009; Peeters & Liebig, 2009; Ferveur & Cobb, 2010; Liebig, 2010; Fedina et al., 2012). In advanced eusocial insect colonies (i.e., those with extreme queen–worker dimorphism), queen-produced pheromones that maintain worker sterility are thought to be taxonomically widespread, as queens, their eggs and queen-derived chemicals have been shown to reduce or eliminate worker reproduction, and because queens typically produce chemicals that are absent or minimally expressed in workers (e.g. Wilson, 1971; Michener, 1974, 1990; Fletcher & Ross, 1985; Wheeler, 1986; Bourke, 1988a; Vargo, 1992; Peeters et al., 1999; Dietemann et al., 2003; Cuvillier-Hot et al., 2004a; Endler et al., 2004; Monnin, 2006; Korb et al., 2009; Bhadra et al. 2010; Holman et al., 2010b; Kocher & Grozinger, 2011).

Queen pheromones are strikingly conserved across at least three independent origins of eusociality, with wasps, ants, and some bees all appearing to use nonvolatile, saturated hydrocarbons to advertise fecundity and/or suppress worker reproduction (Van Oystaeyen et al., 2014). Saturated hydrocarbons were most likely used as fertility cues in the common solitary ancestor of all ants, bees, and wasps, which lived ~145 million years ago (Johnson et al., 2013; Wilson et al., 2013; Van Oystaeyen et al., 2014). Queen pheromones underpin the proximate and ultimate causes of worker sterility: in the honeybee, they cause changes in worker gene expression (Grozinger et al., 2003; Beggs et al., 2007) and physiology (Kaatz et al., 1992; Beggs et al., 2007) and mediate the transition from reproductive suppression to worker reproduction (Wossler & Crewe, 1999; Hoover et al., 2003; Kocher et al., 2008, 2009; Maisonnasse et al., 2010). Moreover, honey and bumble bee queen pheromones inhibit rearing of new queens by workers (Butler et al., 1959; Melathopoulos et al., 1996; Pettis et al., 1997; Lopez-Vaamonde et al., 2007), delay the transition from nursing to foraging (Pankiw et al., 1998), inhibit juvenile hormone synthesis in workers (Pankiw et al., 1998), and influence comb building within the colony (Ledoux et al., 2001). Notably, reproductively active, queenless honeybee workers (and bumble bee workers) revert back to sterility when the queen is reintroduced (Alaux et al., 2007; Malka et al., 2007). Inhibitory queen pheromones and their active compounds have also been found in the ants Camponotus floridanus (Endler et al., 2004), pharaoh's ant Monomorium pharaonis (Edwards, 1987), Aphaenogaster senilis (Boulay et al., 2009), and Lasius niger (Holman et al., 2010b), and the termite Reticulitermes speratus (Matsuura et al., 2010). Among ants, the queen pheromone system of the fire ant Solenopsis invicta is particularly well studied. Pheromones initiate reproductive development in new winged females, called female sexuals (Vargo, 1999). These chemicals also inhibit workers from rearing male and female sexuals, suppress egg production in other queens of multiple queen colonies and cause workers to kill sexual larvae or execute excess queens (Fletcher & Ross, 1985; Vargo & Fletcher, 1986; Vargo, 1999; Klobuchar & Deslippe, 2002). In queenless colonies that lack such pheromones, winged females will quickly shed their wings, develop ovaries and lay eggs. These virgin replacement queens assume the role of the queen and even start to produce queen pheromones (Vargo, 1999). There is also evidence that queen weaver ants Oecophylla longinoda have a variety of exocrine glands that produce pheromones, which prevent workers from laying reproductive eggs (Fletcher & Ross, 1985).

9.3 Excursion: Pleiotropy

Haldane (1955) explained that “two genes are said to be allelomorphic if a nucleus with a single chromosome set, for example that of a spermatozoon, can only contain one of the two.” Even where both the alleles are present in a single diploid organism (heterozygote), only one (the dominant allele) is expressed. In incomplete dominance, there is usually a blended simultaneous expression of the morphological phenotypes. Selfish mutants that cooperate less (cheats) can be isolated from natural populations of microorganisms or artificially generated (Velicer, 2003; West et al., 2006; Foster et al., 2007; Ross-Gillespie et al., 2009). Accordingly, extant models of kin selection suppose that the "altruistic" and selfish traits are allelomorphs. There is never environmentally contingent expression of one or the other in an alternating fashion. Where the latter happens, in modern genetics parlance this is described as phenotypic plasticity (Pigliucci, 2001). Recently, Yakubu (2012, 2013) postulated a pleiotropic plastic “sociality trait” that can express either "altruism" or selfishness depending upon the circumstances. This trait requires the capacity to judge when to be "altruistic" or behave selfishly. This suggests plasticity.

The theory of pleiotropy assumes that a particular gene may have an effect not only on one feature but on several traits of an organism. Pleiotropy is one of the most commonly observed attributes of genes, with broad implications in genetics, evolution, development, aging, disease, and drug discovery (Wright, 1968; Barton, 1990; Cheverud, 1996; Hodgkin, 1998; Waxman & Peck, 1998; Brunner & van Driel, 2004; Otto 2004; van de Peppel & Holstege, 2005; He & Zhang, 2006; Stearns, 2010; Wang et al., 2010). Pleiotropy causes compromises among adaptations of different traits, because a genetic change beneficial to one trait may be deleterious to another (Barton, 1990; Otto, 2004). Hence, pleiotropy is under strong stabilizing selection (Barton, 1990). A model showed that, when three or more characters are affected by a mutation, a single optimal genetic sequence may become common. This result may explain the low levels of variation and low rates of substitution that are observed at some loci (Waxman & Peck, 1998). Pleiotropy is expected to constrain the rate of evolution (Otto, 2004), consistent with the observation that broadly expressed genes evolve more slowly (Hughes & Hughes, 1995; Hastings, 1996; Hurst & Smith, 1999; Duret & Mouchiroud, 2000; Hirsh & Fraser, 2001; Jordan et al., 2002; Subramanian & Kumar, 2004; Zhang & Li, 2004; Wall et al., 2005; Zhang & He, 2005; Liao & Zhang, 2006; Liao et al., 2006; Larracuente et al., 2008). Importantly, the expression pattern-dependent substitution rates in mammalian genes are determined by selection intensity but not mutation rate (Duret & Mouchiroud, 2000). He and Zhang (2006) found that 39.5 ± 0.8% of nonpleiotropic yeast genes (no phenotype in any condition) have detectable homologs in the fruit fly D. melanogaster. In comparison, 49.2 ± 2.3% of low pleiotropic genes (with phenotypes in one to two conditions) and 54.7 ± 3.6% of high pleiotropic genes (with phenotypes in more than two conditions) have fruit fly homologs. Pleiotropic genes are significantly more likely to be retained in long-term evolution than nonpleiotropic genes (χ = 29, P< 10-7).

Social behavior is a plastic trait that is regulated by environmental cues such as stress, and various other, particularly metabolic (see chapter 10), signaling networks. In eusocial insects, environmental effects on worker ovary activation are obvious, as pheromones emitted from the queen (Voogd, 1956; Hoover et al., 2003) and her brood (Jay, 1972; Mohammedi et al., 1998) strongly suppress worker reproduction. Queen pheromones are pleiotropic agents that induce “altruistic” sterility in workers and selfish monopolization of reproduction in queens (Wossler & Crewe, 1999; Hoover et al., 2003; Kocher et al., 2008, 2009; Holman et al., 2010b; Maisonnasse et al., 2010). Within queenright colonies, worker reproduction is rare (Ratnieks, 1993) but in queenless colonies, some workers activate their ovaries and begin to lay unfertilized eggs that develop into haploid males (Velthuis, 1970; Page & Robinson, 1994). Worker reproduction is therefore responsive to pheromone signals, and variation in ovary activation among workers is best modelled as a threshold response (Barron & Robinson, 2008). Beyond pheromones, environmental factors such as nutrition (Jay & Jay, 1993; Pernal & Currie, 2000), season (Lin & Winston, 1998; Hoover et al., 2006) and the presence and qualities of other workers within the colony (Wegener & Bienefeld, 2009) can likewise influence worker reproductive state. In termite colonies, workers are reproductively suppressed in queenright colonies and accordingly to current terminology are considered as “altruists”. When a breeder’s position becomes available, a subset of workers develops into neotenic replacement reproductives. Should more individuals develop than are needed to replace the queen or king, these neotenics will fight with their siblings until death to take over the queenless colony. During the process of caste determination, genetically identical individuals develop into queens or workers (LeBoeuf et al., 2013), the former selfish and the latter “altruistic”. In singular cooperative breeders, the death of the breeding female is often followed by intense fighting between her daughters and the death or eviction of unsuccessful competitors (Clutton-Brock et al., 2006; Sharp & Clutton-Brock, 2011). In the primitive ant species, Harpegnathos saltator, queens are relatively short-lived, and the loss of a queen reliably triggers highly aggressive, long-lasting (up to 4 weeks) dominance fights among the remaining workers (Liebig et al., 2000; Peeters et al., 2002).As a result of these dominance fights, only one or a few inseminated workers eventually become reproductive and turn into functional queens. These examples demonstrate the pleiotropy, context dependency of alternative selfish and “altruistic” behavior.

Genetic effects on worker ovary activation also occur, as is apparent from colony patrilines that show variation in their threshold response to pheromone signals. In such cases, workers sired by a particular drone are more likely to activate their ovaries in queenless colonies (Inbar et al., 2008), or even when a queen is present—a condition described as worker ‘anarchy’ (Barron et al., 2001). A meta-analysis of published studies that measure worker ovary activation as a function of genetic or environmental manipulations, revealed that environmental sources of variation, including real or synthetic queen pheromone, brood pheromone, social interaction, season, diet and CO2 narcosis, have had a larger overall effect on ovary scores than have strictly genetic effects (Backx et al., 2012).

Epigenetics is closely linked to environmental conditions and mitochondrial bioenergetics (Xie et al., 2007; Naviaux, 2008; Smiraglia et al., 2008; Wallace & Fan, 2010; Minocherhomji et al., 2012). The epigenome provides the interface between the environment and the regulation of nuclearDNA gene expression (Feinberg, 2007, 2008). 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 & Fan, 2010; Cyr & Domann, 2011). Stress-induced changes of DNA methylation are common (Chinnusamy & Zhu, 2009; Verhoeven et al., 2010; Richards, 2011; Verhoeven & van Gurp, 2012). While mitochondrial DNA replication appears to be required for prespore differentiation in Dictyostelium (Araki & Maeda, 1995, 1998; Shaulsky & Loomis, 1995), pleiotropic DIF-1 that is released by prespore cells (Kay & Thompson, 2001) suppresses prestalk cell energy metabolism. DIF-1 and DIF-3, the first metabolite produced during the degradation of DIF-1, are mitochondrial uncoupler that inhibit mitochondrial respiration and suppress cell growth (Shaulsky & Loomis, 1995; Thompson & Kay, 2000b; Kubohara et al., 2013). Honeybee worker development is associated with increased DNA methylation; larvae reared in vitro with inhibited DNA methylation show greatly increased probabilities of developing into queens (Kucharski et al., 2008).

Stress-dependent bacterial quorum sensing is a pleiotropic regulatory network of social interactions (Czárán & Hoekstra, 2009, 2010) that regulates both cooperative biofilm formation (Suntharalingam & Cvitkovitch, 2005; Irie & Parsek, 2008; Goo et al., 2012; Li & Tian, 2012) and selfish bacteriocin production (Kleerebezem & Quadri, 2001; van der Ploeg, 2005; Cornforth & Foster, 2013) (see chapter 5.3). As in many other bacteria, Pseudomonas aeruginosa has a Fur protein that functions as a global regulator of iron-responsive genes (Vasil & Ochsner, 1999; Ochsner et al., 2002). Since a critical level of intracellular iron serves as the signal for biofilm development, Fur is a key regulator of biofilm maturation (Banin et al., 2005). In P. aeruginosafur appears to be an essential gene, and the only fur mutants available either produce reduced levels of wild-type Fur or have missense mutations that exhibit high reversion rates (Barton et al., 1996; Ochsner et al., 1999; Banin et al., 2005). Thus, pyoverdine defective mutants, which have been isolated in natural populations, are potential cheats (De Vos et al., 2001; Visca et al., 2007). Clones that defected from cooperative production of iron-scavenging siderophores were deficient in biofilm formation (Banin et al., 2005; Harrison & Buckling, 2009). Thus fur has a pleiotropic role in iron homeostasis and social behavior of P. aeruginosa (Banin et al., 2005; Harrison & Buckling, 2009). Iron metabolism is regulated in response to oxidative stress coordinated with oxidative stress defenses (Pantopoulos & Hentze, 1995; Hanson & 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) and modulate social behavior. Myxococcus fulvus develops by social fruiting body formation in social habitats, but switches pleiotropically to asocial spore formation from vegetative cells in asocial habitats (Zhang et al., 2005; Li et al., 2011). Deleting the gene dimA in D. discoideum allows cells to avoid death, but leads to a great reduction in spore production (Foster et al. 2004). Hence, the pleiotropic effects of dimA stabilize social behavior among amoeba. Given the key role of the metabolism-stress signaling network for life history strategies and their trade-offs in multicellular organisms (Dallman et al., 1995; Dantzer & Swanson, 2012; Heininger, 2012; Sparkman et al., 2012), pleiotropic social interactions add multiple dimensions of complexity to this life history network (Sinervo et al., 2008).

Non-coding small RNAs that modulate gene expression at the post-transcriptional level are plausible candidates for the regulation of pleiotropic behavior (Altuvia et al., 1997). In bacteria, regulatory small RNAs (sRNAs) are often associated with stress responses in changing environments, such as oxidative stress, iron limitation, glucose-phosphate stress and growth-substrate deprivation (Altuvia et al., 1997; Masse & Gottesman, 2002; Vanderpool & Gottesman, 2004; Yu et al., 2010; Chen et al., 2014; Taylor, 2014), and have been shown to regulate social traits such as biofilm formation (Chambers & Sauer, 2013, Taylor, 2014), quorum sensing (Bejerano-Sagie & Xavier, 2007), adaptive resistance (Taylor, 2014) and fruiting body development (Yu et al., 2010). A non-coding small RNA molecule is a key developmental gatekeeper that blocks M. xanthus development when food is abundant (Yu et al., 2010; Chen et al., 2014).

Reactive oxygen/nitrogen species are prototypic pleiotropic agents. Candidate genes associated with variation in sensitivity to oxidative stress form networks of genes involved in development, immunity, and signal transduction (Jordan et al., 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 & Holbrook, 2000; Heininger, 2001; Mittler, 2002; Mikkelsen & Wardman, 2003; Sørensen et al., 2003; Apel & Hirt, 2004; Ardanaz & Pagano, 2006; Rollo, 2007; Miller et al., 2008; Slos & Stoks, 2008; Jaspers & Kangasjärvi, 2010; Steinberg, 2012; Choudhury et al., 2013). There is no doubt that reactive oxygen species 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 & Barrett, 2002; Martindale & Holbrook, 2002; Fruehauf & Meyskens, 2007; Moreira da Silva et al., 2010; Maryanovich & Gross, 2013).

10. Resources and social interactions:a tale of murder, affiliation and indifference


Summary
Behavioral ecologists have long recognized a relationship between social behavior and the distribution and predictability of resources. In extreme resource-limited environments both kin and conspecifics can serve as food. Filial cannibalism, siblicide and sibling cannibalism are a common phenomenon in several orders of taxa. Even matricidal behavior occurs in some taxa living in extremely resource-limited habitats.
In unpredictable or dangerous environments, the chances of successfully raising offspring independently are limited either by predator pressure or harsh environmental conditions. Under such conditions groups are better able to find scattered food, repel enemies, care for young, and thus reduce the risk of brood loss.
In social bacteria and amoeba, nutrient-rich environments select for asociality.An excessive abundance of common resources deters socially responsible actions on structured populations and the excess allows defectors to free-ride well over the time horizon that is required for cooperators to die out.A similar development may be observed in modern human societies with their abundant resources that appear to promote the loss of social coherence and solidarity.

Resource availability follows a conservation law, implying that the gains or losses in resources to individual are balanced by the losses or gains to others (Lehmann & Rousset, 2010). Behavioral ecologists have long recognized a relationship between social behavior and the distribution and predictability of resources (Crook, 1964; Horn, 1968; Bradbury & Vehrencamp, 1976; Clutton-Brock & Harvey, 1977), and increasing empirical evidence points to resource-based explanations of social organization in a variety of species (Waser, 1981; Macdonald, 1983; Wrangham et al., 1993; Chapman et al., 1994; Creel & Macdonald, 1995; Powell et al., 1997; Johnson DDP et al., 2002). In a simulation using an adaptive multiagent system, the effect of environmental conditions on the success of distinct strategies for optimal resource management was investigated. Interestingly, in an environment with patchy and limited resources agents learned to cooperate and to harvest their resources in a moderate way, thus avoiding population crashes due to uncontrolled exploitation (Krink, 2000). Resource supply has been found to alter the cost of cooperation, which decreases with increasing resource supply, such that public-goods cooperation is more likely to evolve when resources are more abundant (Brockhurst et al., 2007, 2008, 2010). The resources hypothesis of sociality suggests that social groups form and are maintained in response to the cooperative benefits that can be obtained by extracting certain types of resource distributions (Crook, 1965; Slobodchikoff, 1984). These groups may consist of either kin or nonkin. Following and extending Crook’s arguments (1965, 1970, 1972), Slobodchikoff (1984) suggested that ecological factors such as resource abundances and distributions play a role in determining whether or not a particular group of animals is going to be social (Slobodchikoff & Schulz, 1988). The relationship between resources and sociality is expressed in several different models (Crook, 1965, 1972; Wrangham, 1980, 1983; Terborgh, 1983; Van Schaik & van Hooff, 1983; Slobodchikoff, 1984). While research on non-humans demonstrates that the extent of nepotism among kin can depend critically on resources available to parents or sibships in insects (Griffin, West, and Buckling 2004; West et al., 2001) our understanding of the sensitivity of human kinship relations to resource competition derives largely from folklore.

10.1 Scarce resources and social behavior

In extreme resource-limited environments both kin and conspecifics can serve as food. Several theorists (Eickwort, 1973; Crespi, 1992; Mock & Parker, 1997; Pfennig, 1997; Perry & Roitberg, 2005) restated Hamilton’s rule for the spread of a selfish act to investigate the conditions favoring sibling cannibalism. For example, in a dyadic interaction a cannibal must increase survival by at least 50% to profit from consuming a full sibling. In my opinion this is teleological thinking par excellence (see also chapter 19). How should organisms, given the huge stochasticity of environments and their chances of survival to maturity (see chapter 15.1; Heininger, 2015), be able to estimate the inclusive-fitness effects of their actions? Already by taking frequency-dependent effects into account (information that is often not available to cannibals) the dynamic stochasticity of the social interactions and their fitness value become evident.

Cannibalism is a well-documented phenomenon that has been described for many animal species (Polis, 1981). Several forms of cannibalism can be distinguished (Rohwer, 1978; Smith & Reay, 1991), for example, heterocannibalism (also called ‘nonkin cannibalism’; Smith & Reay, 1991) and filial cannibalism. While the former describes the eating of unrelated conspecifics, the latter describes the consumption of own offspring, either eggs or young. Filial cannibalism is extremely common in teleost fish (e.g. Dominey & Blumer, 1984; Smith & Reay, 1991; FitzGerald, 1992), having been recorded in 17 families (Manica, 2002), and in a wide range of other taxa (Polis, 1981; Elgar and Crespi, 1992; Klug & Bonsall, 2007), including snakes (Lourdais et al., 2005), skinks (Huang, 2008), mammals (Elwood, 1991; Burn & Mason, 2008), birds (Gilbert et al., 2005), crustaceans (Dumont & Ali, 2004; Gallucci & Ólafsson, 2007), spiders (Anthony, 2003) and insects (Bartlett, 1987; Thomas & Manica, 2003; Creighton, 2005). Population size in the flour beetle Tribolium is mainly governed by the cannibalism rate rather than differences in fertility or fecundity (Park et al., 1961, 1965; McCauley & Wade, 1980; Stevens & Mertz, 1985). Several studies suggest that filial cannibalism increases when the costs of providing care increase or the benefits of providing care decrease (Marconato et al., 1993; Neff, 2003a, b; Thomas & Manica, 2003; Manica, 2004; Lourdais et al., 2005; Klug et al., 2006; Frommen et al., 2007; Gray et al., 2007; Huang, 2008; Chin-Baarstad et al., 2009). Population-level resource competition appears to play an important role in the evolution of both parental care and filial cannibalism (Klug & Bonsall, 2007).

Mock and Parker (1997) broadly defined sibling rivalry as “any features of animals or plants that have the effect of promoting individual survival and/or reproduction at the expense of siblings.” Siblicide (also termed brood reduction) falls within this definition (Hofer & East, 2008). The theory for resource-based sibling rivalry has been well developed (O’Connor, 1978; Stinson, 1979; Dickins & Clark, 1987; Parker et al., 1989; Godfray & Harper, 1990; Godfray & Parker, 1992; Forbes, 1993; Rodríguez- Gironés, 1996; Mock & Parker, 1997; Mock et al., 1998; Perry & Roitberg, 2005). Siblicide and sibling cannibalism are a common phenomenon in several orders of arthropods (Elgar & Crespi, 1992; Fincke & Hadrys, 2001; Gagne et al., 2002; Iida, 2003; Michaud & Grant, 2004; Dobler & Kölliker, 2010; Noriyuki et al., 2011, 2012; Collie et al., 2013), fishes (Hecht & Appelbaum, 1988; van Damme et al., 1989; Smith & Reay, 1991; Bry et al., 1992; Fitzgerald & Whoriskey, 1992; Hecht & Pienaar, 1993; Folkvord, 1997; Baras, 1999), amphibians (Crump, 1992, 1996), birds (Mock & Parker, 1997; Watson et al., 1999; Drummond, 2001; Morandini & Ferrer, 2014), and mammals (Frank et al., 1991; Mock & Parker, 1997; Smale et al., 1999) and is termed facultative when it occurs only occasionally (when food resources are low) and obligate when at least one sibling is killed in virtually all broods/cohorts.

Bird rearing decisions can be described by portfolio theory on the premise that higher returns are generally associated with greater risk, and that portfolio diversification reduces risk (Forbes, 2009). Parents of altricial birds structure their brood into core (first-hatched) and marginal (later-hatched) elements that differ in risk profile. Their initial investment, in the egg and the resources required to sustain the offspring during a period of clemency (Forbes & Ydenberg, 1992), is modest as brood reduction usually occurs early before food demands are high (Forbes et al., 2002a). Thus parents can await unfolding ecological conditions to determine if the option should be exercised. Under conditions of plenty, the option is called, and marginal progeny reared. If food is insufficient, they cut their losses by allowing marginal progeny to perish early via a fatal sibling rivalry (Mock & Forbes, 1995; Mock & Parker, 1997; Forbes et al., 2001). The natural magnitude of hatching asynchrony may be adaptive in part because it moderates sibling competition and chick losses but also because it may give parents greater control over the outcome of sibling competition (Merkling et al., 2014). Death can be caused by physical damage inflicted by a sibling or by aggressive and/or non-aggressive monopolisation of food resources provided by parents (Forbes, 1993; Mock & Parker, 1997; Drummond, 2001; Drake et al., 2008; Hofer & East, 2008; Hudson & Trillmich, 2008; Trillmich & Wolf, 2008; Morandini & Ferrer, 2014). Avian models of facultative siblicide predict, and are supported by empirical evidence that fitness benefits acquired by dominant offspring through despotic behavior should increase as parental provisioning rates decline. In the blue-footed booby (Sula nebouxii), differential mortality of junior chicks was associated with a 20-25% weight deficiency of the senior sib, implying siblicidal brood reduction triggered at a weight threshold. Parents and senior chick cooperate in the siblicide, as if their fitness interests were congruent (Drummond et al., 1986). In contrast, fitness costs should increase and benefits associated with despotic behavior should decline as parental provisioning rates rise (Parker et al., 1989; White et al., 2010). Food amount also influences aggression in obligate brood reducers (Osorno & Drummond, 2003). These findings have been confirmed in a mammalian context of facultative siblicide (Wachter et al., 2002; Hofer & East, 2008; Morandini & Ferrer, 2014) and non-lethal aggressive competition between offspring (Hodge et al., 2009).

Siblicide and cannibalism are a primary cause of mortality throughout nymph development (Dobler & Kölliker, 2010, 2011) in the European earwig (Forficula auricularia), an insect species with uniparental maternal care. Interactions among unrelated nymphs frequently occur under natural conditions due to brood mixing (Kölliker & Vancassel, 2007; Wong & Kölliker, 2013). Cues of maternal condition affect offspring behavior in terms of sibling cannibalism: cues of poor maternal condition enhanced nymph survival in early broods, but reduced nymph survival in late broods, and vice versa for cues of good condition (Wong et al., 2014). Sibling interactions also reflect cooperative behaviors in the form of food sharing in nonderived families of the European earwig. Food ingested by individual offspring was transferred to their siblings through mouth-to-anus contacts and active allo-coprophagy. Relatedness influenced the strategy used by nymphs to provide and/or obtain food from other nestmates, but the two strategies resulted in equivalent outcomes in terms of amount of food transferred (Falk et al., 2014). Together with maternal sensitivities to condition-dependent nymph chemical cues (Mas et al., 2009; Wong & Kölliker, 2012; Meunier & Kölliker, 2012), these findings establish a context-dependent reciprocal information exchange about condition between earwig mothers and their offspring that are potentially mediated by cuticular hydrocarbons and affect the rate of siblicide or offspring cooperation (Wong & Kölliker, 2012; Gómez & Kölliker, 2013).

Even matricidal behavior occurs in extremely resource-limited habitats in which the parental organism serves well to nurture the progeny. Several examples of maternal death (endotokia matricida) that occurs during matricidal hatching are provided by Caleb Finch (1990, p. 102). In a variety of spiders, juveniles eat their mother before dispersing from the communal nest (Wise, 2006). Females of the Japanese foliage spider, Chiracanthium japonicum, are eaten by their offspring at the end of the maternal care period (Toyama, 2003). Facultative endotokia matricida has been observed in a variety of parasitic and free-living oviparous nematodes as a response to food limitation (Ayalew & Murphy, 1986; Johnigk & Ehlers, 1999; Baliadi et al., 2001; Hirao & Ehlers, 2010).

10.2 Unpredictable or limited resources and sociality

Throughout this work (see chapters 4.2, 5.3.1 and 6) cumulative evidence has been presented indicating that more challenging environments with limited resources select for cooperativity. Meta-analyses and models show that many mammals, birds, fishes and insects are found living at densities at the carrying capacity of their habitats (Sibly et al., 2005; Brook & Bradshaw, 2006). Thus, in most animal species a population's growth rate is a decreasing function of density (Sibly et al., 2005) which explains the relative stability of animal populations that do not increase at rates their fertility would allow. In unpredictable or dangerous environments, the chances of successfully raising offspring independently are limited either by predator pressure or harsh environmental conditions. Under such conditions groups are better able to find scattered food, repel enemies, care for young, and thus reduce the risk of brood loss. This scenario has long been a favored explanation for evolution of eusociality in insects (Wilson, 1971; Lin & Michener, 1972), and has been termed ‘life insurer’ eusociality (Queller & Strassmann, 1998) in Hymenoptera, in which helpless larvae require extended care and foraging adults are under constant threat from enemies (Queller, 1989; Gadagkar, 1991b). A similar explanation has been offered for sociality in vertebrates threatened by harsh environments with unpredictable food supply. In both naked mole-rats (Faulkes et al., 1997) and certain cooperatively breeding bird families (Rubenstein & Lovette, 2007), phylogenetically controlled comparative analyses identified significant associations between sociality and unpredictable environmental conditions (Duffy & Macdonald, 2009). Habitat saturation in favorable environments with predictable resources but dense populations and strong pressure from competitors or other enemies may also limit breeding opportunities, a finding that was conceptualized by the long-standing habitat saturation hypothesis (Selander, 1964; Brown, 1974; Emlen, 1982b). Particularly in long-lived species (Hatchwell & Komdeur, 2000), offspring have few opportunities for independent breeding and little option but to live instead as helpers at the nest of parents or other relatives. A global comparative analysis of 182 species of birds supported this hypothesis, confirming that cooperative breeding was significantly associated with sedentary life in warm, invariable climates (Arnold & Owens, 1999). A somewhat similar argument has been made for sociality in ‘fortress defender’ insects, which include certain gall-forming aphids and thrips, and many termites, that live in protected habitats providing both food and shelter (Alexander et al., 1991; Crespi, 1994; Queller & Strassmann, 1998). Based on findings in theoretical and experimental research (MacLean & Gudelj, 2006; Andras et al., 2007; Gore et al., 2009; Requejo & Camacho, 2011, 2012a, b, 2013; Smaldino et al., 2013a, b; Chen & Perc, 2014), cooperation has been found to be the more likely behavior if initially the common resources are limited rather than abundant. In models, harsher environments led to higher long-term frequencies of cooperators (Smaldino et al., 2013a, b), and select for reduced parasitism (Hochberg et al., 2000; Requejo & Camacho, 2011) lending theoretical support to Kropotkin’s (1902) proposal that harsh environments should select for cooperation. Cheater phenotypes of bacteria and amoebae have a high risk to go extinct under conditions of unpredictable and fluctuating resource availability such as in feast and famine cycles (Hilson et al., 1994; Pál & Papp, 2000; Ennis et al., 2003; Fiegna & Velicer, 2003; Rainey & Rainey, 2003; Castillo et al., 2005; Kuzdzal-Fick et al., 2011). These findings are in agreement with data from experiments conducted on yeast (MacLean & Gudelj, 2006; Gore et al., 2009) as well as on social vertebrates (Shen et al., 2012). Similarly, experiments on social vertebrates indicate that unfavorable environmental conditions, where resources are limited, reduce social conflict and make social vertebrates more cooperative (Shen et al., 2012).

10.3 Abundant resources and sociality

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 (MacArthur & Wilson, 1967). These alternative constellations of life-history traits became known as life-history strategies (Pianka 1970; 1974b; Reznick et al., 2002). The life history strategies of cooperators are biased towards the K-selected end of the r-K continuum (Brown, 1974, 1987; Gaston, 1978; Russell, 1989; Rowley & Russell, 1990; Poiani & Jermiin, 1994; Arnold & Owens, 1998, 1999). On the other hand, current abundance of resources, respectively a history of resource abundance or stress, may select for asociality.

Already bacteria and protozoa have options for either social or asocial behavior dependent on the inverse relationship of resource availability and intercellular communication. In chemostats (well-mixed, continuous systems), the selfish high growth rate strategist (at the cost of low growth yield) will always outcompete the cooperative high growth yield strategist (at the cost of low growth rate), because high growth rate strategists grow faster at every substrate concentration above zero (Kreft, 2004b). In the unstructured environment of suspension cultures, resources are a public good and accessible to every individual. In the case of batch cultures, this will select for the fastest growing, selfish organism (Bachmann et al., 2013). The asocial alternative may be pursued in stressed, socially deprived microorganisms (Boesen et al., 1992). In constant, nutrient-rich environments where benefits associated with sporogenesis (a social behavior) are absent and no longer balance the cost of constructing spores, sporulation ability of Bacillus subtilis was lost over 6,000 generations (Maughan et al., 2009). Propagation of B. subtilis for less than 2,000 generations in a nutrient-rich environment where sporulation is suppressed led to rapid onset of genomic erosion including biosynthetic pathways, sporulation, competence, and DNA repair (Brown et al., 2011). The social prokaryote Myxococcus xanthus loses its social behavior when propagated in nutrient-rich habitats in which their social behaviors for starvation-induced spore production or predatory efficiency were not under positive selection (Velicer et al., 1998; Velicer & Stredwick, 2002). This selective regime places no positive selection on M. xanthus social traits but rather on competitiveness under asocial growth conditions. The replicate lineages improved their maximum growth rates an average of 37% over 1,000 generations, but all populations incurred partial or complete losses in their capacity for social motility and social development during this period of adaptation (Velicer et al., 2006). Bacteria or protists which are shear-stressed by constant shaking (Hellung-Larsen & Lyhne, 1993; Fang et al., 1997) are deprived of humoral or cellular contacts with their siblings (Velicer et al., 1998). Bacteria cultured in such an asocial environment but exposed to abundant nutrient resources exhibit substantially attenuated capacities for social behavior following starvation (Velicer et al., 1998). Similarly, disruption of cell-cell contacts that mediate social control (Hanahan & Weinberg, 2000; Bornstein & Sage, 2002; Giancotti & Tarone, 2003) appears to be a basic requirement of carcinogenesis (Yamasaki et al., 1995; Trosko & Ruch, 1998) which can be viewed as social cheating taking advantage of the abundant nutrient resources of the internal milieu of a metazoan organism. The model amoeba Dictyostelium discoideum is single-celled under conditions of nutritional abundance but aggregate under metabolic stress. Following growth with glucose, D. discoideum cells exhibit lowered DIF sensitivity, cheat and are overrepresented in the spores when mixed with cells grown without glucose (Leach et al., 1973; Thompson & Kay, 2000b). When the amount of glucose available in the media is increased, defective yeast that do not pay a cost for producing invertase can spread faster than cooperative yeast, even driving cooperative yeast to extinction (MacLean & Gudelj, 2006; Gore et al., 2009). At either extreme, nutrient-rich or very nutrient-poor conditions, greater numbers of cells are in the planktonic phase where they have greater access to the local nutrients or can be distributed to a new environment (Stanley & Lazazzera, 2004). Biofilm formation is inhibited and biofilm depth reduced by catabolite repression (Jackson et al., 2002; Stanley et al., 2003), possibly as an indicator to the cells that they are in a nutrient-rich environment where there is no growth advantage to being in a biofilm. When the cells are in the planktonic phase, there is greater cell surface area exposed to the nutrients (Stanley & Lazazzera, 2004).

It has been shown repeatedly that the optimal strategies in social games depend on the availability of resources (Burtsev & Turchin, 2006). A key finding is that the investment into helping decreases, and the probability of escalated conflicts increases, for individuals that are likely to inherit a valuable nest (Abbot et al., 2001; Cant et al., 2006a; Field & Cant, 2009), epitomizing a general evolutionary link between resource value and societal conflict (Hoffmann et al., 2012). Individuals following a minimal-effort (plastic) strategy are the ones that most strongly bring sustainability in commons structured in large groups and abundant public goods (Cavaliere & Poyatos, 2013).

Subjected to environmental conditions that do not require cooperation, social deficits become common; all of the experimental lineages developed such deficits (Lyon, 2007). An excessive abundance of common resources deters socially responsible actions on structured populations (Chen & Perc, 2014). If either the common resources are initially too many, if the cooperators are too effective in refilling the pool, or if the maximally allowed endowments are too low for allowing an immediate dissemination of accumulated goods, the defectors are able to take full advantage of their refrain from contributing without suffering the consequences. If sufficiently abundant, the excess allows defectors to free-ride well over the time horizon that is required for cooperators to die out (Chen & Perc, 2014). A similar development may be observed in modern human societies with their abundant resources that appear to promote the loss of social cooperation and solidarity. Putnam (2000) argued that social capital “has eroded steadily and sometimes dramatically over the past two generations” (Putnam, 2000, p. 287). Other evidence supports his claim, including a 2006 study which concluded that not only have informal and formal networks declined, as Putnam notes, but close personal ties have also diminished in recent years (Kibert et al., 2011).

10.4 Resources and sociality in plants

Most plants require a similar balance of resources–energy, water, and mineral nutrients–to maintain optimal growth. Natural environments, however, differ by at least two orders of magnitude in the availability of these resources (Chapin et al., 1987). Light intensity varies 100-fold from the canopy to the floor of a rainforest (Björkman, 1981); annual precipitation ranges 500-fold (10-5000 mm/yr) from deserts to tropical rainforests; and the amount of nitrogen available to plants varies from 0.09 g/m2 x yr in polar desert (Dowding et al., 1981) to 22.8 g/m2 x yr in a rich tropical rainforest (Vitousek, 1984). Among botanists there is a broad consensus that the balance between negative and positive interactions should shift along environmental gradients (Tielbörger & Kadmon, 2000), with competition prevailing under environmentally benign conditions and positive interactions dominating under harsh conditions (Bertness & Callaway, 1994; Bertness & Hacker, 1994; Bertness & Leonard, 1997; Callaway & Walker, 1997; Brooker & Callaghan, 1998; Callaway et al., 2002). The importance of facilitation should increase with increasing abiotic stress, while negative interactions should prevail under benign conditions (Bertness & Callaway, 1994; Callaway & Walker, 1997; Holmgren et al., 1997; Brooker & Callaghan, 1998; Holmgren & Scheffer, 2010). Empirical evidence for this hypothesis stems from a variety of studies investigating interactions between organisms along spatial gradients in the physical environment (Walker & Chapin, 1987; Bertness & Shumway, 1993; Bertness & Hacker, 1994; Bertness & Yeh, 1994; Greenlee & Callaway, 1996; Bertness & Leonard, 1997). Importantly, positive interactions are more prominent under moderately stressful rather than under extreme conditions (Holmgren & Scheffer, 2010).

Clonal growth in plants produces iterated modular units (ramets), each with the same genotype and each potentially capable of independent growth and reproduction. Individuals of clonal plants consist of physically and physiologically connected ramets. In modular, clonal organisms the ability to maximise resource capture as an important characteristic of highly competitive organisms may be enhanced by the physiological integration of modules (ramets) making up the clone. In splitters, they are integrated for a time shorter than ramet generation time (i.e. the time it takes to produce the first offspring ramet), whereas in integrators connections between ramets persist for a longer time (Klimeš, 2008). Clonal integration categories are necessarily heterogeneous, as they include plants with a contrasting architecture of below-ground organs, role of vegetative multiplication and reproduction in their life cycle, storage organs, etc (Jónsdóttir & Watson, 1997). It has been predicted that integrators should prevail in stressful environments, such as habitats poor in nutrients, whereas splitters are expected to dominate in benign habitats, such as fertile areas with a moderate climate (Mágori et al., 2003). Several groups reported a prevalence of clonal integrators in stressful environments (Jónsdóttir & Watson, 1997; Klimeš, 2008; Klimešová et al., 2011). In contrast, in completely homogeneous environments, models show that plants with no clonal integration are usually favored (Oborny et al., 2000). Plants at lower temperatures and higher altitudes, where abiotic stress is high, compete less and cooperate more with their neighbors (Callaway et al., 2002). However, daughter ramets in multiple ramet clones may also compete for resources and inhibit each other (Abrahamson et al., 1991; Hellström et al., 2006; Wang P et al., 2012).

Patchiness or heterogeneity of the habitat arises due to the uneven distribution of environmental features impinging on the organism (e.g. soil quality, topography, temperature, light, food, enemies, etc.). The extent of heterogeneity may vary depending on two factors, patch size and the relative quality differences between patches (Gillespie, 1974). Numerous studies have shown that clonal integration can buffer against environmental heterogeneity (Hartnett & Bazzaz, 1985; Pitelka & Ashmun, 1985; Salzman, 1985; Alpert & Mooney, 1986; Hutchings & Bradbury, 1986; Landa et al., 1992; Stuefer et al., 1994; Wijesinghe & Handel, 1994). Physiological integration, or the sharing of resources between ramets, is considered to be of particular benefit when resources are distributed in discrete patches within the habitat (Cook, 1983; Harper, 1985; Pitelka & Ashmun, 1985; Hutchings & Bradbury, 1986). A number of studies have shown that in clonal plants water, nutrients and photosynthates can be redistributed from ramets growing in favorable patches to those growing in poor patches (Noble & Marshall, 1983; Salzman & Parker, 1985; Slade & Hutchings, 1987; Chapman et al., 1992; Hutchings & de Kroon, 1994) and hence to allow the plants to grow better (Alpert, 1991; Alpert & Stuefer, 1997; Hutchings & Wijesinghe, 1997; Du et al., 2009). This may be true of clonal animals as well (Crowell, 1957; Rees et al., 1970; Jackson, 1977; Best & Thorpe, 1985). This type of physiological integration is thought in some cases to buffer the entire clone against adverse effects resulting from spatial and temporal changes in the quality of a habitat (Hartnett & Bazzaz, 1983; Pitelka & Ashmun, 1985; Hutchings & Bradbury, 1986; Hutchings, 1999). Intriguingly, clones of the herb Fragaria chiloensis from patchier habitats, where high levels of different resources tend to be spatially separated, have a greater capacity for division of labor (Roiloa et al., 2007). Populations of the same clonal species from more heterogeneous habitats can show a higher capacity for resource sharing and cooperative behavior, suggesting that physiological integration has been selected for in response to resource patchiness and that patterns of integration may thus to some extent match patterns of patchiness in the habitat of a species (Alpert, 1999; Alpert et al., 2003; Roiloa et al., 2007; Nilsson & D’Hertefeldt, 2008; He et al., 2011; Wang et al., 2011). When the relative abundance of essential resources varies within a clone each ramet specializes in a division of labor to acquire a locally abundant instead of a scarce resource (Alpert & Stuefer, 1997; Magyar et al., 2007). For example, in habitats consisting of patches with high availability of light but low availability of nutrients, and patches with high nutrients but low light, a ramet growing in high light and low nutrients will allocate a relatively high proportion of biomass, and thus of carbon, to shoots if it is connected to a ramet growing in low light and high nutrients, and that ramet in high nutrients will allocate a high proportion of biomass to roots (e.g. Alpert et al., 2003; Roiloa et al., 2007). Simulations have pointed out a number of habitat types characterized by low patch density, and high spatial and temporal variation of the patches where a splitter (displaying competitive behavior) was not viable at all, and only an integrator (displaying cooperative behavior) could persist. In constant environments, the splitter could survive being confined to clusters of good patches (Kun & Oborny, 2003).

In a uniform habitat, ramets will occupy microsites with similar levels of external resource availability. Clones from populations in relatively uniform habitats such as grassland are selected for low levels of resource sharing between ramets (Alpert, 1999). Based on literature survey and field observations, Pitelka and Ashmun (1985) expected splitting in stressful and resource-poor environments, because maintenance of connections between ramets is costly (Caraco & Kelly, 1991). In such environments, dispersal is favored, both by seeds and vegetative propagules, which are sometimes highly specialized, such as turions, bulbils and detachable buds. The adaptive advantage of clonal splitting in desert shrubs appears to be largely a risk-spreading mechanism that enables independent mortality of integrated hydraulic units or ramets. This should be especially advantageous in heterogeneous, water-limited environments, where soil water occurs in pockets too small to support a large shrub-genet and clonal splitting may cause an increase in intraclonal competition among ramets (Schenk, 1999).

Many clonal plants are colonized by arbuscular mycorrhizal fungi. Arbuscular mycorrhizal fungi are likely to reduce contrasts in effective resource levels and have been shown to partly replace the effects and benefits of clonal integration in low-nutrient and heterogeneous habitats (Du et al., 2009). The interaction between colonization by arbuscular mycorrhizal fungi and physiological integration in a clonal plant provides an intriguing example of how biotic and abiotic factors could interact in the evolution of cooperative and mutualistic behavior (Du et al., 2009).

11. The genetics of social interactions


Summary
A key goal of sociogenomics is to determine how genes and patterns of gene expression change with the shift from solitary to social life. Complex traits are products of a large number of loci with individually small and possibly context-dependent effects. Social behavior is a complex trait. Even in simple organisms such as yeast and social amoeba, the regulation is cooperation is highly polygenic and complex. Aggressive behavior is a complex trait with substantial effects on sociality, regulated by multiple genetic factors as well as by a set of neurotransmitters and neurohormones. Possibly hundreds of candidate genes have been associated with aggressive behavior in fruit flies, zebrafish, mice and humans, many of which, in Drosophila, have pleiotropic effects on metabolism, development, and/or other behavioral traits.
Insulin and insulin-like messengers signal nutritional status to the reproductive axis and thus regulate reproductive functions such as maturation and gametogenesis in metazoans. In eusocial bees, wasps, ants and termites metabolism signaling pathways also regulate the division of labor. Insulin-like peptides mediate a direct and positive long-range effect on ovarian activity. Worker ovaries are inhibited by pheromones of queen and brood. Some workers dramatically increase the activity of their ovarioles in the absence of the queen indicating that signals from the queen, either directly or indirectly, suppress insulin signaling. Honeybee queen pheromones suppressed solitary fruit fly ovaries in much the same way as it suppresses bee worker ovaries. Queen-worker pheromone communication is a multicomponent, labile dialog between the castes that is independent of genetic relatedness and even works across large phylogenetic distances.

The theories of social evolution rely on the assumption that social traits have a simple genetic basis that can be treated as a mechanistic ‘black box’ and largely ignored (Grafen, 1982). The main reason for this is the paucity of classical and molecular genetic tools available to biologists studying classical social organisms (Foster et al., 2007). A key goal of sociogenomics is to determine how genes and patterns of gene expression change with the shift from solitary to social life (Robinson, 1999; Robinson et al., 2005; Owens, 2006; Foster et al., 2007; Smith et al., 2008; O’Connell & Hofmann, 2011; Rittschof & Robinson, 2014). Variation in behaviors in natural populations arises from complex networks of multiple segregating polymorphic alleles, characterized by pleiotropy and widespread epistasis. These networks are sexually dimorphic and sensitive to environmental modulation. Since behaviors reflect dynamic interactions between organisms and their environments, they are central targets for adaptive evolution (Bendesky & Bargmann, 2011; Anholt & Mackay, 2015).

11.1 Social behavior, a complex trait

Tremendous progress has been made resolving certain well-studied “simple” traits at the molecular level (Han & Sternberg, 1990; Hobbs et al., 1992; Carroll et al., 1994; Dilda & Mackay, 2002; Rogers et al., 2003; Gutteling et al., 2007a, b; Kammenga et al., 2007; Debat et al., 2009; Dreyer & Shingleton, 2011; Manceau et al., 2011; Reed et al., 2011; Tang et al., 2011); however, an understanding of the specification of complex traits is generally lacking (Rockman & Kruglyak, 2006). Much of this complexity results from the phenotype being a convolution of the genotype and the specific environment experienced by the organism (Grishkevich & Yanai, 2013). It has been argued that systems genetics approaches are necessary to understand complex traits (Civelek & Lusis, 2014). Genome-wide association studies have identified thousands of genetic loci that contribute to the complex trait ‘disease susceptibility’ in humans (Glazier et al., 2002; Hindorff et al., 2009; Civelek & Lusis, 2014). For example, more than 100 genomic cancer susceptibility regions have been identified (Stadler et al., 2010). An increasing large number of polymorphisms associated with resistance and susceptibility to malaria have been mapped both using population studies and experimental models of malaria (Fortin et al., 2002). However, aside from their typically modest value for predicting future disease occurrence, this information will provide little mechanistic insight until the loci are translated into genes and pathways. Beyond that, it will be important to understand how the alleles interact with each other or with environmental factors (Civelek & Lusis, 2014).

With regard to social behavior these approaches are still in their infancy. In budding yeast, non-sexual cell aggregation has a polygenic molecular architecture (Li et al., 2013). A genome-scale investigation of the genetic opportunities for cheating in social amoeba showed that cheating is multifaceted and complex by revealing cheater mutations in well over 100 genes of diverse types (Santorelli et al., 2008). Multiple mechanisms and pathways may be involved in cheating behavior, including GTPase regulatory activity, polyketide synthesis, nucleotide binding, and phosphoric ester hydrolase activity (Santorelli et al., 2008). If even in this rather simple organism, the actions of so many genes can be exploited for cheating, cooperation can be expected to be complex and highly polygenic (Buttery et al., 2009, 2010). It appeared that response of D. discoideum to social interaction involves many genes with individually small transcriptional effect (Li et al., 2014). These findings corroborate a suggestion already made by Fisher (1918) that most complex traits are products of a large number of loci with individually small and possibly context-dependent effects (Lynch & Lande, 1993). This premise has stood up to a substantial body of empirical work indicating that the number of genes affecting complex traits in populations–or even in simple mouse crosses–range from the tens to hundreds or even thousands (Falconer, 1989; Sjöblom et al., 2006; Wang et al., 2006; Chen Y et al., 2008; Reed et al., 2008; Ayroles et al., 2009; Anholt & Mackay, 2015). In Drosophila, quantitative analyses of new mutations have revealed large numbers of loci affecting quantitative traits (Mackay & Anholt, 2006), as have high-resolution maps of segregating quantitative trait loci (QTL) in Drosophila (Mackay & Anholt, 2006) and mice (Valdar et al., 2006). A systems-genetics analysis of the genetic basis of complex traits in Drosophila implicated several hundred candidate genes that form modules of biologically meaningful correlated transcripts affecting each phenotype (Ayroles et al., 2009). Comparative and integrative studies suggest that social behaviors most likely evolved by acquiring new social roles for ‘‘old’’ genes rather than the evolution of entirely new sets of ‘‘social’’ genes (Robinson & Ben-Shahar, 2002). Evidence indicates that the distribution of allelic effects of quantitative traits is exponential (Mackay, 2001). A few loci with large effects (major genes) influence most of the genetic variation and an increasingly large number of loci with increasingly smaller effects (minor genes) influence the remaining variation. Numerous genes with large effects on behavior have been identified by mutation (Sokolowski, 2001), Mendelian analysis of behavioral variants (Sokolowski, 2001), QTL mapping (Anholt & Mackay, 2004) and the identification of differences in RNA or protein expression between behavioral variants (Insel & Young, 2000). These are all candidate genes for natural variation in behavior (Fitzpatrick et al., 2005). Mutations generally have small effects on social tendencies in animals, and the resulting effects of these mutations on social tendencies have been considered to be too small for selection to act strongly (Sinervo & Lively, 1996; Sinervo & Calsbeek, 2006; Sinervo et al., 2006; Ross-Gillespie et al., 2007; Wild & Traulsen, 2007).

11.1.1 Aggression

Aggressive behavior is a complex trait with substantial effects on sociality, regulated by multiple genetic factors as well as by a set of neurotransmitters and neurohormones (Popova, 2006). Aggressive behavior plays a significant role in the fitness of animals, and it is widespread in the animal kingdom. Animals use aggression to defend themselves and their progeny from attack by predators, to fight for females, to feed, and to maintain the social hierarchy. The study of aggressiveness is complicated, however, by the fact that aggressive behavior is not a unitary trait (Blanchard et al., 2003). An influential classification related to aggression (Moyer, 1968) is based on the eliciting stimuli and included the following types of aggression: predatory (attacks on prey), intermale, fear-induced, irritable, territorial, maternal and instrumental aggression. The problem of searching for the genetic and neurochemical determinants of aggressive behavior is complicated by their apparent heterogeneity (Popova, 2006). Correlating variation in transcript abundance with variation in complex trait phenotypes in 40 wild-derived inbred lines of D. melanogaster, Edwards et al. (2009a) identified 266 novel candidate genes associated with aggressive behavior, many of which have pleiotropic effects on metabolism, development, and/or other behavioral traits. Behavioral tests of mutations in 12 of these candidate genes showed that nine indeed affected aggressive behavior. In a screen of 170 Drosophila P-element insertional mutations for quantitative differences in aggressive behavior, Edwards et al. (2009b) identified 59 mutations in 57 genes that affected aggressive behavior, none of which had been previously implicated to affect aggression. Thirty-two of these mutants exhibited increased aggression, while 27 lines were less aggressive than the control. Most of the mutations had pleiotropic effects on other complex traits (Edwards et al., 2009b). Mapping QTL affecting variation in aggression between two wild-type Drosophila strains, Edwards and Mackay (2009) identified a minimum of five QTL with extensive epistasis in a genomewide scan. In D. melanogaster, a systems genetics approach with mutational analyses together with genome-wide transcript analyses, artificial selection studies, and genome-wide analysis of epistasis revealed that a large segment of the genome contributes to the manifestation of aggressive behavior with widespread epistatic interactions (Anholt & Mackay, 2012). Combining gene expression profiling, behavioral analyses, and pharmacological manipulations, Filby et al. (2010) identified candidate genes and pathways that appear to play significant roles in regulating aggression in the zebrafish (Danio rerio), an animal model in which social rank and aggressiveness tightly correlate. The patterns of differentially-expressed genes between dominant and subordinate males implied multifactorial control of aggression in zebrafish, including the hypothalamo-neurohypophysial-system, serotonin, somatostatin, dopamine, hypothalamo-pituitary-interrenal, hypothalamo-pituitary-gonadal and histamine pathways (Filby et al., 2010). Comparing the genome-wide profiles of chicken brain samples from aggressive and receiver hens in a pecking hierarchy, Buitenhuis et al. (2009) observed that there were 40 differentially expressed genes. In rodents, maternal defense behavior (maternal aggression) involves attacks against intruders by lactating females that are hypothesized to protect the offspring from potential harm (Wolff, 1993; Agrell et al., 1998; Gammie & Lonstein, 2005). Gammie et al. (2007) identified ~200 genes that were differently expressed in the continuous portion of the CNS (including preoptic and hypothalamic regions) between a line of mice selected for high maternal aggression relative to a nonselected control line. At least 39 genes have been associated with some aspects of offense in mice and humans (Maxson, 1999; 2009; Liu et al., 2014). To provide a broader view of gene-environment interaction, a recent study examined the relationship between 403 genetic variants from 39 aggression-related genes (Miczek et al., 2001; Maxson & Canastar, 2003; Maxson, 2009) and youth delinquency and violence (Liu et al., 2014). Low social control was associated with greater genetic risk for delinquency and violence and high/moderate social control with smaller genetic risk for delinquency and violence. A key implication of these findings is that the expression of multiple genes related to delinquency depends on the social environment: gene expression is likely to be amplified in low-social-control environments but tends to be suppressed in high/moderate-social-control environments (Liu et al., 2014). Intriguingly, microsatellite length polymorphisms marking regulatory variation of gene expression appear to be associated with divergent social structure and sociobehavioral traits, including aggression and dispersal propensity development, in various mammals (Trefilov et al., 2000; Hammock & Young, 2004, 2005; Krackow & König, 2008).

11.2 Insulin/IGF-like signaling pathways

Reproduction is an energetically costly process and its linkage to nutrient sensing is evolutionarily conserved (Hietakangas & Cohen, 2009; Jasper & Jones, 2010). Insulin-like and TOR signaling are the metazoan organisms’ monitors of resource utilization. Insulin signaling is a conserved pathway in all metazoans (Barbieri et al., 2003; Piper et al., 2008). Insulin, insulin-like growth factor (IGF) and insulin/IGF-like signaling (IIS) evolved with the appearance of multicellularity, allowing primordial metazoans to respond to a greater diversity of environmental signals. The IIS pathway is split into two complementary and interacting subsystems. The functional separation of IGF and insulin signaling that is seen in mammals dates to approximately 600 million years ago, as the two distinct types of molecules are already present in the lower metazoan tunicate phylum (McRory & Sherwood, 1997). Insects have a single insulin/IGF system that may correspond to the ancestor of the dual insulin/IGF system. The pathway has diverse functions in multicellular organisms, and mutations in IIS can affect growth, development, metabolic homoeostasis, fecundity and stress resistance, as well as lifespan (Broughton & Partridge, 2009). Insulin and insulin-like messengers signal nutritional status to the reproductive axis and thus regulate reproductive functions such as maturation and gametogenesis in nematodes (Heininger, 2012), insects (Riehle & Brown, 1999; Drummond-Barbosa & Spradling, 2001;LaFever & Drummond-Barbosa, 2005; Richard et al., 2005; Tu et al., 2005; Narbonne & Roy, 2006; Hsu & Drummond-Barbosa, 2009), fishes, birds, and mammals (Heininger, 2012). Both at the gonadal and CNS/hypothalamic level, the insulin/IGF system is involved in energy homeostasis, germ cell maturation and reproductive processes in all invertebrate and vertebrate taxa (Brüning et al., 2000; Burks et al., 2000; Lackey et al., 2000; Narbonne & Roy, 2006; Jasper & Jones, 2010). In the honeybee, differential feeding of queen- and worker-destined larvae by nurse bees in the early larval stages triggers a major endocrine response, marked by pronounced differences in the hemolymph juvenile hormone (JH) titer (Rachinsky et al., 1990; Hartfelder & Engels, 1998), as well as in IIS and TOR signaling pathways (Wheeler et al., 2006; Patel et al., 2007; Azevedo & Hartfelder, 2008). JH and ecdysteroid production are dependent on IIS (Riehle & Brown, 1999; Tatar et al., 2001; Tu et al., 2002, 2005; Wu & Brown, 2006), and, in addition, the IIS controls fertility by autonomous effects on the ovaries (Richard et al., 2005). JH shows higher titers during the fourth to fifth instar in queen-destined larvae (Rachinsky & Hartfelder, 1990; Rachinsky et al., 1990). JH affects ovary differentiation from the third larval instar until the onset of metamorphosis: high JH titers in queen larvae prevent autophagic programmed cell death in the ovary (Schmidt Capella & Hartfelder, 2002), thus sustaining tissue survival and differentiation into the large queen ovaries, whereas low JH titers in worker larvae cannot inhibit programmed cell death, which removes 95–99% of the ovariole primordia and leads to the small worker-type ovaries. Functionally, JH application induces queen-like traits in larvae fed a restricted diet (Goewie, 1977; Rembold et al., 1974).

Current evidence indicates that caste development in honeybees involves a complex interaction network composed of the IIS/TOR/epidermal growth factor receptor pathways, JH and ecdysteroids (Wang et al., 2013), which are classic developmental and reproductive hormones in Drosophila (Mirth & Riddiford, 2007) and other insect species (Chapman, 1998). Studies of eusocial bees, wasps, ants and termites have shown that nutritional physiology and some conserved nutrient signalling pathways, especially the insulin and TOR pathways, also regulate the division of labor between queens, foraging and non-foraging individuals (Wheeler et al., 2006; Patel et al., 2007; Ament et al., 2008; Azevedo & Hartfelder, 2008; Okada et al., 2010; Azevedo et al., 2011; Daugherty et al., 2011; Mutti et al., 2011; Wolschin et al., 2011; Hattori et al., 2013; Shao et al., 2014). Developing queens show upregulation of several IIS genes (Wheeler et al., 2006), and knockdown of the bee orthologue of the IIS gene TOR by RNAi blocks queen development (Patel et al., 2007). Wang et al. (2013) performed a functional study on insulin-like peptide genes (AmILP1 and AmILP2) in honeybee larvae by using a double-stranded RNA (dsRNA)-mediated gene knockdown approach and found that JH levels were diminished by AmILP1 dsRNA treatment, while the AmILP2 knockdown caused a reduction in ovary size. During reproductive differentiation in late oogenesis of the queenless ponerine ant, Diacamma sp., IIS pathways are expressed in nurse cells, oocytes, and upper germarial regions of reproductives but not of workers (Okada et al., 2010). In insects, insulin-like peptides are produced by a specialized cell cluster in the brain and mediate a direct and positive long-range effect on ovarian activity (Cao & Brown, 2001; Riehle et al., 2006). Transcriptomics data from the brain of a social wasp demonstrated that the brain of the queen expresses much higher levels of the insulin-like peptide2 than that of the workers (Toth et al., 2007).

Queen-produced pheromones that maintain worker sterility are thought to be taxonomically widespread, as queens, their eggs and queen-derived chemicals have been shown to reduce or eliminate worker reproduction (Vargo, 1992; Peeters et al., 1999; Dietemann et al., 2003; Hoover et al., 2003; Cuvillier-Hot et al., 2004a; Endler et al., 2004; Monnin, 2006; Dengler-Crish & Catania, 2007; Korb et al., 2009; Bhadra et al., 2010; Holman et al., 2010a, 2010b). Programmed cell death is observed in the ovary of workers inhibited by pheromones of queen and brood (Schmidt Capella & Hartfelder, 1998). Some workers dramatically increase the activity of their ovarioles in the absence of the queen indicating that signals from the queen, either directly or indirectly, suppress IIS and its effect on JH activity in the workers (Plettner et al., 1993; Arnold et al., 1994; Winston & Slessor, 1998; Khila & Abouheif, 2010). Intriguingly, sufficient doses of honeybee queen mandibular pheromone suppressed solitary Drosophila melanogaster ovaries in much the same way as it suppresses bee worker ovaries. Exposed fruit flies showed a reduction in ovary size, produced fewer eggs, and generated fewer viable offspring, relative to unexposed controls, indicating that phylogenetically distant solitary and social insects share regulatory pathways associated with female reproduction (Camiletti et al., 2013, 2014). These findings gain plausibility by the recent evidence of putative orthologues of cuticle forming enzymes in Drosophila melanogaster and Apis mellifera (Wang et al., 2014) and of a striking, ~145-million-year-long conservation of nonvolatile, saturated hydrocarbons (HC) as queen pheromone fertility signals (Van Oystaeyen et al., 2014). In D. melanogaster, long-chain HC on the adult fly cuticle are perceived by contact or at a short distance by other flies (Jallon, 1984; Ferveur, 2005), are thought to be related to ovarian activity (Cuvillier-Hot et al., 2001; Liebig et al., 2009; Peeters & Liebig, 2009; Ferveur & Cobb, 2010; Liebig, 2010; Fedina et al., 2012) and IIS signaling (Richard et al., 2005; Fedina et al., 2012) and the large bouquet of HC is affected by mating (Everaerts et al., 2010). There is cumulative evidence that queen-worker pheromone communication is a multicomponent, labile dialog between the castes, rather than a simple, fixed signal-response system (Kocher & Grozinger, 2011) that is independent of genetic relatedness and even works across large phylogenetic distances.

12. The neurobiology of social interactions


Summary
Neurobiological signal systems are pleiotropic; social and asocial effects are context-specific and dynamic. The neurotransmitters serotonin, γ-aminobutyric acid, noradrenaline and dopamine, the steroids corticosteroids, testosterone and estradiol, and the peptides oxytocin and arginine vasopressin have established effects on animal and human social behaviors. There appears to be marked conservation in the molecular mechanisms regulating social behavior across diverse species. However, species and gender differences in receptor distribution and density may contribute to species/gender differences in the effects of neurotransmitters and neurohormones on behavior. Particularly, homologs of oxytocin and vasopressin play a general role in the modulation of social and reproductive behaviors in such diverse organisms as hydra, worms, insects, and vertebrates. Maternal brood care is the nucleus of any complex sociality. There is a continuum of neurobiological processes regulating maternal behavior, pair bonding and social behaviors. Intriguingly, closely related vole species differ in oxytocin and vasopressin signaling in a gender-specific fashion with profound consequences for pair bonding and sociality. Social experiences occurring during postnatal, juvenile, and in some cases, early adulthood impact on several neurobiological systems with long-term, even transgenerational consequences on sociality. Like the vertebrate neurohormonal system, insect endocrine systems are plastic in response to the social environment. Both theory and empirical evidence suggest that the origin of hymenopteran worker behavior lies in parental care redirected towards siblings. From a neurobiological and physiological point of view it appears as if in workers brood care, and division of labor have taken the role of a substitute or surrogate reproductive activity.

The term ‘‘sociality’’ encompasses a wide diversity of behaviors that do not evolve in a linked fashion across species. Thus grouping, monogamy, paternal care, cooperative breeding/alloparental care, and various other forms of social contact are evolutionarily labile and evolve in an almost cafeteria-like fashion, indicating that relevant neural mechanisms are at least partially dissociable (Goodson, 2013). However, there is one constant: not a single more complex social system (I exclude the temporary, targeted, alliances such as swarms or herds as anti-predator defense) evolved without maternal care as the core social relationship within these social groups. Hormones often are the proximate mechanisms by which social traits critical to life-history trade-offs are expressed, and can be helpful in gaining insights into ultimate function (van Anders et al., 2011). The social behaviors are clearly opposed to the ancient self-preservative behaviors. When these balanced social interactions are disturbed, e.g., by a stressor, the self-preservative, fight-flight response pattern takes priority (Henry & Wang, 1998). Evolutionary theories of pair bonds highlight the similarity of these types of bonds and those that exist between parents and offspring; both include attachment, intimacy, and overlapping hormonal mechanisms (Carter, 1998). Given that parent-offspring bonds are likely to be evolutionarily older, pair bonds may be predicated upon a neuroendocrine system that evolved to support parent-offspring bonds, but in general promotes nurturance (Fisher, 1992; Carter, 1998; Fernandez-Duque et al., 2009).

12.1 Vertebrates

The neurotransmitters serotonin, γ-aminobutyric acid, noradrenaline and dopamine, the steroids corticosteroids, testosterone and estradiol, and the peptides oxytocin and arginine vasopressin have established effects on animal and human social behaviors (Miczek et al., 2002; Kosfeld et al., 2005; Donaldson & Young, 2008; Heinrichs et al., 2009; Ebstein et al., 2010; Bos et al., 2012; McCall & Singer, 2012). Social behavior is affected by a variety of factors, e.g. environmental conditions acting through sensory inputs, the hypothalamic-pituitary-gonadal axis, the hypothalamic-pituitary-adrenal stress response axis (Gordon et al., 2011), and genetics (Ebstein et al., 2010). Studies in rodents suggest that the neurohypophysial hormones in concert with steroids are key components in the central mediation of complex social behaviors, including affiliation, parental care, sexual behavior, mate guarding, and territorial aggression (Gimpl & Fahrenholz, 2001; Insel, 2010). Abundant evidence exists for hormonal control of aggression (Koolhaas et al., 1990; Ferris et al., 1997), partner preference (Arletti et al., 1992; Insel & Young, 2001), maternal bonding and affiliative behavior (Kendrick, 2000; Lim & Young, 2006; Neumann, 2008; Ross & Young, 2009; Insel, 2010), trust (Kosfeld et al., 2005; Zak et al., 2005; Baumgarter et al., 2008), and social status (Winslow & Insel, 1991; Sapolsky, 2004). From the huge amount of neurobiological data, only a limited selection will be used to outline some general patterns.

Particularly, aggression is amenable to experimental investigation. Both the serotonergic and dopaminergic systems influence aggressive behaviors in both mammals and birds (Nelson & Chiavegatto, 2001; Van Hierden et al., 2002; Kjaer et al., 2004). Elevated CNS serotonin levels lead to decreased aggression in many different species (Asberg et al., 1987; Linnoila & Virkkunen, 1992; Chiavegatto & Nelson, 2003), while high dopamine levels are associated with aggression and dominance behavior (van Erp & Miczek, 2000; Ryding et al., 2008). Both human and animal studies have also identified the GABAergic system (de Almeida et al., 2004, 2005; Gourley et al., 2005), the noradrenergic system (Miczek & Fish, 2005) and the glutamatergic system, particularly N-methyl-D-aspartate (NMDA) receptors (Belozertseva & Bespalov, 1999; Duncan et al., 2004, 2009), to be related to cognitive-behavioral performance and social interaction, including prosocial, antisocial, and aggressive behaviors (Comai et al., 2012). Observations in several species indicate a role for serotonin in the modulation of prosocial behavior (Wood et al., 2006). Investigation of both peripheral and central indices of serotonin function have shown negative associations with violent and aggressive behavior in rhesus monkeys (Higley et al., 1992, 1996) and human clinical samples (Brown et al., 1979; Virkkunen et al., 1994), as well as positive associations with socially affiliative behaviors (such as grooming and approach) (Raleigh et al., 1981). Pharmacological interventions that increase serotonergic activity also attest to a prominent role in important aspects of social behavior and aggression (Fuller, 1996). The selective serotonin reuptake inhibitor fluoxetine, increases affiliative behaviors in rodents (Knutson & Panksepp, 1996), while each of fluoxetine, quipazine (a 5-HT1 and 5-HT2 receptor agonist), and the amino-acid serotonin precursor, L-tryptophan, have been found to increase affiliative behaviors and decrease nonsocial behaviors (such as vigilance and avoidance) in primates (Raleigh et al., 1985). In humans, serotonin promotes social cooperation (Wood et al., 2006).

A number of studies indicate that the serotonin and dopamine systems interact closely at a basic neurophysiological level (Wong et al., 1995; Kapur & Remington, 1996; Daw et al., 2002; Seo et al., 2008). These general actions, however, are modulated by receptor subtypes and their brain regional variation (Cologer-Clifford et al., 1997; Nelson & Chiavegatto, 2001; Olivier, 2004; de Almeida et al., 2005; de Boer & Koolhaas, 2005; Nelson & Trainor, 2007; Comai et al., 2012). The interaction between low serotonin and high testosterone levels in the central nervous system has a significant effect on the neural mechanisms involved in the expression of aggressive behavior (Birger et al., 2003). Testosterone is associated with social aggression in a wide range of species, affecting such behaviors as mate guarding, territorial and dominance aggression (Mazur & Booth, 1998; Nelson & Trainor, 2007). By contrast, testosterone appears to be less involved in other non-social forms of aggression, such as predatory and anti-predatory aggression (Wingfield et al., 2000). Exposure to embryonic testosterone in birds, in particular, increases aggression and dominance in adults and hence alters fitness, even though embryonic exposure does not affect the levels of testosterone circulating in adults (Partecke & Schwabl, 2008). Testosterone disrupts collaboration by increasing egocentric choices (Wright et al., 2012). The role of testosterone in human social behavior might be best understood in terms of the search for, and maintenance of, social status (Eisenegger et al., 2012). In terms of the stress response, testosterone is traditionally associated with inhibiting the hypothalamic-pituitary-adrenal axis (Handa et al., 1994).

12.1.1 Oxytocin/vasopressin

Within vertebrates, a majority of work relating neuropeptides to social behavior has focused on members of the oxytocin/vasopressin family. Homologs of oxytocin and vasopressin existed at least 700 million years ago and have been identified in such diverse organisms as hydra, worms, insects, and vertebrates (Minakata, 2010). Among these distant taxa, oxytocin- and vasopressin-related peptides play a general role in the modulation of social and reproductive behaviors. In contrast to this apparent conservation in function, the specific behaviors affected by these neuropeptides are notably species-specific (Donaldson & Young, 2008). The mammalian oxytocin (OT) and arginine vasopressin (AVP) nonapeptides, so called for their nine–amino acid composition, differ from each other at only two amino acid positions. OT, AVP, and their respective nonmammalian vertebrate lineages are thought to have arisen from a gene duplication event before vertebrate divergence. Within these lineages, peptides vary by a single amino acid, and their genes are found near each other on the same chromosome. Invertebrates, with few exceptions, have only one OT/AVP homolog, whereas vertebrates have two (Acher et al., 1995; Caldwell & Young, 2006). OT is a very abundant neuropeptide. This became obvious in a study where the most prevalent rat hypothalamic-specific mRNAs were analyzed. OT was found to be the most abundant of 43 transcripts identified (Gautvik et al., 1996). Brain OT receptor (OTR)-mediated actions were shown to be significantly involved in the regulation of a variety of behaviors. A broad variety of behaviors, including maternal care and aggression, pair bonding, sexual behavior, and social memory and support, as well as anxiety-related behavior and stress coping, are modulated by brain OT (Kenrick, 2000; Numan & Insel, 2003; Pedersen, 2004; Lonstein & Morrell, 2007; Campbell, 2008; Neumann, 2008, 2009; Insel, 2010; Anacker & Beery, 2013). While the importance of OT for social functions appears nearly universal, central OTR distribution varies between species and may relate to species-typical social behavior (Insel et al., 1994; Young, 1999; Insel & Young, 2000; Donaldson & Young, 2008). Likewise, there are substantial gender-related differences of both OT and AVP brain innervation (de Vries, 1999). Gender-specific aspects of mammalian social behavior are regulated by the steroid hormones, notably estrogens and progestogens, which determine the expression and distribution of receptors for neuropeptides such as OT and AVP (Dantzer 1998; Kalamatianos et al. 2004).

OT and OT-like hormones facilitate reproduction in all vertebrates at several levels (Gimpl & Fahrenholz, 2001). Peripheral and intracerebroventricular administration of OT reliably leads to maternal behavior in virgin mammals who would ordinarily ignore or attack pups (Pedersen & Prange, 1979; Pedersen et al., 1982; Kendrick et al., 1987; Kendrick, 2000; Francis et al., 2002). Conversely, experimental manipulations that decrease OT levels or block OTR activation within the brain reduce maternal behaviors (Pedersen et al., 1985; Pedersen & Boccia, 2003; Caldwell & Young, 2006). OT has been shown to play a key role in processes of parent-infant bonding across a range of mammalian species, including rats, prairie voles, sheep, and primates (Kendrick et al., 1987; Holman & Goy, 1995; Kendrick, 2000; Neumann, 2008; Maestripieri et al., 2009). Pregnancy, lactation, maternal behavior and exogenous treatments with estrogen and progesterone increase OT immunoreactivity, OT mRNA expression, OTR mRNA expression and OTR binding in brain areas central for parenting and the reward parents derive from their infants (Ross & Young, 2009). AVP is involved in erection and ejaculation in species including humans, rats, and rabbits (Segarra et al., 1998; Gupta et al., 2008), and it mediates a variety of male-typical social behaviors including aggression, territoriality, and pair bonding in various species. The brain OT system is also involved to a significant extent in female sexual behavior, at least in rats, specifically promoting lordosis behavior in an estrogen-dependent manner. Moreover, in late pregnancy and during lactation, central OT is involved in the establishment and fine-tuned maintenance of maternal care. In addition to maternal care, most lactating mammals show a remarkable level of aggression (with a link to anxiety; Bosch et al., 2005), thus protecting their offspring against potential social threats.

Pair bonding is exclusive to the 3 to 5% of mammalian species that are socially monogamous. Traditional laboratory organisms such as rats and mice, however, do not display mate-based pair bonds. The mammalian genus Microtus provides an excellent model for investigating the evolution of social behavior. Rodents of the genus Microtus show dramatic species differences in social structure: prairie voles (Microtus ochrogaster) exhibit a monogamous social structure in nature with either communal nesting or cooperative breeding (Getz & McGuire, 1997; Roberts et al., 1998), whereas closely related meadow voles (Microtus pennsylvanicus) and montane voles (Microtus montanus) are solitary and polygamous (Gruder-Adams & Getz, 1985; Shapiro & Dewsbury, 1990). In monogamous prairie voles, both OT and AVP were demonstrated to play a major role in pair bonding in a gender-specific fashion. Although both peptides may facilitate pair-bond formation in either sex (Winslow et al., 1993; Insel & Hulihan, 1995; Cho et al., 1999; Young & Wang, 2004; Young et al., 2010), AVP seems to be more important in males, whereas OT is more critical in females. In male prairie voles, partner loss elicits anxiety-like and depression-like behaviors, disrupts bond-related behaviors, and alters neuropeptide systems that regulate such behaviors (Sun et al., 2014). Central infusions of OT facilitate (Williams et al., 1994) whereas a selective OTR antagonist inhibits pair-bond formation in female prairie voles (Insel & Hulihan, 1995). Moreover, there are notable differences in OTR distribution patterns among prairie voles and montane voles mainly in the brain region nucleus accumbens (Insel & Shapiro, 1992), and an OTR antagonist applied directly to this region blocks mating-induced partner preference formation (Young et al., 2001). Overexpression of the OTR in nucleus accumbens accelerates partner preference formation in female prairie voles (Ross et al., 2009). Differences in potential regulatory elements in the OTR gene, which could reflect variation in gene expression, have been found between prairie and montane voles (Young et al., 1996). A recent study has found that manipulations of OT activity alter partner-directed social behavior during pair interactions in the pair-bonding primate Black-tufted Marmoset (Callithrix penicillata) (Smith AS et al., 2009). In humans, single nucleotide polymorphisms in OTR are also associated with traits reflecting pair-bonding in women (Walum et al., 2012), empathy and stress reactivity (Rodrigues et al., 2009).
Repetitive microsatellites mutate at relatively high rates and may contribute to the rapid evolution of species-typical traits. Microsatellites in the cis-regulatory regions of genes may significantly enhance the rate of evolution of gene expression patterns and selectable phenotypic traits (King, 1994; Kashi et al., 1997; Li et al., 2004; Kashi & King, 2006). Vasopressin and the vasopressin 1a receptor (V1aR) have been implicated in social recognition and interaction processes in a variety of species (Ferguson et al., 2002; Bielsky & Young, 2004; Bielsky et al., 2005; Egashira et al., 2007; Donaldson et al., 2010). The V1aR is expressed at higher levels in the ventral forebrain of monogamous than in promiscuous vole species (Insel et al., 1994). Intracerebroventricular injection of a V1a receptor antagonist into male prairie voles abolished their mate preference demonstrating the role of vasopressin and its V1a receptor in this species-specific behavior (Winslow et al., 1993). Individual alleles of a repetitive polymorphic microsatellite in the 5’ region of the prairie vole V1aR gene modify V1aR gene expression in vitro. In vivo, this regulatory polymorphism predicts both individual differences in receptor distribution patterns and socio-behavioral traits (Hammock & Young, 2004, 2005). Male prairie voles with the longest DNA microsatellite strings spend more time with their mates and pups than male prairie voles with shorter strings (Hammock & Young, 2005). Transgenic mice that have received the prairie vole V1aR gene affiliate significantly more with their mated partners (Young et al., 1999). By using viral vector V1aR gene transfer into the ventral forebrain, Lim et al. (2004) substantially increased partner preference formation in the socially promiscuous meadow vole, showing that a change in the expression of a single gene in the larger context of pre-existing genetic and neural circuits can profoundly alter social behavior. In conclusion, a change of affiliative behavior, possibly resulting from the evolutionary history of the species (Getz & McGuire, 1997; Gadagkar, 2004b), had significant consequences for both the mating and social systems of Microtus ochrogaster compared to Microtus pennsylvanicus and possibly in humans. Both the findings of Sun et al. (2014) and the Young group argue for the joint malleability of the neuropetide system through environmental and genetic factors.

Another example of behavioral actions of intracerebral OT is the promotion of social memory processes and recognition of conspecifics, as revealed in rats, mice, sheep and voles. OT knockout mice show deficits in social recognition, but normal nonsocial learning and memory abilities (Ferguson et al., 2000; Kavaliers et al., 2003; Takayanagi et al., 2005). These deficits are reversible by OT administration specifically into the central amygdala (Ferguson et al., 2001). OT released in response to social stimuli may be part of a neuroendocrine substrate which underlies the benefits of positive social experiences. In rats, OT exerts potent and long-term physiological antistress effects (Uvnäs-Moberg, 1998) and promotes social affiliation during threatening situations (Bowen & McGregor, 2014). In the highly gregarious zebra finch (Estrildidae: Taeniopygia guttata), blockade of nonapeptide receptors by an OT antagonist significantly reduced time spent with large groups and familiar social partners, independent of time spent in social contact. Opposing effects were produced by central infusions of mesotocin (avian OT homologue) (Goodson et al., 2009). Experimental evidence suggests that, in humans, brain OT exerts similar behavioral effects (Kenrick, 2000; Numan & Insel, 2003; Pedersen, 2004; Lonstein & Morrell, 2007; Campbell, 2008; Neumann, 2008). There is also evidence from neuroimaging studies suggesting effects of OT treatment on amygdala activity (Kirsch et al., 2005; Domes et al., 2007a; Petrovic et al., 2008), a brain region known to be of importance for regulation of social behaviors. Exogenously administered OT increases rates of several cooperative behaviors in genetically related meerkats (Madden & Clutton-Brock, 2011). In wild chimpanzees, OT levels were higher after grooming with bond partners compared with nonbond partners regardless of genetic relatedness or sexual interest (Crockford et al., 2013). In genetically unrelated humans, intranasal OT increases trust (Kosfeld et al., 2005; Baumgartner et al., 2009; Mikolajczak et al., 2010), generosity (Zak et al., 2007), willingness to cooperate (Declerck et al., 2010) and promotes reciprocity (Kosfeld et al., 2005; Baumgartner et al., 2008). However, OT decreases generosity and the adherence to fairness norms in social settings where others are likely to be perceived as not belonging to one’s ingroup (Radke & de Bruijn, 2012).

The commonality of pathways and genes suggests a high degree of conservation in the multi-factorial control of aggression between fish, birds and mammals (Miczek et al., 2001; Caldwell et al., 2008; Buitenhuis et al., 2009; Filby et al., 2010). On the other hand, studies on gene expression in the brain have shown that profound differences in phenotype are often associated with only small changes in gene expression (Nisenbaum, 2002). OT is involved in both acute aggressive interactions and lasting dominance relationships (Anacker & Beery, 2013). The peptide is released in the lateral septum during an acute social defeat in rats (Ebner et al., 2000), which can have lasting effects because initial interactions can be remembered for long periods of time, and thus contribute to a stable social dominance hierarchy (Adkins-Regan, 2005). In established hierarchies, something different is observed. Dominant female rhesus macaques have higher levels of serum OT than subordinates (Michopoulos et al., 2011, 2012), and dominant male squirrel monkeys exhibit greater levels of aggression when given OT infusions (Winslow & Insel, 1991). Conversely, subordinate male rats decrease OTR expression in the long-term establishment of dominance roles (Timmer et al., 2011). Dominant male cichlid fish have higher levels of isotocin, the teleost homolog of oxytocin, in the hindbrain (Almeida et al., 2012), and aggressive dominant three-spined sticklebacks have higher levels of isotocin in the brain as well (Kleszczynska et al., 2012). Also the teleost arginine vasotocin (AVT), and its mammalian homologue AVP, have well established roles in aggression and social position across vertebrate taxa (Caldwell et al., 2008). There have been conflicting findings of effects of AVT on aggression in fish (Lema & Nevitt, 2004; Santangelo & Bass, 2006; Filby et al., 2010). This may be because AVT increases aggression in non-territorial (colonial) species and decreases it in territorial species (Caldwell et al., 2008). Even within a single species, AVT can have opposite effects between behavioral phenotypes; AVT increases aggression in non-territorial males but decreases aggression in territorial males (Semsar et al., 2001; Filby et al., 2010).

12.1.2 Hypothalamic-pituitary-adrenal axis

Mounting evidence suggests that corticotropin-releasing factor (CRF), CRF receptors, and the hypothalamic-pituitary-adrenal (HPA) axis may play an important role in social behaviors (Hostetler & Ryabinin, 2013). Plasma corticosteroids rise fast and early in the course of an agonistic encounter between rats (Schuurman, 1980), even before actual aggressive behavior is observed (Haller et al., 1995). There are several reasons to believe that the surge in plasma glucocorticoids caused by the confrontation with a potential adversary plays an important role in the subsequent aggressive conflict. Inhibiting corticosterone synthesis inhibits aggressive behavior (Haller et al., 1996; Mikics et al., 2004), whereas corticosterone treatment administered to metyrapone-treated rats rapidly reinstates aggressive behavior (Mikics et al., 2004). Also, injecting glucocorticosteroids (GC) into the hypothalamus of the golden hamster rapidly facilitates aggression (Hayden-Hixson & Ferris, 1991). A fast, mutual, positive feedback of the controlling mechanisms within the time frame of a single conflict may contribute to the precipitation and escalation of violent behavior under stressful conditions (Kruk et al., 2004). Stress is a major factor promoting aggression and violence in humans (Barnett et al., 1991; Tardiff, 1992). Of all of the enduring effects of GCs, the most profound might be their capacity to alter their own regulation (Martin et al., 2011). In mammals (Weaver et al., 2004), birds (Love et al., 2008), and amphibians (Denver, 2009), stressors early in life alter HPA and OT/AVP pathway regulation throughout life with long-term, even transgenerational, effects on social behavior (Champagne, 2008; Weinstock, 2008; Curley et al., 2011; Bales & Perkeybile, 2012; Veenema, 2012).

The CRF system has also been implicated in prosocial and affiliative behaviors such as parental care, maternal defense, sexual behavior, and pair bonding (Hostetler & Ryabinin, 2013). Low GC levels are also associated with antisocial behavior (Shirtcliff et al., 2009). Acute stress and activation of the HPA axis may increase prosocial behavior (Taylor et al., 2000; DeVries et al., 2002; de Waal & Suchak, 2010; Buchanan et al., 2012; Von Dawans et al., 2012; Buchanan & Preston, 2014). There is neuroanatomical and neurobiological evidence for reciprocal regulation of the HPA and OT systems balancing stress and affect (Dabrowska et al., 2011). In the rat, physical and mental stresses or central administration of CRF evoke an increase in OT secretion while AVP secretion remains unchanged (Bruhn et al., 1986). OT appears to function as an anxiolytic, suppressing HPA axis function during periods of stress (Smith & Wang, 2012). OT was found to inhibit HPA axis responses to a wide variety of physical, emotional and pharmacological stressors, and may thus make an important contribution to the attenuated stress responsiveness found in pregnancy and during lactation. These neuroendocrine adaptations are mainly reflected by lower peak levels of corticosterone⁄cortisol and corticotropin in response to acute stressor exposure. In humans, OT administration enhances the stress-alleviating effects of social support (Heinrichs et al., 2003).

Social experiences occurring during postnatal, juvenile, and in some cases, early adulthood impact on social competence later in life (mammals including humans: Bastian et al., 2003; Levy et al., 2003; Margulis et al., 2005; Branchi et al., 2006, 2009; Bester-Meredith & Marler, 2007; Flinn et al., 2011; birds: Adkins-Regan & Krakauer, 2000; Bertin et al., 2009; White et al., 2010; fish: Moretz et al., 2007; Arnold & Taborsky, 2010; Taborsky et al., 2012). This early socialization affects several neurobiological systems, including monoamine (e.g. serotonin and dopamine), GABAergic, glutamatergic, vasopressin, oxytocin, estrogen sensitivity and estrogen receptors, and receptors for CRF and GC, changes that may also extend transgenerationally (Curley et al., 2011). Intriguingly, the hormonal system orchestrating social interactions entertains manifold cross-talk to the metabolic system, particularly insulin (Bobbioni-Harsch et al., 1995; Björkstrand et al., 1996). In the coral reef cleaner fish Labroides dimidiatus, pro- and anti-social behaviors are physiological condition-dependent involving metabolic stress response signaling pathways mediated by e.g. cortisol and vasotocin (Bshary, 2002; Bshary et al., 2011; Soares et al., 2012, 2014; Cardoso et al., 2015). These pathways establish the pleiotropic links between resource availability and social behavior (see also chapter 10).

12.2 Invertebrates

Like the vertebrate neurohormonal system, insect endocrine systems also respond to social stimuli (Scott, 2006a, b; Tibbetts & Huang, 2010; Tibbetts & Crocker, 2014), hence are plastic in response to the social environment. In eusocial communities, suppression of aggression is an essential aspect of communal life. Queen pheromone modulates brain dopamine (DA) function in young worker honeybees (Beggs et al. 2007; Beggs & Mercer, 2009). A key component of queen pheromone, homovanillyl alcohol, has been found to lower young worker brain DA levels (Beggs et al. 2007; Beggs & Mercer, 2009). Reducing DA levels, changing receptor expression and activation has been suggested to reduce aggression, stinging, and aversive learning in young bees (the bees closest to the queen, feeding her); this would translate into an increase in both the queen’s and the colony’s fitness (Vergoz et al., 2007; Wright, 2009). In adult honeybees, aggression is associated with juvenile hormone (JH) titers (Pearce et al., 2001). Older bees, which have higher JH levels (Huang et al., 1994), are generally more aggressive than younger bees with lower JH levels (Breed, 1983). Bees treated with a JH analog exhibited an earlier response to alarm pheromone (Robinson, 1987), and the proportion of bees acting as guards also increased (Sasagawa et al., 1989). Queen pheromones suppress worker JH production (Pankiw et al., 1998). Workers reared in isolation showed higher levels of aggressiveness toward other bees (Breed, 1983), and isolated bees have elevated JH levels compared to bees reared in groups or in a colony (Huang & Robinson, 1992). Guards at the entrance that are less exposed to queen pheromones exhibit low thresholds for the expression of aggressiveness (Breed et al., 1992; Pearce et al., 2001) and JH titers of guards are higher than of all other middle-age bees except undertakers (Huang et al., 1994; Pearce et al., 2001).

The hormonal and neurochemical mechanisms modulating aggression in arthropods seem to be quite varied. In many insects, JH has a role in aggression, dominance and reproduction (Röseler, 1991; Larrere & Couillaud, 1993; Bloch et al., 2000a; Cnaani et al., 2000; Scott, 2006a; Kou et al., 2009; Tibbetts & Izzo, 2009; Tibbetts & Huang, 2010; Tibbets et al., 2011), but not in crickets (Adamo et al., 1994). The low DA and JH levels in reproductive queenless ants are very different from what has been observed in fire ant queens, which show a rise in DA with the onset of reproduction (Boulay et al., 2001) and have high JH levels when mature (Brent & Vargo, 2003). Pharmacological reduction of JH levels in Bombus terrestris workers decreased ovarian activation, irrespective of the bees’ dominance rank within the group, an effect that was remedied by JH replacement therapy. Low JH also decreased aggressiveness and increased ester-sterility signal production; these changes were rank-dependent, and affected mainly the most reproductive and the least aggressive workers, respectively, and could not be remedied by JH replacement therapy (Amsalem et al., 2014). Aggressiveness was depressed in crickets depleted of octopamine (a biogenic monoamine structurally related to noradrenaline) and dopamine, and was restored by an octopamine agonist (Stevenson et al., 2005). Octopamine has also been reported to correlate with dominance and aggressiveness in bumblebees (Bloch et al., 2000b); in fruit flies, octopamine increases aggressiveness and dopamine is negatively correlated (Baier et al., 2002) while serotonin seems to promote aggression in crustaceans (Kravitz, 2000; Kravitz & Huber, 2003).

Queenless ant species have colonies composed exclusively of monomorphic workers, each one capable of mating and laying eggs, forming societies with a more plastic mode of organization (Peeters, 1993). In these colonies, the division of labor depends on interindividual behavioral interactions that determine not only gathering tasks but also more importantly, reproductive roles. Since queenless ant species are restricted by ecological constraints from forming new, independent colonies, their access to reproduction is entirely dependent on their rank in the dominance hierarchy (Cuvillier-Hot & Lenoir, 2006). The potential reproducers in a queenless ant colony form a dominance hierarchy through confrontations. Because castes are determined through behavioral interactions, the social position of individual workers changes frequently and rapidly. Streblognathus peetersi is a monogynous queenless ant with colonies organised in three distinct groups: the alpha ant is the most dominant worker and the only one to mate and lay eggs; workers of high social ranks are behaviorally dominant (while submissive to the alpha), are rather young and do not mate or lay eggs; workers of low social ranks are subordinate to the other two groups, can be of any age and also remain infertile. Fertile alpha workers are clearly characterised by higher levels of octopamine; this monoamine, thus, seems to correlate with reproductive activity in Streblognathus.

The present results suggest a shift through which DA no longer regulates reproduction but has been co-opted for polyethism regulation in eusocial species. This major physiological transition of the role of DA is similar to the well-documented change in JH activity in honeybees. In insects, JH serves as a gonadotropin to regulate the biosynthesis of vitellogenin and/or its uptake by the developing oocytes (Nijhout, 1994; Gilbert et al., 2000). JH is broadly correlated with fertility, and treatment with JH increases fertility (e.g. Tibbetts & Izzo, 2009; Tibbetts et al., 2011). JH has also a clear gonadotropic role in primitively eusocial species [Bombus (Bloch et al., 2000a), Ropalidia (Agrahari & Gadagkar, 2003) and Polistes (Tibbetts et al., 2011)]. Queens produce a pheromone that inhibits the activity of workers’ corpora allata, resulting in low JH titers and thus in restricted ovarian development and inhibition of egg laying (van Honk et al., 1980; Röseler et al., 1981). Once the queen loses control over her workers, or when the queen dies, the corpora allata of a few dominant workers are reactivated and this allows them to lay eggs and to inhibit the corpora allata of other workers (van Honk et al., 1980; Röseler et al., 1981; van Honk & Hogeweg, 1981). In at least some insects, high JH seems to inhibit parental care (Rankin et al., 1997). However, in many taxa, JH may have no effect or a positive effect on parental care (Mas & Kolliker, 2008; Trumbo, 2002). JH seems to have lost its reproductive function in honeybees, and instead, regulates colony polyethism (Hartfelder, 2000; Robinson & Vargo, 1997). As opposed to the pattern of JH levels in Bombus, JH levels in both honeybees and queenless ants are low in reproductive castes (and in honeybee nurses) and high in foragers in both cases (Robinson et al., 1991; Sommer et al., 1993). This double inversion of JH and DA roles suggests that DA in queenless ants plays an allatotropic role, as it does in several species [e.g., Locusta migratoria (Lafon-Cazal & Baehr, 1988), Apis mellifera (Kaatz et al., 1994; Rachinsky, 1994)]. Both theory and empirical evidence suggest that the origin of hymenopteran worker behavior lies in parental care redirected towards siblings (Wheeler, 1928; Kennedy, 1966; Michener, 1969; Wilson, 1971; Hamilton, 1972; Alexander, 1974; West-Eberhard, 1987; Alexander et al., 1991; Bourke & Franks, 1995; Linksvayer & Wade, 2005; Amdam et al., 2006; Boomsma, 2007; Toth et al., 2006; Boomsma et al., 2011). Taken together it appears as if in workers, from a neurobiological and physiological point of view, brood care, and division of labor have taken the role of a substitute or surrogate reproductive activity.

13. The ecology of kin recognition


Summary
Social recognition abilities have evolved for a host of functional reasons. Familiarity as the sole mechanism to determine kin and used as a proxy of relatedness has been shown for diverse organisms and contexts. Cooperation based on genetic cues, on the other hand, would be a conundrum because it requires polymorphic recognition loci, yet cooperation is predicted to erode this genetic variation. In many species, olfactory information provides salient information about species identity, kin, sex, social status, and/or reproductive condition. Nestmate recognition based on cuticular hydrocarbons is common among the eusocial insects. Nestmate recognition cues have a substantial hereditary component but are also modulated by environmental factors that may override genetically based ones. A highly polymorphic major histocompatibility complex (MHC) encodes proteins that target foreign molecules for immune cell recognition. Animals are capable of converting this genetic code into olfactory information that is used for social recognition. The odorants are the same MHC peptides used during immune recognition, which provides the molecular logic linking selection acting on MHC-mediated behaviors with selection acting on immune recognition.
Accurate discrimination of conspecifics with regard to genetic relatedness is a crucial prerequisite for nepotistic behaviors (preferential treatment of kin). However, kin discrimination and nepotistic behavior are only rarely associated. Thus, there is no evidence that evolution of kin recognition/discrimination and prosocial behavior are interdependent. This prompts the question whether kin discrimination may serve another purpose. A totalitarian system, such as a unitary metazoan organism or eusocial colony, requires effective means of information and communication and compliant agents that enforce the despotic system. Throughout phylogenesis, allorecognition, the self/nonself discrimination, served the purpose to ensure the reproductive monopoly of the germline and to prevent germline parasitism. The importance of selfish/cheater behavior in the activation of the immune system is demonstrated by the ability of the immune system to tolerate allogeneic cooperative microbiota and fight syngeneic selfish cells (tumors) by responding to ‘‘selfish/cheater signals’’. Thus, similar to metazoan organisms that have evolved systems to control cancer, the evolution of insect societies requires mechanisms to control selfish worker reproduction.There is a nearly perfect analogy of the eusocial nestmate recognition and worker policing systems to the function of the immune system in multicellular organisms.

Social recognition abilities have evolved for a host of functional reasons (Colgan, 1983; Sherman et al., 1997; Pfennig et al., 1999; Mateo, 2004; Tibbetts & Dale, 2007; Gabor et al., 2012) such as inbreeding avoidance (Sherborne et al., 2007), mate choice (Brown, 1997; Tregenza & Wedell, 2000; Candolin, 2003), mother-offspring recognition (Beecher, 1991; Booth & Katz, 2000), dominance hierarchies (Barnard & Burk, 1979; d’Ettorre & Heinze, 2005; Gherardi & Atema, 2005), and territory establishment and defense (Temeles, 1994; Tibbetts & Dale, 2007; Grether, 2011). Visual, olfactory and auditory cues are prominent in many recognitive systems. Kin recognition, the differential treatment of conspecifics varying in genetic relatedness, has been documented in multiple animal and plant taxa (Fletcher & Michener, 1987; Waldman, 1988; Hepper, 1991; Pfennig & Sherman, 1995; Penn & Frommen, 2010). Familiarity as the sole mechanism to determine kin and used as a proxy of relatedness has been shown for diverse organisms and contexts (Schausberger, 2007): Association attractiveness among salmon varies with familiarity and odor concentration (Courtenay et al., 2001), guppies choose shoal mates based on familiarity (Griffiths & Magurran, 1999), familiar piglets show less aggressiveness than unfamiliar ones irrespective of genetic relatedness (Stookey & Gonyou, 1998), juvenile sticklebacks compete less with familiar individuals without any indication that genetic relatedness plays a role (Utne-Palm & Hart, 2000), and penguin chicks discriminate between familiar and unfamiliar calls but not between calls of familiar kin and nonkin (Nakagawa et al., 2001). On the other hand, there is no evidence for the use of genetically determined recognition cues or templates (Komdeur et al., 2008). Kin discrimination as a means to perceive certainty of paternity cannot be exerted by birds (Kempenaers & Sheldon, 1996; DeWoody et al. 2001; Sheldon, 2002; Komdeur et al., 2004) and fish (Svensson et al., 1998). A simple spatially based recognition rule, such as ‘feed anything in my nest or territory’, is widespread and successfully exploited by bird species that are brood parasites or in which extrapair paternity occurs (Komdeur & Hatchwell, 1999; Aktipis & Fernandez-Duque, 2011). In other bird species (e.g. dunnocks, Prunella modularis and alpine accentors, Prunella collaris) males use indirect cues and behavioral ‘rules of thumb’, adjusting their parental care based on the amount of attempted extrapair copulations or exclusive sexual access they had to the female, which correlates with degree of paternity (Burke et al., 1989; Davies et al., 1992; Hartley et al., 1995; Ewen & Armstrong, 2000). There is a continuum of social behavior in spiders from small subsocial mother-offspring-sibling groups to complex, cooperative societies of thousands of individuals (Buskirk, 1981; Avilés, 1997; Lubin & Bilde, 2007). Most social spiders, despite multiple evolutionary origins, share a suite of traits that includes the acceptance of alien spiders (unrelated and unfamiliar conspecifics) into the group without overt aggression (Lubin & Bilde, 2007). These social spider species do not appear to differentiate between conspecific aliens and members of their own colony (Pasquet et al., 1997; but see Yip et al., 2009), between silk from kin or nonkin (Bilde et al., 2002; Buser, 2002), or even between heterospecific from conspecific spiders in the same genus (Seibt & Wickler, 1988).

Ever since Hamilton (1964) proposed his concept of kin selection to explain social evolution and the evolution of ‘‘altruistic’’ behaviors, there has been a strong interest in kin recognition and the mechanisms by which animals may be able to distinguish between their kin (genetically-related individuals) and nonkin (genetically-unrelated individuals) (e.g., Holmes & Sherman, 1983; Sherman & Holmes, 1985; Waldman, 1988; Waldman et al., 1988; Hepper, 1991; Sherman et al., 1997; Neff & Gross, 2001; Tang-Martinez, 2001; Krupp et al., 2011). However, Hamilton later modified his thinking to suggest that an innate ability to recognize actual genetic relatedness was unlikely to be the dominant mediating mechanism for kin altruism: “But once again, we do not expect anything describable as an innate kin recognition adaptation, used for social behaviour other than mating…” (Hamilton 1987, p. 425).

The prevailing categorization lists four different mechanisms of ‘kin recognition’: (i) spatially based recognition; (ii) association or familiarity; (iii) phenotype matching (Hauber & Sherman, 2001); and (iv) recognition alleles, sensu Hamilton (1964). The classical mechanisms of kin recognition have been criticized (Fletcher, 1987; Waldman, 1987; Waldman et al., 1988; Barnard, 1990). As Grafen (1990) pointed out, the conclusions that can be drawn when an animal is observed to behave differently on first encounter towards differently related groups of conspecifics are limited. We can conclude that individuals are able to detect some feature of conspecifics or their environment which correlates with their genotype, but it does not necessarily follow that this ability has evolved to distinguish kin from nonkin or has anything to do with kin selection (Grafen, 1990). Thus, although there are undoubtedly many examples of kin bias (e.g. cooperative breeding in certain bird species), only those involving recognition of relatedness per se (i.e. genetic similarity), with the purpose of biasing responses towards relatives (kin discrimination), deserve the label ‘kin recognition’ (Grafen, 1990, 1991; Barnard, 1991; Barnard et al., 1991). Barnard (1999) argued that the four categories of recognition mechanisms are not helpful for two reasons: first they do not constitute a set of mutually exclusive alternatives (Fletcher, 1987; Waldman, 1987), and second they confuse questions about the development and expression of traits indicating genetic similarity with those about the processes of perceiving and acting upon this information (Fletcher, 1987; Waldman, 1987; Barnard, 1990).

Cooperation based on genetic cues, if it exists, would be a conundrum because it requires polymorphic recognition loci, yet cooperation is predicted to erode this genetic variation (e.g. Crozier, 1986; Grosberg & Quinn, 1989; Tsutsui, 2004; Gardner & West, 2007; Rousset & Roze, 2007; Boomsma & d’Ettorre, 2013; Holman et al., 2013). This problem, sometimes termed Crozier’s paradox, applies whenever individuals with common recognition cues receive greater average fitness returns from social interactions. Crozier pointed out that those individuals bearing a common marker more readily enter into social interactions, and hence enjoy a higher reproductive success than those individuals bearing rare markers. Disproportionate fitness benefits for individuals with common recognition alleles should produce positive frequency-dependent selection at recognition loci, depleting the genetic variance necessary for kin recognition (Holman et al., 2013). Thus, already common markers become more common still, and eventually all individuals carry the same marker. At this point, it fails to be diagnostic of kinship, and there has been a breakdown of kin recognition. Alexander and Borgia (1978) pointed out that “fixation for ‘genetic recognition’ alleles would lead to rather uniform distribution of benefits to all interactants, and relatives would be preferred only while such traits were on their way to fixation. Such systems could account for variations in nepotistic behavior associated with social structure only if there is (a) rapid recurrence of mutations leading to genetic recognition or (b) common association with disadvantageous characters through either pleiotropy or linkage. The first condition is unlikely because of the necessity of complex recognition mechanisms, and there is no reason to expect the second.” On the other hand, theoretical models suggest that many recognition loci likely have some primary function such as histocompatibility or inbreeding avoidance that is subject to diversifying selection, keeping them variable (Holman et al., 2013). Overall, genetic kin recognition is inherently unstable, explaining its rarity (Gardner & West, 2007).

13.1 Kin recognition by olfactory cues

In many species, olfactory information provides salient information about species identity, sex, social status, and/or reproductive condition (Brown, 1979; Halpin, 1986; Johnston, 2003; Wyatt, 2003; Kavaliers et al., 2005). Members of vertebrate species commonly use olfactory signatures to distinguish between individual members and to denote various subgroups, including the recognition of kin versus nonkin (Porter, 1998; Yamazaki et al., 1999, 2000; Mateo, 2002; Reynolds & Sheldon, 2003). Kin recognition is triggered by olfactory cues in a variety of vertebrates (e.g. Porter et al., 1978; Holmes 1984; Quinn & Busack, 1985; Waldman, 1985; Todrank et al., 1998; Gerlach et al., 2008; Mehlis et al., 2008; Bonadonna & Sanz-Aguilar, 2012; Krause et al., 2012). Several of these species do not form cooperative alliances with their kin. Olfactory cues appear to be important in kin recognition mechanisms as they function to facilitate inbreeding avoidance and the attainment of optimal outbreeding (e.g., in mice, Gilder & Slater, 1978; D'Udine & Partridge, 1981).

13.1.1 Hydrocarbons as olfactory cues in nestmate recognition

Nestmate recognition is common among the eusocial insects (Wilson, 1971; Singer & Espelie, 1992; Clément & Bagnères, 1998; Vander Meer & Morel, 1998; Strassmann et al., 2000; Tarpy et al., 2004). Eusocial insects commonly rely on olfactory cues for making social discriminations. The olfactory pathway of Hymenoptera is well investigated (Kirschner et al., 2006; Zube et al., 2008; Kleineidam & Rössler; 2009; Galizia & Rössler, 2010; Nakanishi et al., 2010). For example, all annotated ant, bee and wasp species have several-fold more odorant receptors than most solitary insects, suggesting that enhanced olfactory abilities may contribute to complex social organization (LeBoeuf et al., 2013).

The most studied, and probably the most widespread, intracolony chemical messengers are cuticular hydrocarbons (Richard & Hunt, 2013). Insects cover their cuticles with a complex mixture of organic chemicals that is thought to have evolved to protect against desiccation and/or to prevent microbial infection (Gibbs, 2002; Howard & Blomquist, 2005). These chemicals are predominantly hydrocarbons, though substances such as wax esters, aldehydes, ketones, sterols and alcohols are also often found (Lockey, 1988; Copren et al., 2005; Monnin, 2006; van Wilgenburg et al., 2011). Many of these cuticular chemicals have been evolutionarily co-opted to function in communication, and, as signals for species-, gender-, and mate-recognition, dominance and fertility cues, have pivotal roles in the social lives of their bearers (Howard & Blomquist, 2005; Blomquist & Bagnères, 2010; Babis et al., 2014). In eusocial insects these cuticular hydrocarbons have been co-opted to function in recognition and are perceived by other individuals by direct antennal contact or at a short distance (Cuvillier-Hot et al., 2005; Brandstaetter et al., 2008). The pattern of cuticular hydrocarbons can be complex and dynamic, with many compounds varying both qualitatively (different species typically have different compounds) and quantitatively (different colonies of the same species have different relative proportions of the same hydrocarbons) (Guerrieri et al., 2009). In ants (Beye et al., 1998; Lahav et al., 1999; Wagner et al., 2000; Roulston et al., 2003; Martin et al., 2008), bees (Paxton et al., 1999), wasps (Dani et al., 2001; Ruther et al., 2002) and termites (Polizzi & Forschler, 1999; Weil et al., 2009) the determination of self versus nonself in the context of nestmate recognition is frequently based on the expression of cuticular hydrocarbon profiles (Liang et al., 2001; Suarez et al., 2002; Blomquist & Bagnères, 2010). Nestmate recognition cues have a substantial hereditary component (van Zweden et al., 2009, 2010) but are also modulated by environmental factors. Three major sources of nestmate recognition cues are identified as: (i) queen dominant (Carlin & Hölldobler, 1983, 1986); (ii) worker-produced (Stuart, 1988; Soroker et al., 1994; Dahbi & Lenoir, 1998; Lahav et al., 1998; van Zweden & d’Ettorre, 2010), and (iii) environmentally derived (Obin, 1986; Crosland, 1989; Heinze et al., 1996). These chemicals may then be distributed throughout the membership of the colony via allogrooming and trophollaxis resulting in a colony-specific odor blend or “gestalt” (Crozier & Dix, 1979; Crozier, 1987; Dahbi & Lenoir, 1998).

The presence of the queen’s pheromones that serve as recognition cues can influence nestmate recognition and intraspecific interactions in several ant species (Carlin & Hölldobler, 1987; Provost, 1987; Keller & Passera, 1989; Lahav et al., 1998; Vander Meer & Alonso, 2002). In orphaned Solenopsis invicta colonies, worker aggression toward non-nestmate conspecifics drops (but still exists with interspecific interactions) and re-establishes following acceptance of a newly mated queen (Vander Meer & Alonso, 2002). The new queen exposes workers to her queen primer pheromone, the level of which correlates with workers’ sensitivity to colony-level differences in cuticular hydrocarbons. Increased sensitivity is linked to higher levels of octopamine in workers’ brains of queenright colonies compared with orphan colonies (Vander Meer et al., 2008). However, the presence of the queen does not influence nestmate recognition cues in several other ant species (Boulay et al., 2004; Caldera & Holway, 2004; Richard et al., 2004; van Zweden et al., 2009).

Sources for environmental cues include diet (Gamboa et al., 1986; Obin & Vander Meer, 1988; Crosland, 1989; Le Moli et al., 1992; Liang & Silverman, 2000; Silverman & Liang, 2001; Florane et al., 2004; Richard et al., 2004; Buczkowski et al., 2005; Sorvari et al., 2008) nesting substrate (Jutsum et al., 1979; Crosland, 1989; Heinze et al., 1996; Richard et al., 2007) and abiotic conditions (Wagner et al., 2001). Environmental cues may override genetically based ones (Gamboa et al., 1986; Bennett, 1989; Beye et al., 1997, 1998; Downs & Ratnieks, 1999). Due to their strong environmental modulation, cuticular hydrocarbons do not appear to be an optimal choice as a tool for genetic kin recognition. Thus, within-colony kin discrimination does not appear to be an important factor organizing honeybee societies (Breed et al., 1994; Keller, 1997; Kryger & Moritz, 1997; Underwood et al., 2004; Rangel et al., 2009). Likewise, studies of ants (Carlin et al., 1993; Snyder, 1993; DeHeer & Ross, 1997; Holzer et al., 2006; vanWilgenburg et al., 2007; Zinck et al., 2009; Kellner & Heinze, 2011; Friend & Bourke, 2012), wasps (Queller et al., 1990; Strassmann et al., 1997, 2000b; Solís et al., 1998; Goodisman et al., 2007), termites (Clément & Bagnères, 1998; Kitade et al., 2004; DeHeer & Vargo, 2006; Atkinson et al., 2008), and aphids (Aoki et al., 1991; Foster & Benton, 1992; Miller, 1998; 2004; Shibao, 1999) failed to demonstrate within-colony kin discrimination. In ant colonies, “workers are able to discriminate between nestmates and intruders, but they also tend to treat all nestmates as colony members, regardless of the degree of relatedness” (Hölldobler & Wilson, 1990, p. 197). Virtually all recognition studies involve between-colony discrimination (nestmate recognition) (Torres et al., 2007; Brandt et al., 2009; Guerrieri et al., 2009; van Zweden & d’Ettorre, 2010) instead of within-colony discrimination (kin recognition) (Vander Meer & Morel, 1998). However, comparison of two allodapine bees, the highly eusocial allodapine Exoneurella tridentata and the facultatively social Exoneura robusta, using a standardised circle-tube apparatus found that (i) discrimination between nestmates and non-nestmates was much more strongly expressed in the facultatively social species and (ii) principal components analyses did not indicate suites of behaviors that permit clear interpretations as being agonistic, cooperative, or avoidance (Dew et al., 2014). The authors concluded that nestmate recognition is not an essential ability for social species (Dew et al., 2014).

Social parasites are able to break the hydrocarbon code and integrate into the normally closed colony of a social insect (Lenoir et al., 2001; Nash & Boomsma, 2008; Bagnères & Lorenzi, 2010; van Zweden & d’Ettorre, 2010). Social parasite species can use either chemical camouflage (recognition cues are acquired from its host) or chemical mimicry (recognition cues are synthesized by the parasite), or a combination of both (Lenoir et al., 2001). However, the result is the same: they overcome detection as non-nestmates. The staphylinid termitophile beetles Trichopsenius frostiT. depressusXenistusa hexagonalis, and Philotermes howardi are guests in termite nests and were found to have hydrocarbon profiles similar to those of their respective hosts (Howard et al., 1980, 1982). Similarly, the paper wasp Polistes atrimandibularis is an obligate social parasite of another paper wasp, P. biglumis bimaculatus. It was shown that, at the point of the colony life cycle that the parasite needs to integrate into its host colony, the hydrocarbon profile of the parasite changes from one that is characterized by unsaturated hydrocarbons into one that matches the host’s profile: characterized by the same saturated methyl-branched hydrocarbons, without any unsaturated hydrocarbons (Bagnères et al., 1996). Likewise, another paper wasp, P. sulcifer, was found to adopt a colony-specific hydrocarbon profile of its host, P. dominulus (Sledge et al., 2001). The ant Polyergus rufescens can achieve social integration in nests of multiple hosts of the genus Formica (subgenus Serviformica) by adopting a hydrocarbon profile that is similar to that of their host, even if the experimental host is not a natural host, like Formica selysi (d’Ettorre et al., 2002).

Drifting, that is leaving the nest to enter a foreign one, is a surprisingly widespread phenomenon among flying social insects (Free 1958; Jay, 1966; Akre et al., 1976; Kasuya, 1981; Paar et al., 2002; Seppa et al., 2002; Birmingham et al., 2004; Lopez-Vaamonde et al., 2004; Blacher et al., 2013) that may be made possible either by a lack of nestmate recognition or social tolerance of colonies. In Apis mellifera drifting can be extensive, with drifters sometimes comprising more than half of the workforce (Free 1958; Jay, 1966; Pfeiffer & Crailsheim, 1998; Neumann et al., 2000, 2001; Chapman et al., 2010). In Bombus terrestris, depending on colony orientation, up to 50% of the workforce can be composed of drifters (Lefebvre & Pierre, 2007; Blacher et al., 2013; O’Connor et al., 2013). In a natural population of the eusocial paper wasp Polistes canadensis 56% of females drifted (Sumner et al., 2007). For Bombus occidentalis and Bombus impatiens, an average of 28 % of workers was discovered to be drifters (Birmingham & Winston, 2004).

13.1.2 Histocompatibility polymorphism

Allorecognition–the ability of an individual organism to distinguish its own tissues from those of another–has been described in all uni- and multicellular phyla (Burnet, 1971), including bacterial self/non-self recognition systems (Gibbs et al., 2008; Budding et al., 2009), kin discrimination in social amoebae (Ostrowski et al., 2008; Benabentos et al., 2009), fungal mating types and heterokaryon incompatibility (Buss, 1982; Glass et al., 2000; Fraser & Heitman, 2003, 2004; Idnurm et al., 2008), and plant self-incompatibility systems (Takayama & Isogai, 2005). For example, self/nonself discrimination and social recognition has been reported in bacteria Myxococcus xanthus (Fiegna & Velicer, 2005; Vos & Velicer, 2009), Proteus mirabilis (Pfaller et al., 2000; Gibbs et al., 2008) and Pseudomonas aeruginosa (Munson et al., 2002). In the ascidian Botryllus schlosseri, histocompatibility assays between randomly collected colonies suggest the presence of a highly polymorphic locus with hundreds of different allotypes (Karakashian & Milkman, 1967; Mukai & Watanabe, 1975; Scofield et al., 1982a, b; Grosberg & Quinn, 1986; Rinkevich et al., 1992; Yund & Feldgarden, 1992; Rosengarten & Nicotra, 2011). Sibling planktonic larvae of the sessile colonial settle in aggregations that are much stronger than expected from dispersal distance effects alone. Laboratory experiments indicate that larvae distinguish kin on the basis of shared alleles at the histocompatibility locus known to regulate fusion between adult colonies. This kin recognition mechanism, along with limited dispersal of larvae, promotes co-settlement of histocompatible individuals (Grosberg & Quinn, 1986). Likewise, in the cnidarian Hydractinia symbiolongicarpus allorecognition specificity is determined by highly polymorphic cell-surface molecules that control fusion or rejection when colonies grow into contact (Rosengarten & Nicotra, 2011). Allorecognition systems are thought to have evolved to limit fusion to self or close kin, thereby reducing the potential cost of germline parasitism (Buss, 1982, 1987).

In vertebrates, histocompatibility is a function of the immune system controlled by a highly polymorphic major histocompatibility complex (MHC), which encodes proteins that target foreign molecules for immune cell recognition. The association of the MHC and immune function suggests an evolutionary relationship between metazoan histocompatibility and the origins of vertebrate immunity. The hypothesis that there is a connection between the immune system and social behaviors such as mate choice mediated via olfactory cues is now four decades old (Thomas, 1975). Specifically, it was proposed that the MHC is involved in cell-cell recognition and the adaptive immune response is in addition responsible for the production of characteristic chemical signals in bodily secretions that enable individual recognition in social contexts such as mate or kin recognition by olfaction (Yamazaki et al., 1976; Beauchamp et al., 1985; Brown & Eklund, 1994; Eggert et al., 1998; Penn & Potts, 1998; Penn, 2002; Leinders-Zufall et al., 2004; Boehm, 2006; Ruff et al., 2012; Overath et al., 2014). The MHC, designated H2 in mice and HLA in humans, is a genomic region on chromosomes 17 and 6, respectively, which harbors the highly polymorphic MHC class I and class II genes as well as many non-polymorphic genes (Murphy, 2012). The MHC is a family of ~50 genes that are known in mammals for their interindividual variability. Although MHC social signaling is thought to occur in over 20 species of vertebrates (Ruff et al., 2012), studies into its molecular mechanism have been predominantly performed with mice and more recently in fish, reptiles, birds and man (Wedekind & Penn, 2000; Milinski, 2006). At three of the more variable human MHC loci, HLA-A, HLA-B and HLA-DRB1, 243, 499 and 321 alleles have been resolved worldwide, respectively, and nucleotide diversity in the human MHC is up to two orders of magnitude higher than the genomic average (Gaudieri et al, 2000; Garrigan & Hedrick, 2003). It seems that animals are capable of converting this genetic code into olfactory information (Hurst et al., 2001; Leinders-Zufall et al., 2004) that is able to activate olfactory sensory neurons at exceedingly low concentrations (Leinders-Zufall et al., 2004). MHC differences are correlated with social recognition in many vertebrate species. The prevalence of this phenomenon (e.g. Olsén et al., 1998; Villinger & Waldman, 2008; Zelano & Edwards, 2002) suggests that the use of MHC-correlated odors in social recognition may be a characteristic in this clade. MHC genes can affect the concentration of volatile acids that produce odor in sweat or urine (Yamazaki et al, 1979; Singh et al, 1987; Wedekind et al, 1995; Wedekind & Furi, 1997; Hurst et al, 2001; Beauchamp & Yamazaki, 2003; Santos et al, 2004; Piertney & Oliver, 2006). Individual variation in volatiles and peptides, which becomes larger with a decrease in the degree of relatedness, can explain attraction or aversion in behavioral contexts such as mating, kin and parent-progeny recognition, as well as inbreeding avoidance without invoking a specific influence of the MHC class I and II genes (Overath et al., 2014). It has been pointed out (Boehm, 2013) that this variability may provide “a universal mechanism of olfactory assessment of genetic individuality even for animals that do not possess an MHC”. The finding that specialized sensory neurons can bind peptides in an MHC-like fashion (Boehm & Zufall, 2006; Spehr et al., 2006) has revealed the long-sought odorants used to recognize the MHC genotype and phenotype of other individuals. The odorants are the same MHC peptides used during immune recognition, which provides the molecular logic linking selection acting on MHC-mediated behaviors with selection acting on immune recognition (Slev et al., 2006).

13.2 Kin discrimination and nepotism

Accurate discrimination of conspecifics with regard to genetic relatedness is a crucial prerequisite for nepotistic behaviors (preferential treatment of kin) (Hamilton, 1987). There is only very rare unambiguous evidence for both kin discrimination and nepotism. Nepotism among closely related female kin (‘close’ kin), including cooperative defense of territories has been well documented in Belding’s ground squirrels, Spermophilus beldingi (Sherman, 1977, 1980, 1981). Females with close kin (e.g. mothers, daughters, sisters) are more likely to give potentially risky alarm calls than females with no close kin. Although S. beldingi adults do not treat distant female kin (e.g. grandmothers, aunts or cousins) and male kin nepotistically, they are able to recognize these individuals as relatives (Mateo, 2002). These findings indicate a dissociation in the evolution of recognition components, such that all S. beldingi kin classes produce discriminable cues, but only close female kin are recipients of nepotism. Males, like females, produce kin labels and can recognize their kin but are not treated nepotistically, nor do they act nepotistically (Mateo, 2002). Golden-mantled ground squirrels (S. lateralis), an asocial species which are closely related to S. beldingi but are not nepotistic, are also able to discriminate among classes of their kin (Mateo, 2002).

In many species, the overall rates at which adult individuals direct aggression towards kin and nonkin are simply indistinguishable. This is the case for meerkats, Suricata suricatta (Madden et al., 2012), European badgers, Meles meles (Hewitt et al., 2009), ringtailed coatis, Nasua nasua (Hirsch et al., 2012), spotted hyaenas (Smith et al., 2010; Wahaj et al., 2004), bonnet macaques, Macaca radiata (Silk et al., 1981b) and rhesus macaques (Widdig et al., 2002). Kinship also fails to curtail aggression in Belding’s ground squirrels (Holmes, 1986; Mateo, 2003) despite strong evidence indicating that kin selection favors alarm calling in this species (Sherman, 1977, 1980, 1981). Kinship also generally fails to protect individuals from becoming victims of coalitionary attacks in mammals with strict linear dominance hierarchies for which social ranks are often established and reinforced within maternal lineages (e.g. baboons: Alberts, 1999; rhesus macaques: Widdig et al., 2001; spotted hyaenas: Smith et al., 2010; Wahaj et al., 2004). Although inter-pride relatedness can be high, response to intruders was influenced by both environmental and social factors, whereas kinship had no detectable effect on territorial conflicts among groups of lions (Spong & Creel, 2004). On the other hand, the low level of aggression within coalitions of male lions is not affected by the degree of genetic relatedness (Packer & Pusey, 1982). Overall, rates of aggression were reduced among kin for only 8 out of the 31 (26%) species reviewed (Smith, 2014) such that species were significantly less likely to be socially tolerant of genetic relatives than expected by chance (binomial test: N = 31 species, P < 0.0001). That is, most species either directed higher rates of aggression towards kin or failed to preferentially tolerate groupmates on the basis of kinship. Moreover, this lack of nepotistic tolerance was statistically indistinguishable between primates and nonprimates (N = 31 species, P = 0.926).

Many social insect colonies are composed of multiple matrilines (progeny from different mothers and, typically, fathers) and/or patrilines (progeny from different fathers). Inclusive fitness theory (Hamilton, 1964), therefore, predicts that nepotism within full-sister groups would be beneficial if this does not incur excessive costs. In eusocial species, however, unambiguous direct evidence for nepotism has so far only been found in drywood termites (Korb, 2006), whereas evidence for the absence of nepotism has been found in numerous species (Queller et al., 1990; Woyciechowski, 1990; Carlin et al., 1993; Breed et al., 1994; Balas & Adams, 1996; Bernasconi & Keller, 1996; DeHeer & Ross, 1997; Keller, 1997; Strassmann et al., 1997; Solís et al., 1998; Mohammedi & Le Conte, 2000; Tarpy et al., 2004; Châline et al., 2005; Ratnieks et al., 2006; Goodisman et al., 2007; Koyama et al., 2007, 2009; Atkinson et al., 2008; Rangel et al., 2009; Zinck et al., 2009; Kellner & Heinze, 2011). In contrast to the lack of evidence for a role of kin recognition in nestmate nepotism, there is abundant evidence for the role of kin recognition in non-nestmate aggression and intracolonial “cheater” detection (e.g. Wilson, 1971; Henderson et al., 1990; Adams, 1991; Vander Meer et al., 1998; Liebig et al., 1999, 2000; Monnin & Peeters, 1999; Dani et al., 2001; Ruther et al., 2002; Roulston et al., 2003; Endler et al., 2004; Dietemann et al., 2005; Cuvillier-Hot et al., 2004a, b; Buchwald & Breed, 2005; Ozaki et al., 2005; Cini et al., 2009; Smith SM et al., 2009; Sturgis & Gordon, 2012). This pattern is exactly what can be expected when kin discrimination is only used to eliminate non-self as can be found in the interactions of metazoan immune-competent cells with intruders (e.g. parasites) and social cheaters (e.g. cancer cells).

In conclusion, there is no evidence that evolution of kin recognition/discrimination and prosocial behavior are interdependent. All conceivable combinations of the two factors occur: (i) kin recognition with and without cooperation; and (ii) cooperation, even eusociality, with and without kin recognition. Thus, kin recognition is neither a necessary nor sufficient condition for prosocial behavior.

13.3 Excursion: An Orwellian Big Brother society needs an effective system of communication and surveillance

In an Orwellian society individuals are oppressed with the mechanics of fear and/or the sedatives of pleasure (Orwell’s “1984” and Huxley’s “A Brave New World” are good examples of this). A totalitarian system requires effective means of information and communication and compliant agents that enforce the despotic system. Most importantly, it should be made sure that the information is understood correctly by each member of the society, so that transgressions of the “rules” can be prevented in the first place. In metazoan organisms, a common “language” as effective communication system is ensured by common genetic descent. At least as important as discrimination of “self” and “nonself” is for the fight against external hazards, it is for the maintenance of internal homeostasis. In multicellular organisms, the immune system performs a large list of ‘‘nonimmunological’’ tasks such as (i) management of the cooperation of syngeneic cells, which may explain the numerous functions of the immune system in the development and maintenance of organisms in the absence of infection; (ii) detection of selfish/cheater behavior of syngeneic or allogeneic cells; and (iii) elimination and memorization of cheaters (Muraille, 2013). The importance of selfish/cheater behavior in the activation of the immune system is demonstrated by the ability of the immune system to tolerate allogeneic cooperative microbiota and fight syngeneic selfish cells (tumors) by responding to ‘‘selfish/cheater signals’’ (Matzinger, 2002; Cremer & Sixt, 2009; Vance et al., 2009; Muraille, 2013). Each unitary metazoan organism (that falls under Weismann’s doctrine) is a totalitarian society in which the soma and each of its cells ultimately serve the final mission of the germline cells, reproduction. It is obvious that autonomy of cell growth is potentially very dangerous to the survival of the organism if not closely regulated (Hull, 1982; Chigira et al., 1990). Therefore, autonomous proliferation of each somatic cell that does not serve this ultimate goal is strictly controlled. Normal tissues carefully control the production and release of growth-promoting signals that instruct entry into and progression through the cell growth-and-division cycle, thereby ensuring a homeostasis of cell number and thus maintenance of normal tissue architecture and function (Hanahan & Weinberg, 2000, 2011). Moreover, cell adhesion molecules and the extracellular matrix milieu keep the cells under tight social control (Hanahan & Weinberg, 2000; Bornstein & Sage, 2002; Giancotti & Tarone, 2003). This social control is even exerted in bacterial communities (Hori & Matsumoto, 2010). Circumvention of this proliferation control is the basic cause of carcinogenesis (Hull, 1982; Chigira et al., 1990; Hanahan & Weinberg, 2000, 2011). Cancer can be considered a social parasite and cheat (Nunney, 1999; Crespi & Summers, 2000; Michor et al., 2004; Dobata & Tsuji, 2009; Bourke, 2011b; Ghoul et al., 2014).

A defining characteristic of eusocial colonies (as of metazoan organisms) is their closed membership. Like a multicellular organism, membership of the colony is “self” and nonmembers, even if they are of the same species, are treated as “nonself” (Krasnec & Breed, 2012). Separation of “self” from “nonself” allows eusocial colonies to prevent invasion by parasites and predators (Vander Meer et al., 1998). Kin discrimination and nepotistic behavior are only rarely associated. This prompts the question whether kin discrimination may serve another purpose. Beehives, wasp and ant nests are police states (Foster & Ratnieks, 2001; Monnin & Ratnieks, 2001; Cuvillier-Hot et al., 2004b) in which selfish and despotic queens, by means of their pheromones, suppress the reproductive activity of their daughters and enforce their reproductive monopoly by murder, torture and imprisonment (Whitfield, 2002). Just as a metazoan organism is vulnerable to parasitism by renegade cells that form tumors (Weinberg, 1998), an insect society is susceptible to exploitation by rogue reproductive workers that can harness the brood-rearing capacity of a colony to enhance their personal reproductive success. It is thought that an insect colony is more likely than a metazoan organism to experience such problems because although the cells of an individual are considered clonal (but see e.g. Muotri et al., 2005; Flores et al., 2007; Lam & Jeffreys, 2007; Bruder et al., 2008; Liang et al., 2008; Piotrowski et al., 2008; Coufal et al., 2009; Quinlan & Hall, 2012), an insect colony comprises a genetically heterogeneous family in which the interests of individuals are not identical, especially in reproductive matters (Beekman & Ratnieks, 2003). Thus, similar to metazoan organisms that have evolved systems to control cancer, the evolution of insect societies requires mechanisms to control selfish worker reproduction (Frank, 1995; Amdam & Seehuus, 2006; Ratnieks et al., 2006; Cremer et al., 2007; Cremer & Sixt, 2009). There is a nearly perfect analogy of the eusocial nestmate recognition and worker policing systems to the function of the immune system in multicellular organisms.

14. Stochasticity and selection: duality in evolution


Summary
Here I refer to the abstract of my recent work (Heininger, 2015) elaborating a cybernetic theory of evolution that was already outlined in an earlier work (Heininger, 2013). Stochasticity and natural selection are the input, respectively, output levels of the evolutionary Black Box. Stochastic environments fundamentally change the rules how evolution works (compared to Darwinian theory that assumed stable environments). Ashby’s ‘Law of Requisite Variety’ describes the behavior of cybernetic systems in variable environments: create internal variety that matches the external variety to fight variation with variation. Particularly, individual-level and population-level selective pressures act jointly dependent on ecological variables. Stochastic environments coerce organisms into lotteries. Risk-avoidance or -spreading are individual- or population-level biological insurances in response to the vagaries of an uncertain, unpredictable environment. The evolution of cooperation is a bet-hedging (risk spreading) strategy of risk-averse individuals.

Darwin’s theory is based on the assumption of a stable environment. Stable environments favor selfish individuals that try to maximize their fitness (the survival of the fittest). Hamilton’s rule allegedly explains the paradox of altruism, that selfish individuals forego their fitness for the benefit of others. Kin selection/inclusive fitness was his explanation for this conundrum. 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 whose Black Box can be understood as learning automaton with separate input and output channels (Heininger, 2013, 2015). 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 input signal is given by a complex biotic and abiotic environment. Natural selection is the output/outcome of the learning automaton.

Abiotic and biotic environments are not stable, but variable, often unpredictable. Particularly, density- and frequency-dependent coevolutionary interactions generate chaotic and stochastic dynamics. Environmental stochasticity changes fundamentally the rules for the “gamble of life”. Stochastic environments coerce organisms into risky lotteries. Chance favors the prepared. Ashby’s (1956) ‘Law of Requisite Variety’ holds that cybernetic systems must have internal variety that matches their external variety so that they can self-organize to fight variation with variation. Both conservative and diversifying bet-hedging are the risk-avoiding and -spreading insurance strategies in response to environmental uncertainty. The bet-hedging strategy tries to cover all bases of plausible evolutionary scenios in an often unpredictable environment where it does not make sense to “put all eggs into one basket”. In this sense, variation is the bad/worst-case insurance strategy of risk-aversive individuals. Variation is pervasive at every level of biological organization and is created by a multitude of processes: mutagenesis, epimutagenesis, recombination, transposon mobility, repeat instability, gene expression noise, cellular network dynamics, physiology, phenotypic plasticity, behavior, and life history strategy. Importantly, variation is created condition-dependently, when variation is most needed – in organisms under stress. The bet-hedging strategy also manifests in a multitude of life history patterns: turnover of generations, reproductive prudence, iteroparity, polyandry, and sexual reproduction.

Cybernetic systems are complex systems. A system is a set of inter-related elements and a complex system is one in which, in plain English, the whole is greater than the sum of its parts (Byrne & Callaghan, 2014). Complexity is conceived as a system’s potential to assume a large number of states, i.e., variety. Nothing novel can emerge from systems with high degrees of order and stability — for example, crystals, incestuous communities, or regulated industries. On the other hand, complete chaotic systems, such as stampedes, riots, rage, or the early years of the French Revolution, are too formless to coalesce. Generative complexity takes place in the boundary between rigidity and randomness (Pascale et al., 2000). Only within an intermediate level of stochastic variation, somewhere between determined rigidity and literal chaos, local interactions can give rise to complexity. Complex systems have both stochastic and deterministic properties and, in fact, generate order from chaos. Nonlinearity, criticality, self-organization, emergent properties, scaling, hierarchy and evolvability are features of complex systems. Emergent properties are features of a complex system that are not present at the lower level but arise unexpectedly from interactions among the system’s components. Nonlinearity is a product of emergence (Byrne & Callaghan, 2014). The duality of stochasticity and selection is the organizing principle of evolution. Both are interdependent. The feedback between output and input signals inextricably intertwines both stochasticity and natural selection, and the individual- and population-levels of selection. Sexual reproduction with its generation of pre-selected variation is the paradigmatic bet-hedging enterprise and its evolutionary success reveals the selective signature of stochastic environments. Sexual reproduction is the proof of concept that (epi)genetic variation is no accidental occurrence but a highly regulated process and environmental stochasticity is its evolutionary “raison d’être”.

Stochastic environments change the rules of evolution. Lotteries cannot be played and insurance strategies not employed with single individuals. These are emergent population-level processes that exert population-level selection pressures generating variation and diversity at all levels of biological organization. Together with frequency and density-dependent selection, lottery- and insurance-dependent selection act on population-level traits. Within this theoretical framework, Spencer’s (1864) “the survival of the fittest” and Kimura’s “survival of the luckiest” (Kimura, 1989) are replaced by the “survival of the fitter and luckier”, emphasizing both the duality of stochasticity and natural selection.

14.1 Cooperation and multilevel selection

The phrase ‘group selection’ still evokes shudder and frowning among some evolutionary biologists (and incredible polemics, e.g. Dawkins, 2012) yet when properly applied it is mathematically equivalent to kin selection (Wade, 1978, 1979, 1980a, 1980b, 1985; Grafen, 1984; Queller, 1992b; Dugatkin & Reeve, 1994; Frank, 1998; Michod, 1999; Rousset, 2004; Foster et al., 2006; Lehmann et al., 2007b; Marshall, 2011). Thus, that kin selection is a type of group selection is defended by Hamilton (1975a, b), Uyenoyama and Feldman (1980), Nunney (1985), Queller (1992b), Sober and Wilson (1998) and others (Okasha, 2005) (but see van Veelen et al., 2012). Traulsen (2010), however, argued that the underlying mathematics of game theory as the basis of many multilevel selection models is fundamentally different from the approach of inclusive fitness.

Multilevel selection operates if, in addition to competition between individuals in a group, there is also competition between groups (Wilson, 1975; Boyd & Richerson, 1990; Sober & Wilson, 1998; Boyd et al., 2003; Traulsen & Nowak, 2006; Bowles, 2009; Bowles & Gintis, 2011). Typically, individual fitness and population fitness are in conflict. While selfish behavior is favored by individual selection, cooperation can evolve in many models of multilevel/group selection (Eshel, 1972; Wilson DS, 1975; Uyenoyama, 1979; Slatkin, 1981; Leigh, 1983; Wilson, 1983; Boyd & Richerson, 1990, 1992, 2002; Binmore, 1992, 1994a, b; van Baalen & Rand, 1998; Bergstrom, 2002; Goodnight, 2005; Killingback et al., 2006; Traulsen & Nowak, 2006; Nowak et al., 2010; see Heininger, 2015). According to Goodnight (2005), the evolution of an “altruistic” trait is driven by the ratio of the heritability at the group level to the heritability at the individual level. Experimental studies have shown that both artificial and natural group selection can be effective (Wade, 1976, 1977, 1982; Colwell, 1981; Craig DM, 1982; Goodnight, 1985; Breden & Wade, 1989; Bourke & Franks, 1995; Seeley, 1995; Stevens et al., 1995; Craig & Muir, 1996; Goodnight & Stevens, 1997; Miralles et al., 1997; Brookfield, 1998; Swenson et al., 2000; Goodnight, 2005).

Although couched in terms of inclusive fitness, the design of Griffin et al.’s (2004) and Kümmerli et al.’s (2009b) experiments are nearly identical to the classic group selection experiments of Wade (1976, 1977, 1982). As Queller (2004) pointed out in a comment to the work of Griffin et al. (2004): “Curiously, however, their experiment is perhaps more easily understood from a group-selection standpoint. The conditions of low and high relatedness correspond exactly to the presence and absence of within-group selection. The conditions of global and local competition correspond exactly to the presence and absence of between-group selection. The two-by-two crossing of these treatments therefore leads to the most basic group-selection experiment possible. The results confirm that cooperation is favored by between-group selection and disfavored by within-group selection.” The local competition treatment is analogous to the non-group selection controls of Wade, while the global competition treatment mirrors Wade’s group selection protocol. As predicted, when group selection is turned off, cooperation cannot evolve. These experiments, along with the present theoretical analysis, support the theoretical conclusion of Bijma & Wade (2008) that kin selection requires both relatedness and group selection.

15. Stochasticity and cooperation


Summary
In stochastic environments, reproductive success is unpredictable and highly variable.In taxa without parental brood care, particularly insects, larval survivorship is typically low. Even in eusocial insects with independent colony foundation, the vast majority of attempts to establish a colony will fail. The extent to which we opt for immediate gains over future rewards is known as future discounting. Future discounting is a functional and adaptive response to specific situations, particularly uncertainty or the low probability of reaping delayed benefits. Future discounting in adverse environments with uncertain reproductive success should have hampered the evolution of eusociality. If the basic assumptions of kin selection/inclusive fitness theory would be right, eusociality should have evolved in (i) less adverse environments with more predictable reproductive success, and (ii) in taxa with more predictable reproductive success and, hence, less uncertain inclusive fitness benefits.
An uncertain, unpredictable environment selects for evolutionary gambling, and either conservative or diversifying bet-hedging as individual- or population-level insurance to individual risk. A large body of literature shows taxonomic ubiquitous risk-sensitive behavior that is variance-sensitive behavior. In stochastic environments, cooperation trades individual fitness maximization for less variability and greater reliability of evolutionary outcomes. Both extrinsic and intrinsic noise/stochasticity can promote collective behavior in a variety of model and experimental systems.Cooperative breeding appears to be a risk-averse strategy to optimize fitness by reducing environmentally induced variance of reproductive success.
Given the manifold advantages of prosocial behavior at the population level and the manifold incentives to defect at the individual level, processes have evolved that mediate between the selection level conflicts. From microbes to humans, natural selection stabilizes cooperative behavior in resource-limited environments.A multitude of these interactions have been identified: direct and indirect reciprocity, sanctions, partner choice and fidelity.

An uncertain, unpredictable environment selects for evolutionary gambling, and either conservative or diversifying bet-hedging as individual- or population-level insurance to individual risk. In a model where kin selection and tag-based selection are dominant, mutualism is promoted by introducing environmental fluctuations (Uitdehaag, 2011). These fluctuations cause reduction in reproductive success by the mechanism of variance discount. The best strategy to counter variance discount is to share with agents who experience the most anticorrelated fluctuations, a strategy called bet-hedging. In this way, bet-hedging stimulates cooperation with the most unrelated partners, which is a basis for mutualism. Analytic results and simulations revealed that, if this effect is large enough, mutualistic strategies can dominate kin selective strategies. In addition, mutants of these mutualistic strategies that experience fluctuations that are more anticorrelated to their partner, can outcompete wild type, which can lead to the evolution of specialization. In this way, the evolutionary success of mutualistic strategies can be explained by cooperation based on bet-hedging strategies (Uitdehaag, 2011).

15.1 The unpredictability of reproductive success in stochastic environments

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). A life-history model predicted the occurrence of skipped reproduction only for intermediate environmental qualities, with high reproductive investment being optimal at both ends of a gradient of environmental quality (Fischer, 2009). Skipped reproduction is frequently observed in nature (in fish: Bull & Shine, 1979; Engelhard & Heino, 2005; Rideout et al., 2005; Jørgensen & Fiksen, 2006; Jørgensen et al., 2006; in amphibians: Bull & Shine, 1979; Harris & Ludwig, 2004; in reptiles: Bull & Shine, 1979; Brown & Weatherhead, 2004; in birds: Illera & Diaz, 2006). Poor individual condition and/or poor environmental quality are thought of as the main causes for skipped reproduction (Bull & Shine, 1979; Dutil, 1986; Rideout et al., 2005; Illera & Diaz, 2006). In taxa without parental brood care, particularly insects, the number of embryos entering a habitat is usually far in excess of its carrying capacity, and larval survivorship is typically low (e.g., Berryman, 1988; Ohgushi, 1991; Willis & Hendrick, 1992; Tinkle et al., 1993; Duffy, 1994; Dempster & McLean, 1998; Dixon et al., 1999) and unpredictable (Madsen & Shine, 1998; Fincke & Hadrys, 2001; Haugen, 2001; Rollinson & Brooks, 2007). Ecological factors such as deterioration of larval habitats or fluctuations in the density of food, predators, cannibals, or parasites can result in unpredictable windows of offspring survivorship (e.g., Smith, 1987; Newman, 1989; So & Dugeon, 1989; Morin et al., 1990; Messina, 1991; Anholt, 1994; Dixon et al., 1999). In insects, “while lifetime egg production is largely determined by chance” (Thompson, 1990), the numbers of mature offspring produced (fitness) is largely unpredictable (Fincke & Hadrys, 2001) and in natural populations, crucially, is poorly correlated with behavioral observations of mating, particularly for females (Thompson et al., 2011).

Hamilton’s rule is a deterministic statement: A cooperative behavior can only evolve if the benefit-to-cost ratio of the “altruistic” act is larger than the relatedness between the donor and recipient (Hamilton 1964; Traulsen, 2010). An obstacle for the evolution of “altruistic” behavior in the population is that the costs must be paid now but the payoff in terms of inclusive fitness lies more or less far in the future. Since natural selection lacks foresight and is a blind causal process operating over random mutations, it can only operate on actual past causal contributions (Neander, 1991). But how could a trait evolve in the first place when the costs paid by the donor are certain but the benefits for the recipient (and hence the inclusive fitness gains of the donor) in the future may be highly uncertain, adding to the donor’s uncertainties regarding kin recognition (see chapter 13) and recipient cheating (Sachs & Rubenstein, 2007; Chapuisat, 2009; Heinze & d'Ettorre, 2009)? And the choice between cooperation/defection of the recipient (in terms of its reproductive success) may not even be under the control of the recipient but of an uncertain, unpredictable environment. In addition, the environments that coerce individuals into some type of cooperative behavior/”altruism” can be characterized as uncertain, harsh ecological conditions that constrain independent breeding (see chapter 5) and render reproductive success unpredictable. Particularly in insects, larval survivorship is typically low (e.g., Berryman, 1988; Ohgushi, 1991; Willis & Hendrick, 1992; Tinkle et al., 1993; Duffy, 1994; Dempster & McLean, 1998; Dixon et al., 1999) and unpredictable (Madsen & Shine, 1998; Fincke & Hadrys, 2001; Haugen, 2001; Rollinson & Brooks, 2007).

Eusocial insect colonies typically go through three different phases in their life cycle: a founding stage in which new nests are founded by recently mated queens (and kings, in the case of termites), an ergonomic growth stage in which the colony produces workers only and increases in size, and a reproductive stage in which the colony produces new sexuals (males and gynes) (Oster & Wilson 1978). The transition from ergonomic growth to reproduction is generally marked by the attainment of a certain colony size, which varies from species to species. The cooperation that makes eusocial species so successful is, however, conspicuously absent when new colonies begin: The mortality of the eusocial foundresses is very high from the moment they leave the mother nest and mate until they finish constructing a nest. The agents of the increased mortality during exposure are many, both biological (chiefly through predation) and physical (because of the scarcity of nest sites and hour-by-hour vicissitudes in the environment). The large magnitude of the resulting hecatomb of would-be foundresses has been commonly observed in species that release large numbers in nuptial flights (Hölldobler & Wilson, 2014). Most species use independent colony foundation during which the queen (queen and king in termites) is alone as she attempts to establish a new nest and produce the first generation of workers. This vulnerable solitary stage can last several weeks, and the vast majority of attempts will fail (Myles, 1988; Hölldobler & Wilson, 1990, 2008; Greene, 1991; Tschinkel, 1992a, 1992b, 2006; Herbers, 1993; Peeters, 1997; Brown & Bonhoeffer, 2003; Korb & Schneider, 2007; Cole, 2009; Peeters & Molet, 2010; Cronin et al., 2013). Tschinkel (2006), for example, has described the death of the vast majority of imported fire ant (Solenopsis invicta) queens from the first minute they take flight, subjecting themselves to “predation by birds and insects and the chance of heat death, starvation, execution, and usurpation.” As the queens settle to the ground, “a further fraction is taken by ground-based predators, especially other ants” (Tschinkel, 2006). When predators are abundant, as few as five percent of the queens succeed in building a nest (Whitcomb et al., 1973). Similar mortality rates have been observed in the mating swarms of other ant species with large mature colonies, including representatives of the genera AttaPheidole, and Pogonomyrmex (Hölldobler & Wilson, 2009, 2011). Survival is particularly poor in environments that provide low or unpredictable resources. In a long-term study of colony founding in the red harvester ant, Pogonomyrmex barbatus, less than 1% of queens were estimated to succeed (Gordon & Kulig, 1996; Ingram et al., 2013). High rates of mortality also exist for dispersing queens of primitively eusocial species, such as the facultatively eusocial wasps. Among 19 species studied, 38–100% of the nests constructed by lone foundresses, who then were subjected to high risk both on the nest and during foraging, failed before the first brood emerged (Queller, 1996). Even in honeybees with dependent colony foundation by swarms (Cronin et al., 2013), a 70% mortality of swarms before they found a nest has been reported (Oldroyd et al., 1997). It is hardly conceivable that a behavior, the helper’s “altruism”, could have played a role in the evolution of eusociality when its putative evolutionary raison d’être, the worker’s inclusive fitness gain, can be more or less annihilated by an uncertain and stochastic environment.

15.1.1 Discounting future uncertain benefits

Intertemporal choices–decisions involving tradeoffs among costs and benefits occurring at different times–are important and ubiquitous. It is intuitively obvious that the subjective value of an outcome decreases as the time until its occurrence increases and that subjective value also decreases as the odds against an outcome increase: People would generally prefer to receive $100 now rather than in a month and would prefer $100 for sure rather than a l-in-10 chance of receiving the same amount (Green et al., 1999). When offered the choice of a smaller reward now or a larger reward at a certain point in the future, some are better than others at resisting the temptation of instant gratification. But when the long-term future is uncertain, it’s rational to discount it and take what you can get today (MacGregor, 2014). The extent to which we opt for immediate gains over future rewards is known as future discounting (Frederick et al., 2002; Daly & Wilson, 2005). Future discounting is a functional and adaptive response to specific situations, particularly uncertainty or the low probability of reaping delayed benefits (Rogers, 1994; Charlton, 1996; Wilson & Daly, 1997). Studies with animals, such as wasps, honeybees, and scorpionflies, have shown that organisms tend to discount the future under certain situations (i.e., they bet on risky behaviors and present gains based on environmental cues that indicate low life expectancy or narrow future options) (Roitberg et al., 1992, 1993; Woyciechowski & Kozlowski, 1998; Engqvist & Sauer, 2002). For example, reciprocal altruism (Trivers, 1971) can evolve as long as the cost of aiding another individual is outweighed by the benefit of receiving aid from that individual later, devalued by the probability that aid will be returned (Boyd, 1992). Among nonhuman primates, investigations demonstrated that feeding ecology could explain some differences between species with regard to the capacity to wait for rewards (Stevens et al., 2005a). Other studies verified that monkeys avoid risky options with increasing delay (Hayden & Platt, 2007), and they tend to choose a small reward when the large one is delayed (Hwang et al., 2009). Apes show a higher ability for future-oriented decisions than other animals, presenting temporal discounting preferences that are similar to humans under comparable conditions (Rosati et al., 2007). Individuals who grow up in environments where resources are scarce, competition is intense, and mortality is high should discount the future more heavily than individuals who grow up in abundant, supportive, long-lived locales (Wilson & Daly, 1997). Studies demonstrated an association between socioeconomic deprivation and faster life-history strategies involved in reproductive decisions (Ellis et al., 2009; Nettle, 2010), environmental unpredictability and risk taking (Hill et al., 2008), and between low income and choices for smaller-sooner amounts of money (Lawrance, 1991; Loewenstein & Elster, 1992; Green et al., 1996; Harrison et al., 2002; Reimers et al., 2009; Leitão et al., 2013). A recent study with a Brazilian sample showed that slum-dwelling young people discounted the future more than university students from middle-class neighborhoods in Rio de Janeiro, indicating that the young adapted their psychology and behavior to their living conditions (Ramos et al., 2013). Individuals with low discount rates will be more likely to cooperate than people with high discount rates. In other words, patient people will be more trusting, and more trustworthy (Curry et al., 2005).

In economy, when the costs have to be paid now and the benefits of this investment can only be reaped in the distant, particularly intergenerational, future the benefits have to be discounted taking into account the uncertainty of the benefits of the investment (Arrow et al., 2012). With simplifying assumptions, this leads to the Ramsey discounting formula. The Ramsey formula results in a declining certainty-equivalent discount rate if the rate of growth in consumption is uncertain and if shocks to consumption are correlated over time (Cropper, 2012). Overall, the discount factor, the reciprocal of the factor by which capital would grow over unit time by compound interest, should be rising with the long-term nature of the investment. Brown (1987) suggested that inclusive fitness might be usefully broken into direct and indirect, current and future components. The most difficult components of inclusive fitness to conceptualize and estimate are the future effects; several authors discussed the ways in which ‘future indirect’ fitness components might arise (Mumme et al., 1989; Creel, 1990b; Lucas et al., 1997). Emlen & Wrege (1991) and Creel et al. (1991) suggested partial solutions to the problem of estimating future fitness components, both direct and indirect. However, under certain conditions, Creel’s approach yields even negative fitness (Lucas et al., 1996).

There are clusters of correlated life history traits (e.g., timing of puberty, age at sexual debut and first birth, parental investment strategies) that lie on a slow-to-fast continuum (Ellis et al., 2009; Heininger, 2012). Harshness (externally caused levels of morbidity-mortality) and unpredictability (spatial-temporal variation in harshness) are the most fundamental environmental influences on the evolution and development of these life history strategies (Ellis et al., 2009). This continuum indicates that microbes, animals and plants sense environmental stochasticity and time their reproductive investments according to their future discounting (Ellis et al., 2009). Thus, the standard life history response to increased environmental stochasticity is discounting of future reproduction (Klinkhamer et al., 1997; Williams & Day, 2003; Jones et al., 2008; Ratcliff et al., 2009; Heininger, 2012; Frank, 2014). Insects appear also to be able to discount the future (Roitberg et al., 1992, 1993; Woyciechowski & Kozlowski, 1998; Engqvist & Sauer, 2002). In eusocial societies, individual worker lifespan is significantly shorter (e.g. in honeybees 15 to 38 days) than the reproductive interval of the colony (2–3 swarms/colony per year). Within the conceptual framework of inclusive fitness theory (and if the analogy of economic and evolutionary reasoning holds) agents should exert transgenerational discounting of inclusive fitness benefits. Stochasticity-related devaluation charges of inclusive fitness returns for new workers in newly founded colonies (and their incentive to reproduce) should be higher than that for workers towards the end of a reproductive cycle caring for new reproductives, resulting in an increased selfish reproduction of workers during the founding period of a colony when the long-term inclusive fitness benefits of helping are particularly uncertain. However, no such colony-reproductive-cycle-specific worker behavior in terms of selfish reproduction has been reported to date. This, again, questions the validity of the inclusive fitness theory. If the basic assumptions of kin selection/inclusive fitness theory would be right, eusociality should have evolved in (i) less adverse environments with more predictable reproductive success, and (ii) in taxa with more predictable reproductive success and, hence, less uncertain inclusive fitness benefit. Future discounting in adverse environments with uncertain reproductive success should have hampered the evolution of eusociality.

Imagine, a close relative asks you for a big favor (so big that if you meet her/his request your own aspirations would be impaired) and she/he promises to reward your investment in the future. You know from past experience that there is a 1:50 chance that she/he will keep her/his promises. Would you do her/him this favor? Probably not if you are not a complete idiot. And for sure, evolution is not such an idiot (remember Orgel’s rule).

15.2 Cooperation as bet-hedging response to environmental stochasticity

15.2.1 Empirical evidence

Uncertainty can be measured as the variance of a distribution of environmental quality, and adversity as the mean (Andras et al., 2003; Fronhofer et al., 2011). Both adversity and uncertainty have been conceptualized as aspects of environmental ‘risk’ (Daly & Wilson, 2002). A large body of literature now has shown taxonomic ubiquitous risk-sensitive behavior (Heininger, 2015). Risk-sensitive behavior is variance-sensitive behavior (Smallwood, 1996; Ydenberg, 2007; Mayack & Naug, 2011; Ratikainen, 2012). Cooperation as a risk-spreading response to environmental stochasticity is a byproduct of selfish action (West-Eberhard, 1975; Brown, 1983; Connor, 1986; Morris et al., 2012). Both extrinsic and intrinsic noise/stochasticity can promote collective behavior in a variety of model and experimental systems (Chen et al., 2005; Rudge & Burrage, 2008; Yates et al., 2009). Environmental variation selects for cooperative behavior (Rubenstein & Lovette, 2007; Rubenstein, 2011). Empirical evidence from plant and animal communities and human societies suggests that there is a correlation between adverse conditions and cooperation (e.g. Knight, 1984; Wilkinson, 1984; Rytkönen & Soppela, 1995; Avilés, 1999; Spinks et al., 2000; De Bono et al., 2002; Andras et al., 2003, 2007; Spieler, 2003; Krams et al., 2010; Shen et al., 2012). Social species disproportionately occur in heterogeneous and unpredictable environments (Wilson, 1971; Solomon & French, 1997; Costa, 2006; Jetz & Rubenstein, 2011). In more adverse and fluctuating environments individuals perceive their resources to be more unpredictable, and this unpredictability favors cooperation (Avilés, 1999; Andras & Lazarus, 2005; Andras et al., 2007; Shen et al., 2012). In addition to the mean amount of food, the variance of a given resource influences foraging decisions (Caraco, 1980; Caraco et al., 1980; Stephens, 1981; Real & Caraco, 1986; Schmitz & Ritchie, 1991). For example, game in the optimal diet breadth has many high-yield but high-variance targets so that one may go several days without capturing sufficient game to meet the dietary needs of a hunter’s family. Sharing among hunters provides a solution to the problem by reducing the variance across hunters through pooling risk (Kaplan et al., 1984; Kaplan & Hill, 1985; Hames, 2004). Using an experimental design in which food distribution was either clumped or dispersed, in combination with individuals that varied in exploratory behavior, Tanner and Jackson (2012) showed that social structure can be induced in the otherwise non-social European shore crab (Carcinus maenas). Prey consumption variance is reduced in group foraging (Caraco, 1981; Clark & Mangel, 1986; Caraco et al., 1995), i.e. foraging success becomes more reliable in groups. Risk-sensitive foraging theory (Bateson, 2002; Bednekoff, 1996; Kacelnik & Bateson, 1996; Smallwood, 1996) can be used to explain cooperative foraging and breeding (Poethke & Liebig, 2008), as in social spiders (Wenzel & Pickering, 1991; Caraco et al., 1995; Avilés, 1997; Whitehouse & Lubin, 2005; Agnarsson, 2006). In the context of risk-sensitive foraging theory, group formation has been interpreted as a mechanism of risk avoidance (Caraco, 1981; Clark & Mangel, 1986; Wenzel & Pickering, 1991; Caraco et al., 1995; Uetz, 1996; Uetz & Hieber, 1997; Fronhofer et al., 2011). Spatial self-structuring leads to selection of cooperative traits (Boerlijst & Hogeweg, 1991; Kümmerli et al., 2009a).

Using a synthetic yeast cooperator-cheater system engineered to supply or exploit essential nutrients, A.J. Waite (2013) found in his doctoral thesis that despite cooperators being less fit than cheaters, some cocultures rapidly purged cheaters, while in others cheaters dominated and drove cocultures and themselves extinct. The nutrient-limited cooperative environment acted as a stress on both cooperators and cheaters and this stress created a selective pressure which triggered an “adaptive race” between cooperators and cheaters to sample from the same pool of fitness-enhancing mutations. Although cooperators and cheaters sample from the same pool, this symmetry is soon broken: The best cooperators purge cheaters and continue to grow, whereas the best cheaters cause rapid self-extinction due to the tragedy of the commons (Waite & Shou, 2012). In a model, Waite (2013) showed that stress can synergize with a spatially-structured environment to promote cooperation. Together with Ben Kerr’s lab at University of Washington, he showed that cooperation in the natural cooperative Pseudomonas aeruginosa can be stabilized even when spatial structure is random and transient, but only when an environmental stress–antibiotics in this case–elicits an adaptive race between cooperators and cheaters (Waite, 2013). Social unicellular prokaryotes and eukaryotes (see chapter 5) and the nematode Caenorhabditis elegans aggregate in response to stressors (De Bono et al., 2002); animals form more cohesive or larger groups, with consequent greater mutualistic benefits under greater predation risk (Seghers, 1974; Farr, 1975; Dunbar, 1988; Lima & Dill, 1990; Spieler, 2003; Heg et al., 2004; Zöttl et al., 2013b); mole-rats delay dispersion more in arid than in mesic habitats (Spinks et al., 2000); human in-group solidarity is greatest when the group is under threat or in a harsh environment (Levine & Campbell, 1972; Goody, 1991; Hogg, 1992; Hewstone et al., 2002).

15.2.1.1 Cooperative breeding

Cooperative breeding behavior appears to be a risk-averse strategy to maximize fitness by reducing environmentally induced fecundity variance. Such a within-generation bet-hedging hypothesis for social evolution predicts that (i) variance in reproductive success should be related to environmental variation, (ii) variance in reproductive success should be related to the potential for cooperation in a group, and (iii) the potential for cooperation should be related to environmental variation (Rubenstein, 2011). It has been proposed that there may be a link between bet-hedging as response to environmental uncertainty and cooperative breeding (Cockburn & Russell, 2011).

Consider a situation where two individuals nesting together can rear more brood than the sum of their individual capabilities and further that the roles of fertile queen and sterile worker among these two individuals is decided by the toss of a coin, so to speak (Gadagkar & Bonner, 1994). In such a situation an average individual that nests in the group may obtain more fitness than a solitary individual. If we replace the concept of alleles programming individuals into workers with alleles programming individuals to take the risk of being part of the group, then it is possible to show that under certain ecological conditions the ‘gamblers’ will be fitter than the risk-averse solitary individuals (Gadagkar, 1990d, 1991a, 1994b). The losers in the ‘gamble’ will leave behind no offspring and we will see them as sterile workers. Nevertheless, it is a form of mutualism that has given rise to this situation. One advantage of such a model is that it requires no assumption of increased genetic relatedness or parental manipulation. The fact that foundress associations in primitively eusocial species may consist of distant relatives and that workers can become queens before they die suggests that the possibility of direct reproduction at some future point of time may itself drive the evolution of group living. Likewise, Seger (1993) argued that cofoundresses that initially join the group may put their lives at risk in a gamble in which, however, they may be also the one to survive (Seger, 1993). An important assumption of this so called ‘gambling hypothesis’ is that the per capita productivity in the group mode is higher (or more reliable; see Wenzel & Pickering, 1991) than in the solitary mode. Even if the probability of becoming a queen may be quite small for any given individual, the fitness gained by those who succeed can be so great that it offsets the cost incurred by the other, unsuccessful bearers of the hypothetical ‘gambling’ allele, which makes its bearers stay in a social group and await their chances of becoming queen (Gadagkar, 1990d, 1991a, 1994b).

Increased group productivity and strong ecological constraints have previously been suggested to favor pleometrotic associations in social insects, for example faster colony growth (Bartz & Hölldobler, 1982; Tschinkel & Howard, 1983; Rissing & Pollock, 1987; Sommer & Hölldobler, 1995), increased protection against predators and usurpation (Gamboa, 1978; McCorquodale, 1989; Balas & Adams, 1996) and a shortage of nest sites (e.g. Pfennig, 1995). In the wasp Ropalidia marginata a solitary foundress produces on the average no more than one or two offspring whereas a queen of a multi-female colony produces on the average 76 offspring (Gadagkar 1990a). Simultaneous polygyny in R. cyathiformis (Gadagkar & Joshi, 1982), serial polygyny due to queen supersedure, the process by which a colony of bees/wasps replaces its queen without swarming, in R. marginata (Gadagkar et al., 1993), the causes and consequences of colony fission in R. cyathiformis (Gadagkar & Joshi, 1985), the complex behavioral caste differentiation in R. marginataR. cyathiformis (Gadagkar & Joshi, 1983, 1984), and in Polistes versicolor (Gadagkar, 1990b) all add up to make mutualism and gambling plausible models for the origin of sociality in primitively eusocial species such as R. marginata and R. cyathiformis (Gadagkar, 1990d, 1991a, 1994b). In the primitively eusocial wasp species Belonogaster juncea juncea multiple foundress colonies were significantly more successful than single foundress colonies in producing at least one adult (Tindo et al., 2008). The total productivity of the colonies increased significantly with the number of associated foundresses, but the productivity per capita did not. No single foundress colony (out of 13) reached the sexual phase, while eight (out of 36, 21.6%) multiple foundress colonies did. The decreasing per capita productivity concomitant with an increasing number of females noted in this study illustrates once again Michener’s paradox (1964). The coefficient of variance of the per capita productivity significantly decreased with group size, as Wenzel and Pickering (1991) suggested in the model they created to explain the paradox (Tindo et al., 2008). Several studies reported a survival advantage of multiple foundress colonies compared with single foundress colonies of the genera Polistes (Metcalf & Whitt, 1977; Gibo, 1978; Tibbetts & Reeve, 2003), Belonogaster (Keeping & Crewe, 1987; Tindo et al. , 1997a), and Ropalidia (Shakarad & Gadagkar, 1995), Allodapine bees (Hogendoorn & Zammit, 2001) and social shrimps (Duffy, 2002). Likewise, the increase in total productivity as a function of group size is in line with previous findings reported on primitively eusocial species (Michener, 1964; Shakarad & Gadagkar, 1995; Tindo et al., 1997b; Tibbetts & Reeve, 2003). The decrease in per capita productivity as a function of group size, however, illustrates Michener’s paradox in primitively eusocial insects (Michener, 1964; Noonan, 1981; Strassmann et al., 1988; Shakarad & Gadagkar, 1995; Gadagkar, 1996; Hogendoorn & Zammit, 2001; Seppä et al., 2002; Soucy et al., 2003). A widely accepted explanation for Michener’s paradox is the model by Wenzel and Pickering (1991) suggesting that individuals in larger groups might trade lower per capita productivity for less variability and greater predictability. Risk-sensitive reproductive strategies may reduce the average (arithmetic mean) of individual reproductive output, while yet maximizing the population geometric mean; this trade-off in terms of average reproduction is ‘bet hedging’ (Slatkin, 1974; Seger & Brockmann, 1987; Philippi & Seger, 1989; Simons, 2009). To understand the adaptive value of cooperative breeding behavior in heterogeneous and unpredictable environments, both the mean and environmentally induced variance in reproductive success has to be taken into account. In addition to within-generation bet-hedging, social evolution has to be considered in the context of among-generation bet-hedging, or risk spreading over multiple generations to maximize geometric mean lifetime reproductive success (Rubenstein, 2011).

Stochasticity promotes cooperation at least at two levels. Stochastic availability of resources stabilizes cooperative acquisition of resources. Stochastic determination of fate e.g. in the spore-stalk lottery of D. discoideum (Chattwood & Thompson, 2011) and in cell differentiation in organisms ranging from bacteria to humans (Losick & Desplan, 2008) ensures participation in the lottery. Lotteries are considered fair procedures. In fact, when a fair (unbiased) procedure is feasible and a fair allocation is not, selection of a fair procedure is acceptable whereas there is resistance to the imposition of biased outcomes (Bohnet et al., 2008), much as when a fair (unbiased) allocation is available (Rawls, 1971 p. 86; Bolton et al., 2005). Certainly, none of the losers would voluntarily take part in a “lottery” when the outcome is biased, e.g. if the winners and losers in that lottery would be determined by a weighted dice.

15.2.2 Evidence from models

It is now generally accepted that stochastic models are necessary to properly capture the multiple sources of heterogeneity needed for modeling biosystems in a realistic way (Wilkinson, 2009). Cooperating groups might provide mutual advantages for their members, so that the net benefits to all participants outweigh the costs. In essence, game theoretic models of cooperation tacitly postulate (see e.g. Axelrod & Hamilton, 1981; Maynard Smith, 1984, 1989), that cooperation is not equivalent to altruism and does not by definition require sacrifices, or genes for altruism, which is why game theory formulations are largely indifferent to the degree of relatedness, if any, between the cooperators (Corning, 1997).

As the most stringent situation of reciprocal behavior through pairwise interactions, the prisoner’s dilemma has long been considered as a paradigmatic example for studying the dilemmas between individual interests and collective welfare. In its original form, the prisoner’s dilemma is a two-player non-zero-sum game, where each player decides simultaneously whether to cooperate (C) or defect (D) without knowing a priori how its opponent will act. There are four possible outcomes for this game: (1) mutual cooperation (C, C) yields the largest collective payoff by offering each a reward R, (2) mutual defection (D, D) pays each a punishment P, and (3) the mixed choices (C, D) or (D, C) give the defector a temptation T and the cooperator the sucker’s payoff S (Zhang et al., 2011; Stewart & Plotkin, 2013), with the payoff ranking satisfying T>R>P>S and 2R>T+S (Li & Duan, 2014). A basic evolutionary problem posed by the Iterated Prisoner’s Dilemma game is to understand when the paradigmatic cooperative strategy Tit-for-Tat can invade a population of pure defectors. Deterministically, this is impossible (Doebeli et al., 1997). Historically, the evolution of cooperation has been studied in the absence of diversity (Hardin, 1968; Axelrod & Hamilton, 1981; Hofbauer & Sigmund, 1998; Nowak & Sigmund, 2004; Nowak, 2006a, 2006b; Sigmund, 2010). It was found more reasonable to deal with the evolution of cooperation in a population of (primitively) identical individuals, as defined by conventional evolutionary game theory (EGT) (Maynard Smith, 1982; Hofbauer & Sigmund, 1998; Sigmund, 2010). However, the dynamical randomness stemming from diverse factors of stochasticity in a variety of interactions and biological systems constitutes an important mechanism to establish and maintain cooperation and mutualism (Fudenberg & Maskin, 1990; Nowak, 1990; Nowak & Sigmund, 1990, 1992, 1993a, 1993b; Bendor et al., 1991; Goodnight, 1992; Nowak et al., 1994, 1995, 2004; Wu & Axelrod, 1995; Doebeli & Knowlton, 1998; Wahl & Nowak, 1999; Mitteldorf & Wilson, 2000; Eriksson & Lindgren, 2002; Lorberbaum et al., 2002; McNamara et al., 2004; Imhof et al., 2005; Szabó et al., 2005; Lehmann et al., 2006; Perc, 2006a, 2006b, 2007a; Perc & Marhl, 2006; Vukov et al., 2006; Ren et al., 2007; Tanimoto, 2007; Chen X et al., 2008a, b; Perc & Szolnoki, 2008; Santos et al., 2008, 2012; Szolnoki & Perc, 2008; Szolnoki et al., 2008; Helbing & Yu, 2009; Wu & Holme, 2009; Iliopoulos et al., 2010; Imhof & Nowak, 2010; Jia et al., 2010; Wu et al., 2010; Yaari & Solomon, 2010; Delton et al., 2011; Fronhofer et al., 2011). It was noted that “…stochasticity itself is a factor favoring the selection of altruism” (Mitteldorf & Wilson, 2000).

Cumulative evidence from models supports the intuition that complex networks are also entities with some degree of stochastic dynamics such that the coevolution of the networks with the dynamical processes they host strongly influences the overall system behavior (Bornholdt & Rohlf, 2000; Skyrms & Pemantle, 2000; Ito & Kaneko, 2001; Skyrms, 2004; Gross et al., 2006; Holme & Newman, 2006; Kossinets & Watts, 2006; Pacheco et al., 2006; Santos et al., 2006b; Garlaschelli et al., 2007; Szabó & Fáth, 2007; González et al., 2008; Gross & Blasius, 2008; Fehl et al., 2011; Rand et al., 2011; Wang J et al., 2012). Complex networks, having the connectivity structure and stochastic dynamics similar to that of social networks, are very beneficial for the evolution of cooperation (Abramson & Kuperman, 2001; Santos & Pacheco, 2005; Ohtsuki et al., 2006; Santos et al., 2006a, 2008, 2012; Tang et al., 2006; Gómez-Gardeñes et al., 2007, 2008; Lehmann et al., 2007a; Taylor et al., 2007; Gómez-Gardeñes et al., 2008; Poncela et al., 2007; Tomassini et al., 2007; Kuperman & Risau-Gusman, 2008; Lozano et al., 2008; Perc & Szolnoki, 2008; Wang S et al., 2008; Du et al., 2009; Pacheco et al., 2009; Szolnoki & Perc, 2009; McNamara & Leimar, 2010; Tomassini & Pestelacci, 2010; Fu et al., 2010; de Weerd & Verbrugge, 2011; Du & Fu, 2011; Pruitt et al., 2012). Thus, populations in which individuals exhibit diversity in behavior or handling their social contacts end up being more cooperative than those in which individuals exhibit no such diversity (McNamara et al., 2004; Van Segbroeck et al., 2009).

More generally, heterogeneity or diversity allows for cooperative behavior to prevail even if the temptations to defect are large (Fishman et al., 2001; Brockhurst et al., 2006; Perc, 2007b; Szabó & Fáth, 2007; Fu & Wang, 2008; Perc & Szolnoki, 2008; Santos et al., 2008, 2012; Szolnoki et al., 2008). Likewise, heterogeneity/diversity of resource availability, acquisition or investment (Killingback et al., 1999; Killingback & Doebeli, 2002; Chadefaux & Helbing, 2010; Santos et al., 2012; Kun & Dieckmann, 2013) will promote the evolution of cooperation. In nature, payoffs for certain acts vary according to the levels of supply and demand for various commodities–e.g. the need for cooperation or help (Noë et al., 1991; Dunbar, 1992; Barrett et al., 1999; Stopka & Macdonald, 1999)–and variation in the payoffs are more likely to coerce individuals into mutual cooperation (Stopka & Johnson, 2012). Likewise, spatial structure will promote the evolution of cooperation (Nowak & May, 1992; Nowak et al., 1994; Ferriere & Michod, 1996; van Baalen & Rand, 1998; Brauchli et al., 1999; Pfeiffer et al., 2001; Boza & Scheuring, 2004; Kreft, 2004b; Grim et al., 2006; Langer et al., 2008). Again, there is “overwhelming evidence indicating that heterogeneity, almost irrespective of its origin, promotes cooperative actions” (Perc & Wang, 2010). On the other hand, cheating will prevail in homogeneous populations (Axelrod & Hamilton, 1981; Pfeiffer et al., 2001; Frick & Schuster, 2003; Kreft, 2004b).

After years of attempting to explain cooperation by dyadic models assuming static individuals (both developmentally within the same individual and across different individuals), it has been acknowledged that cooperation may involve more than two individuals (Brosnan & Bshary, 2010; Connor, 2010; Earley, 2010). Game theory and other models that incorporate choice among multiple potential partners (and hence another stochastic element) are more realistic representations of many mutualisms than are the two-player Prisoner’s Dilemma scenarios (Noë & Hammerstein, 1995; Leimar & Hammerstein, 2010). Game theory models allow for broader stability of mutualistic interactions when multiple players are incorporated into the model (Bull & Rice, 1991; Noë & Hammerstein, 1994, 1995). While models of mutualistic communities show little dependence on stochastic population fluctuations, predator-prey models show strong dependence on the stochasticity that is a major cause of population extinctions (Murase et al., 2010).

Stochasticity, although only rarely explicitly stated, is also intrinsic to model scenarios like iterated games (Axelrod, 1984; Aumann & Maschler, 1995; Nakamaru et al., 1997). Cooperation can be supported by repeated interactions (Trivers, 1971; Axelrod & Hamilton, 1981; Axelrod, 1984). The most important condition necessary for the evolution of direct reciprocity is that interactions between pairs of agents be sufficiently repeated (Axelrod & Hamilton, 1981). For reciprocity to operate, after one agent delivers a benefit, the partner must forgo the immediate gain offered by cheating–that is, of not incurring the cost involved in returning a comparable benefit. In general, selection can only favor forgoing this gain and incurring the cost of reciprocating when the net value to the partner of the future series of exchange interactions (enabled by reciprocation) exceeds the benefit of immediate defection (which would terminate that future series). If there were no future exchanges–if an interaction was one-shot–then the equilibrium strategy would be always defect. In deciding whether to engage in dyadic reciprocity, these systems must balance (i) the costs of mistaking a one-shot interaction for a repeated interaction (hence, risking a single chance of being exploited) with (ii) the far greater costs of mistaking a repeated interaction for a one-shot interaction (thereby precluding benefits from multiple future cooperative interactions). This asymmetry builds organisms naturally selected to cooperate even when exposed to cues that they are in one-shot interactions (Delton et al., 2011). Stochasticity may, at least in part, be introduced by the compelling empirical evidence of random variation in individuals’ decisions (Boyd, 1989; Camerer, 1989; Starmer & Sugden, 1989; Hey & Orme, 1994; Wu, 1994; Wu & Axelrod, 1995; Ballinger & Wilcox 1997; Hey, 2001; Eriksson & Lindgren, 2002; Blavatskyy, 2007; Van Segbroeck et al., 2009; Iliopoulos et al., 2010; Cavaliere & Poyatos, 2013) and as many competitive games favor “mixed” (stochastic) strategies (Miller, 1997). For example, Camerer (1989) reports that 31.6% of subjects reversed their choices when presented with the same binary choice problem for the second time. Starmer and Sugden (1989) found that 26.5% of all choices are reversed on the second repetition of a decision problem. Hey and Orme (1994) report an inconsistency rate of 25% even when individuals are allowed to declare indifference. Wu (1994) found that 5% to 45% of choice decisions are reversed (depending on a lottery pair) when a decision problem is repeated. Ballinger and Wilcox (1997) report a median switching rate of 20.8%. There are a number of systems from single neurons and synapses (Lowen et al., 2001; Stein et al., 2005) to invertebrate (Brembs et al., 2002; Briggman et al., 2005; Maye et al., 2007) and vertebrate animals including humans (Grobstein, 1994; Glimcher, 2005; Raichle, 2006), which even generate variable output despite no variations in input at all, leading to difficulties reproducing even tightly controlled experiments (Crabbe et al., 1999). As such, dynamic heterogeneity/diversity is instrumental not only to promote cooperation, but also to sustain it, even in the absence of complex community enforcement mechanisms, reputations or punishment (Hamilton, 1964; Wilson DS, 1975; Ostrom, 1990; Fehr & Gächter, 2000; Milinski et al., 2002; Skyrms, 2004; Nowak, 2006a; West et al., 2007a; Sigmund et al., 2010).

While in an additive context the emergence and survival of cooperation requires special conditions (especially some level of reward, punishment, reciprocity), Yaari & Solomon (2010) found that in the multiplicative random context the emergence of cooperation is much more natural and effective. The fundamental feature of multiplicative processes is the fact that the expected gain of the players taking part in this iterative process depends in a crucial way on the number of players considered (number of independent realizations) and the number of time steps that the game is played. For long times (the number of time steps played in the game), the expected wealth of the players follows the geometric mean and not the arithmetic mean of the game (keep in mind that “geometric mean ≤ arithmetical mean”) (Yaari & Solomon, 2010).

15.3 Natural selection stabilizes cooperation

Given the manifold advantages of prosocial behavior at the population level (Wilson & Wilson, 2007a, b) and the manifold incentives to defect at the individual level, processes should have evolved that mediate between the selection level conflicts. In microbes, sociality is under strong natural selection. In constant, nutrient-rich environments where benefits associated with sporogenesis (a social behavior) are absent and no longer balance the cost of constructing spores, sporulation ability of Bacillus subtilis was lost over 6,000 generations (Maughan et al., 2009). Propagation of B. subtilis for less than 2,000 generations in a nutrient-rich environment where sporulation is suppressed led to rapid initiation of genomic erosion including biosynthetic pathways, sporulation, competence, and DNA repair (Brown et al., 2011). The social prokaryote Myxococcus xanthus loses its social behavior when propagated in nutrient-rich habitats in which their social behaviors for starvation-induced spore production or predatory efficiency were not under positive selection (Velicer et al., 1998; Velicer & Stredwick, 2002). Social behavior is an evolutionary burden in asocial habitats (Velicer et al., 1998; Velicer & Stredwick, 2002; Zhang et al., 2005; Li et al., 2011). Kraemer et al. (2010) argued that the globally widespread presence of developmentally competent strains (Vos & Velicer, 2008) indicates that social proficiency is highly beneficial in many habitats. Natural variation in developmental timing of strains recovered from different sites, suggests that variation in selective forces across different environments may contribute to the persistence of such variants (Kraemer et al., 2010).

Corning (1996) listed some of the common assumptions in Prisoner’s Dilemma games: “The games are always voluntary and “democratic”; each player is free to choose his/her own preferred strategy, and the opposing player has no means available for coercing choices, or compliance. Also, the players are not allowed to communicate with one another in an effort to reduce the uncertainties in the interactions. Furthermore, defectors are usually rewarded handsomely for cheating while the co-operators are denied the power to prevent defectors from enjoying the rewards, much less punishing them for defection. Such “grade inflation” for defection biases the game in favor of cheating. Worse yet, in iterative games the players are forced to continue playing; they cannot exclude or ostracize a defector. They can only retaliate by themselves defecting and hoping thereby to penalize the other player. […] This reasoning is reflected in a new Prisoner's Dilemma model developed by Nowak and Sigmund (1993b) called “Pavlov”, which the authors suggest can outperform Tit-For-Tat. They call their strategy “win-stay, lose-shift,” and the significance of this innovation is that, in contrast with an iterated game in which the players must continue playing regardless of the outcome, in Pavlov they have the choice of leaving the game if they don’t like the results. In other words, a player may also have the power to exercise some control over the behavior of a defector by denying to that player future access to the game and its potential benefits. Punishments as well as rewards may be utilized as a means of keeping the game honest and, more important, as a means of restricting the game over time to mutual co-operators.” Not surprisingly, evolution employed a variety of these strategies, e.g. punishment, reward, and partner fidelity/choice to ensure cooperative behavior.

A multitude of interactions can stabilize cooperative behavior: direct and indirect reciprocity, sanctions, partner choice and fidelity (Gurven, 2004; Weyl et al., 2010; Archetti et al., 2011; Nowak & Highfield, 2011; Frederickson, 2013). Several natural mechanisms of promoting/enforcing cooperation have been explored such as sanctions and policing (Axelrod & Hamilton, 1981; Axelrod, 1984; Boyd & Richerson, 1992; Clutton-Brock & Parker, 1995; Denison, 2000; Sigmund et al., 2001, 2010; Yu, 2001; West et al., 2002b, 2002c; Kiers et al., 2003, 2006, 2011a; Gächter et al., 2008; Rockenbach & Milinski, 2009; Boyd et al., 2010; Helbing et al., 2010; Manhes & Velicer, 2011; Hilbe & Traulsen, 2012; Jandér et al., 2012; Raihani et al., 2012; Shutters, 2012). The evolution of individually costly punishment has been difficult to explain. Several studies in isolated groups have found clear evidence that while costly punishment increases the level of cooperation, the net effect in terms of material pay-off is often negative, decreasing the success of groups and individuals (Sefton et al., 2007; Dreber et al., 2008; Egas & Riedl, 2008; Wu et al., 2009). However, there is ample evidence that intergroup competition can be a powerful force for promoting within-group cooperation (Bornstein et al., 1990; Erev et al., 1993; Gunnthorsdottir & Rapoport, 2006; Tan & Bolle, 2007; Puurtinen & Mappes, 2009; Gneezy & Fessler, 2011; Sääksvuori et al., 2011; Voors et al., 2012; Böhm & Rockenbach, 2013; Rand & Nowak, 2013). Generally, group-selected systems can be expected to evolve policing against cheats (Frank, 1995). Group competition enhances the effectiveness of punishment so that when groups are in direct competition, individuals belonging to a group with punishment opportunity prevail over individuals in a group without this opportunity (Sääksvuori et al., 2011). Punishment can function as a form of reproductive leveling that is likely to change the selective environment so that it becomes more favorable to the evolution of behaviors that increase the success of the group relative to other groups (Bowles, 2006). In addition to competitive superiority in between-group competition, punishment reduces within-group variation in success, creating circumstances that are highly favorable for the evolution of accompanying group-functional behaviors (Sääksvuori et al., 2011). Experience with real-world intergroup conflict supports the results of mathematical models (Gneezy & Fessler, 2011; Voors et al., 2012).

Direct reciprocity (Trivers, 1971; Bowles & Gintis, 2004; Kiers et al., 2011b; Fellbaum et al., 2012), indirect reciprocity that works through reputation (Alexander, 1987; Boyd & Richerson, 1989; Nowak & Sigmund, 1998; Leimar & Hammerstein, 2001; Milinski et al., 2002; Fehr & Fischbacher, 2003; Ohtsuki & Iwasa, 2004, 2006; Panchanathan & Boyd, 2004; Suzuki & Akiyama, 2005; Bshary & Grutter, 2006), pleiotropy (Foster et al., 2004; Banin et al., 2005; Xavier & Foster, 2007; Harrison & Buckling, 2009; Dandekar et al., 2012), voluntary participation (Hauert et al., 2002; 2007; Szabó & Hauert, 2002; Semmann et al., 2003; Szabó & Vukov, 2004), or spatial extensions (Nowak & May, 1992) stabilize cooperative behavior. Tit-for-tat-like strategies can sustain mutualistic cooperation in the presence of cheaters (Frank, 1994; Doebeli & Knowlton, 1998; Bergstrom & Lachmann, 2003; Yamamura et al., 2004). Theoretical and empirical studies of indirect reciprocity show that people who are more helpful are more likely to receive help (Wedekind & Milinski, 2000; Leimar & Hammerstein, 2001; Milinski et al., 2002; Fishman, 2003; Hauser et al., 2003; Brandt & Sigmund, 2004; Gurven, 2004; Ohtsuki & Iwasa, 2004; Panchanathan & Boyd, 2004; Nowak & Sigmund, 2005; Nowak, 2006a). In his extensive work about the sharing of blood among vampire bats (their exclusive food-source) (Wilkinson 1984, 1988, 1990), Wilkinson concluded: “Reciprocity is likely to be more beneficial than kin selection – provided that cheaters can be detected and excluded from the system” (1990, p. 82). On the other hand, individuals may prefer close kin to distant kin and nonkin as partners for reciprocity. Relatives can make ideal candidates for reciprocal exchanges due to factors such as familiarity, trust, proximity, a high probability of future interaction, or an expectation that relatives will cooperate (Allen-Arave et al., 2008). When choosing among potential reciprocity partners, individuals should generally prefer partners who will provide the highest expected return benefit. Familiarity and emotional bonds fostered over time may make close kin easier to “read” and trust than distant kin and nonkin (Allen-Arave et al., 2008). Familiarity, emotional bonds, trust, and proximity can promote an expectation among kin that a relative will cooperate, and experimental research has shown that expectations of cooperation promote and stabilize “altruistic” behavior (Dawes, 1980; Messick & Brewer, 1983; Allen-Arave et al., 2008). Would-be transgressors likely experience more guilt from cheating victims with whom they have emotional ties (Frank, 1988).

However, relatively few examples of cooperation and mutualism in non-human species seem to fit reciprocity concepts (Hauser et al., 2003; Stevens & Gilby, 2004; Brosnan et al., 2010b; Connor, 2010; Leimar & Hammerstein, 2010; Melis & Semmann, 2010). It has even been suggested that the specific mechanism of reciprocity is unlikely to be of general importance outside of humans, because the conditions required can be extremely restrictive (Connor, 1995; Dugatkin, 1997; Clutton-Brock, 2002; Hammerstein, 2003; Stevens & Hauser, 2004; Stevens et al., 2005b). Trivers (2004) argued that reciprocity in all its aspects is very difficult to demonstrate, especially in nature. For example, it is not sufficient to demonstrate a positive correlation in beneficent acts across pairs of unrelated individuals (which is sometimes achieved). “Contingency” has also to be shown: that withholding a benefit by one results in a similar action by the other (which is rarely established).

A mechanism that has not been considered sufficiently in all models, at least to my knowledge (but see Queller, 1994a), is that, in a world of limited resources, the fitness gain associated with cooperative behavior results in a relative fitness loss due to competitive/solitary behavior. Resource availability follows a conservation law, implying that the gains or losses in resources to one individual are balanced by the losses or gains to others (Schluter, 1996, 2010). For instance, the huge success of eusocial insects should come at the expense of solitary organisms. Within the nervous system, cooperative synapses also have a competitive side: when some synapses grow stronger and prosper, others, which left to themselves would also have strengthened, instead weaken. Cooperation and competition between postsynaptic cells during correlation-based development determine the arrangement or maps of receptive field properties across the postsynaptic structure (Miller, 1996; Bhaumik & Mathur, 2003).

Many researchers have found it useful to separate reciprocity into two components; repeated interactions of partners, partner fidelity, and the ability of interactants to alter their response based on the other’s behavior–partner choice or sanctions (Noë, 1990; Bull & Rice, 1991; Nowak & May, 1992; Noë & Hammerstein, 1994; Doebeli & Knowlton, 1998; Simms & Taylor, 2002; West et al., 2002a; Sachs et al., 2004). Individuals often discriminate among partners according to the quantity of rewards they provide and associate differentially with higher reward producers (Bull & Rice, 1991; Christensen et al., 1991; Mitchell, 1994; Anstett et al., 1998; Ferriere et al., 2002). Plants and mycorrhizal fungi appear to be able to distinguish and selectively reward better symbionts (e.g., Hammer et al., 2011; Kiers et al., 2011b; Lekberg et al., 2010), although this ability may differ among plant species, depend on the spatial distribution of co-occurring fungi (Bever et al., 2009), and be moderated by bet-hedging (Lekberg & Koide, 2014). Hosts respond differentially to strains varying in their mutualistic benefit, reducing the fitness of less beneficial strains relative to more beneficial strains. Such ‘sanctions’ have been reported in soybean (Kiers et al., 2003), lupine (Simms et al., 2006), and in alfalfa and pea (Oono et al., 2011). Partner choice and partner fidelity are mechanisms for dealing with cheaters, and can theoretically allow mutualisms to persist despite cheaters (Simms & Taylor, 2002; Sachs & Simms, 2006). Partner fidelity involves long-term or repeated interactions among partners that promote correlated fitness interests between them (Trivers, 1971; Axelrod & Hamilton, 1981; Bull & Rice, 1991; Frank, 1994; Doebeli & Knowlton, 1998; Sachs et al., 2004; Foster & Wenseleers, 2006) such as microbial mutualists that are vertically transmitted in a host lineage (Axelrod & Hamilton, 1981; Bull & Rice, 1991; Frank, 1994; Herre et al., 1999; Sachs et al., 2004). By far the most common tool to enforce cooperation is partner choice. Partner choice or sanctions occur when an interacting party is able to alter its response based on the behavior of the other (Bull & Rice, 1991; Denison, 2000; Simms & Taylor, 2002; Kiers et al., 2003; Sachs et al., 2004; Foster & Wenseleers, 2006; Simms et al., 2006). For instance, in the mutualism between client fish and cleaner fish in which cleaner fish obtain food from removing client parasites, client fish recognize and avoid cheating cleaners that also bite healthy tissues (Bshary, 2002). Models in public goods games show that both external rewards and punishment enhance cooperation in a society (Hilbe & Sigmund, 2010; Szolnoki & Perc, 2010). It has been proposed that a good way of increasing cooperation would be to use a reward first (the carrot), and a punishment later (the stick) (Hilbe & Sigmund, 2010).

Punishing non-cooperative partners or rewarding cooperative ones can help maintain cooperation between species (Clutton-Brock & Parker, 1995) if the behavior of the partners can be observed. In some cases discrimination may not be possible and one may only be able to react to the collective action of its partners, either because there is no way to detect the behavior (or its effects) of individual partners or because the differences between cooperators and non-cooperators are too small to be detected. Archetti and Scheuring (2013) showed that if the benefit of the public goods traded by the two species is a nonlinear saturating function of the individual contributions, cooperators and defectors can coexist in sizeable groups, in the absence of assortment and of discrimination mechanisms, and mutualism can be maintained in well-mixed population without punishing the free riders. Momeni et al. (2013) conducted a study with two genetically engineered species of yeast that mutually cooperate, each providing a metabolite that is essential to the other, but are not able to recognize each other (this means that these populations cannot rely on partner choice to combat cheaters), and a third species of yeast that cheated by consuming one of the metabolites without releasing any metabolite of its own. Momeni et al. found that as long as there was space for the yeast cells to grow into, the two species that cooperated self-organized into mixed clusters, with the cheating species being excluded from these clusters. The self-organization was driven by a positive feedback loop involving the two species that cooperated, with each species helping to increase the fitness of the other. The results demonstrate that it is possible for two genetically unrelated populations to cooperate and combat cheaters without the use of partner choice (Momeni et al., 2013).

There is ample evidence that partner choice is a strong coevolutionary force in cooperative and mutualistic systems (Noë & Hammerstein, 1994; Noë, 2001; Bshary & Grutter, 2002; Simms & Taylor, 2002; Simms et al., 2006; Johnstone & Bshary, 2008; Leimar & Hammerstein, 2010). Unsatisfactory relationships can be abandoned and replaced by ones with greater benefits. For example the choice of a more desirable aphid species draws ant protection away from the less desirable aphid species which, as a result, experience a higher level of predation and population decline (Fischer et al., 2001). With each individual shopping for the best partners and selling its own services, the framework becomes one of supply and demand, as formalized in Noë and Hammerstein’s (1994) Biological Market theory. This theory applies whenever trading partners can choose with whom to deal. Market mechanisms are an effective way of sidelining profiteers (de Waal & Suchak, 2010). Partner choice has long been considered sufficient for the maintenance of cooperation both within and among species (Noë, 1990; Bull & Rice, 1991; Noë & Hammerstein, 1994; Dugatkin & Sih, 1995; Noë et al., 2001; Bshary & Noë , 2003; Sachs et al., 2004; Sachs, 2006). Models have predicted that partner choice mechanisms can act alone to support intra-specific cooperation (Eshel & Cavalli-Sforza, 1982; Noë, 1990; Noë & Hammerstein, 1994) and empirical examples appear to fulfil these predictions. Examples include nuptial gifts in scorpionflies (Thornhill, 1976, 1984) and bushcrickets (Gwynn, 1988), as well as coalitions in baboons (Noë, 1990) and common eiders (Ost et al., 2003, 2005). Partner choice in a market of potential symbionts can constrain cheating, but only if the following conditions apply: (i) A range of partners is available, (ii) there is a mechanism for effecting choice, and (iii) the cost of evaluating partners is less than the benefit derived from choosing a good partner (Simms & Taylor, 2002).

16. Stochasticity, complex systems and self-organization


A striking difference between linear and nonlinear laws is whether the property of superposition holds or breaks down. In a linear system the ultimate effect of the combined action of two different causes is merely the superposition of the effects of each cause taken individually. But in a nonlinear system adding two elementary actions to one another can induce dramatic new effects reflecting the onset of cooperativity between the constituent elements. This can give rise to unexpected structures and events whose properties can be quite different from those of underlying elementary laws, in the form of abrupt transitions, a multiplicity of states, pattern formation, or an irregularly markedly unpredictable evolution of space and time referred to as deterministic chaos. Nonlinear science, is therefore, the science of evolution and complexity.
Nicolis 1995, pp 1–2

Summary
Complex systems are characterized by interacting units that display global properties not present at the lower level. Complex adaptive systems also require stochastic factors, e.g. noise and fluctuations. For self-organization to arise, a system needs to exhibit two properties: it must be both dissipative and nonlinear. Generally, the polarity of randomness and law characterizes the self-creating natural world. Only with an intermediate level of stochastic variation, somewhere between determined rigidity and literal chaos local interactions give rise to complexity.Cooperative behavior on various levels is an emergent phenomenon of self-organized systems. Both the formation of swarms and division of labor in a colony are self-organizedbehaviors at the group level as an emergent consequence of individuals’ interactions.

To begin with, the term complex is a relative one. Individual organisms may use relatively simple behavioral rules to generate structures and patterns at the collective level that are relatively more complex than the components and processes from which they emerge. Systems are complex not because they involve many behavioral rules and large numbers of different components but because of the nature of the system’s global response. Complexity and complex systems, on the other hand, generally refer to a system of interacting units that displays global properties not present at the lower level. These systems may show diverse responses that are often sensitively dependent on both the initial state of the system and nonlinear interactions among its components. Since these nonlinear interactions involve amplification or cooperativity, complex behaviors may emerge even though the system components may be similar and follow simple rules (Camazine et al., 2001). Life reflects all the key characteristics of complex systems: living organisms have a large number of interdependent constituents that often behave chaotically, span across several organizational scales, show collectivity and emergence, and maintain a balance between cooperation and competition (Baranger, 2000; Baffy & Loscalzo, 2014). Complex biological systems manifest a large variety of emergent phenomena among which prominent roles belong to self-organization and swarm intelligence (Proulx, 2007).Complexity, in Ashby’s sense, is essentially conceived as a system’s potential to assume a large number of states, and we also have a measure for it: variety, the number of states a system can assume (Schwaninger, 2004).

Complex systems, by their very nature, do not yield to a traditional reductionist approach (Hartwell et al., 1999; Kitano, 2002; Vendruscolo et al., 2003). Such systems are made up of a large number of distinct parts which interact and are often organized hierarchically (Weng et al., 1999; Csete & Doyle, 2002; Milo et al., 2002; Oltvai & Barabasi, 2002). The complexity of a system scales with the number of its elements, the number of interactions between them, the complexities of the elements, and the complexities of the interactions (Gershenson, 2002, 2005). This can be confirmed mathematically in certain systems. As a general example, random Boolean networks (Kauffman 1969; 1993; Gershenson, 2004) show clearly that the complexity of the network increases with the number of elements and the number of interactions. The result is a dazzling multiplicity of possible patterns of behavior that makes it extremely difficulty to predict and control the course of events, as is well known in weather or in financial market forecasting (Goldenfeld & Kadanoff, 1999; Vendruscolo et al., 2003).

For self-organization to arise, a system needs to exhibit two properties: it must be both dissipative and nonlinear. Generally, the polarity of randomness and law characterizes the self-creating natural world (Carr, 2004). Heinz von Foerster (1960) formulated the principle of “order from noise”. He noted that, paradoxically, the larger the random perturbations (“noise”) that affect a system, the more quickly it will self-organize (produce “order”). Another reason for this intrinsic robustness is that self-organization thrives on randomness, fluctuations or “noise”. An important factor to consider when understanding the collective behaviors of animal groups (and self-organized pattern-forming processes in general) is the influence of stochastic (random) events. Animal behavior is inherently probabilistic, and stochastic properties of animal movement are likely to strongly influence the structure of many groups. It is becoming increasingly evident that self-organized patterns often arise because of the amplification of random fluctuation (Nicolis & Prigogine, 1977; Seeley, 1995). By developing stochastic computer models of animal groups the essential statistical mechanics of the system may be captured (Grünbaum, 1998; Couzin & Krause, 2003).

Self-organization is a set of dynamical mechanisms whereby structures appear at the global level of a system from interactions among its lower-level components, without being explicitly coded at the individual level (Garnier et al., 2007). “It relies on four basic ingredients: (1) The first component is a positive feedback that results from the execution of simple behavioral “rules of thumb” that promote the creation of structures. For instance, trail recruitment to a food source is a kind of positive feedback which creates the conditions for the emergence of a trail network at the global level.
(2) Then we have a negative feedback that counterbalances positive feedback and that leads to the stabilization of the collective pattern. In the example of ant foraging, negative feedback may have several origins. It may result from the limited number of available foragers, the food source exhaustion, and the evaporation of pheromone or a competition between paths to attract foragers.
(3) Self-organization also relies on the amplification of fluctuations by positive feedbacks. Social insects are well known to perform actions that can be described as stochastic. Such random fluctuations are the seeds from which structures nucleate and grow. Moreover, randomness is often crucial, because it enables the colony to discover new solutions. For instance, lost foragers can find new, unexploited food sources, and then recruit nest mates to these food sources.
(4) Finally, self-organization requires multiple direct or stigmergic interactions among individuals to produce apparently deterministic outcomes and the appearance of large and enduring structures.
In addition to the previously detailed ingredients, self-organization is also characterized by a few key properties:
(1) Self-organized systems are dynamic. As stated before, the production of structures as well as their persistence requires permanent interactions between the members of the colony and with their environment. These interactions promote the positive feedbacks that create the collective structures and act for their subsistence against negative feedbacks that tend to eliminate them.
(2) Self-organized systems exhibit emergent properties. They display properties that are more complex than the simple contribution of each agent. These properties arise from the nonlinear combination of the interactions between the members of the colony.
(3) Together with the emergent properties, non linear interactions lead self-organized systems to bifurcations. A bifurcation is the appearance of new stable solutions when some of the system’s parameters change. This corresponds to a qualitative change in the collective behavior.
(4) Last, self-organized systems can be multi-stable. Multi-stability means that, for a given set of parameters, the system can reach different stable states depending on the initial conditions and on the random fluctuations” (Garnier et al., 2007).

According to the self-organization approach, by studying interactions on a lower level, the emergence of a macrostructure on a higher level is perceived and, therefore, better understood (Hogeweg, 1988). Patterns of interactions at a group level arise from local interactions between individuals and their environment. By interacting, individuals change each other and, therein, their social environment. In turn, the developing social structure feeds back to the individuals and shapes their interactions, etc. Individual and group behaviors become a product of social dynamics (Moore et al., 1997; Wolf et al., 1998; Fewell & Page, 1999; Camazine et al., 2001; Clark & Fewell, 2014). Experiments with robots have demonstrated that cooperative behavior which looks complex and sophisticated to an observer can be nothing more than the result of straightforward, simple interactions between agents and their local environment (Maris & te Boekhorst, 1996). Consequently, this approach attributes the complexity of social interaction patterns to interactions between entities rather than their internal complexity (Hemelrijk, 1999, 2002). This reflects a shift of focus from objects to relationships (Cohen & Stewart, 1994), whereby relationships are often considered to be self-reinforcing. Moreover, natural selection will often operate, not on single traits, but on self-organized patterns (Boerlijst & Hogeweg, 1991; Hemelrijk, 1999). To gauge the complexity-generating effects of interindividual interactions, Hemelrijk (1996, 1997) investigated the ‘social organisation’ of a group of simple, artificial agents by means of an individual-oriented model. In a virtual world, cooperation emerged as a self-organized phenomenon. These effects arose without considering any particular costs or benefits associated with acts of cooperation and defection but were due to the intertwined effects of displaying dominance and social cohesion (Hemelrijk, 1997). With self-organization, complexity and emergence, the realization increasingly shared by many scientists is that behavior at one ‘level’ is then more than the ‘sum’ of the behavior of individual entities (Skår & Coveney, 2003).

Cooperative behavior on various levels is associated with self-organization phenomena (Deneubourg et al., 2002; Hemelrijk, 2002, 2005; Miramontes & DeSouza, 2014; Szolnoki et al., 2014). Models of self-organized systems are able to explain the emergent cooperative behavior of microorganisms (Ben-Jacob et al., 1994, 2000, 2004) and groups of animals such as insects (Dussutour et al., 2004; Jeanson et al., 2005; Detrain & Deneubourg, 2006), fishes (Couzin et al., 2005; Ward et al., 2008), birds (Hildenbrandt et al., 2010; Hemelrijk & Hildenbrandt, 2012), primates (Puga-Gonzalez et al., 2009; Hemelrijk & Puga-Gonzalez, 2012) and humans (Moussaid et al., 2010). D. discoideum aggregation following metabolic stress (see chapter 5.2) displays emergence of cooperative behavior as a result of self-organization (Marée & Hogeweg, 2001). To this date, the wavelike patterns observed during D. discoideum aggregation represent one of the most beautiful and best understood examples of spatiotemporal self-organization at the cellular level (Goldbeter, 2006; Gregor et al., 2010). The wavelike nature of aggregation (Gerisch, 1968; Alcantara & Monk, 1974; Gerisch et al., 1975) results from the existence of a oscillatory cellular rhythm in the production of cAMP (Gregor et al., 2010; Prindle & Hasty, 2010).

16.1 Stochasticity of swarms

16.1.1 Swarm behavior

Groups of many kinds of fishes show a characteristic social aggregation, school (Shaw, 1978). The problem of why fish groups make schooling behaviors has often been studied based on the evolutional assumption that schools must increase the survivorship or reproductive success of individuals in the school (Breder, 1967; Cashing & Harden-Jones, 1968; Magurran, 1990; Pitcher & Parrish, 1993). Selection forces related to the location and acquisition of food and predator–prey interactions may have played very important roles in shaping the evolution of aggregating behavior (Zheng et al., 2005).

Schools make frequently sharp turns or accelerations in which every individual within the school appears to react simultaneously. However, it has been revealed (Partridge, 1981, 1982) that individuals within a school do not all respond instantaneously to course change by other individuals but there is a well-defined latency or lag. The properties of a shoal are the product of local rules between neighboring fish that, via self-organization, generate behavior at the group level as an emergent consequence of individuals’ interactions (Parrish et al., 2002; Vabø & Skaret, 2008; Rieucau et al., 2014). Intriguingly, the role of stochasticity extends to features like facilitation of coherence in collective swarm motion (Yates et al., 2009). An individual’s response to a loss of alignment in the group is increased randomness of its motion, until an aligned state is again achieved. The phenomena of using randomness to keep the group ordered appears counterintuitive but has been reported at the level of an individual (Douglass et al., 1993) and at the collective level in an ecological system (Yates et al., 2009). Furthermore, the lack of sources of environmental noise indicates that the noise-induced alignment seems to be an intrinsic characteristic of collective coherent motion. The frequency of stochastic asynchronous updating of individual positions and orientations increases with perceived threat and leads to more synchronized group movement, with speed and nearest-neighbor distributions becoming more uniform (Bode et al., 2010a, b). In fact, experiments with three-spined sticklebacks (Gasterosteus aculeatus) that were exposed to different threat levels suggest that the behavior of fish (at different states of agitation) can be explained by a single parameter in the model, the updating frequency.

A widely appreciated benefit of group membership is the reduction in cost of locomotion for individuals that trail behind others, taking advantage of the vortices, e.g. birds flying in a V formation or fish schooling (Weihs, 1973; Cutts & Speakman, 1994; Weimerskirch et al., 2001; Marras et al., 2014; Portugal et al., 2014), or zones of low pressure (e.g. bicyclists drafting in a peloton or vehicles on a motorway) created by their leading group mates (McCole et al., 1990; Dominy, 1992; Fish, 1999). Such energetic savings can be significant enough to be considered one of the main benefits of group membership for schooling fish, flocking birds and cycling humans (Fish, 1999; Krause & Ruxton, 2002; Liao et al., 2003).

16.1.2 Swarm intelligence

Stochastic environments confront organisms with unfamiliar situations. As one solution to the problem of uncertainty, organisms can attempt to reduce the uncertainty associated with key features of their environments by collecting and storing information. By sampling each of its options regularly, animals gain from being able to exploit them when they are productive and avoid them otherwise. In this way, collecting information can be thought of as a solution to the uncertainty problem that maximizes potential opportunities (Stephens, 1989; Mangel, 1990; Dall & Johnstone, 2002). Unpredictable or variable environments favor the evolution of cognition and learning (Bergman & Feldman, 1995; Richerson & Boyd, 2000; Godfrey-Smith, 2002; Mery & Kawecki, 2002; Brown et al., 2003; Kerr & Feldman, 2003; Borenstein et al., 2008; Kotrschal & Taborsky, 2010; Clarin et al., 2013; Tebbich & Teschke, 2014) and cognition/learning is thought to enable organisms to deal with environmental heterogeneity (Godfrey-Smith, 2002). Reviews of the factors contributing to the emergence of social learning emphasize the role played by a spatially and temporally changing environment (Laland et al., 2000; Richerson & Boyd, 2000; Alvard, 2003; Henrich & McElreath, 2003; Aoki et al., 2005). Social learning has been observed in a wide range of species in diverse taxa including mammals (Perry & Manson, 2003; Galef & Laland, 2005; Perreault et al., 2012), birds (Lefebvre, 2000; Benskin et al., 2002), fish (Brown & Laland, 2003), and invertebrates (Leadbeater & Chittka, 2007a, b). Social interactions form one part of a complex environment and intelligence and learning may have evolved to navigate the social world (Humphrey, 1976; Byrne, 1996; Byrne & Bates, 2007; Borenstein et al., 2008; Hamblin & Giraldeau, 2009; Arbilly et al., 2010; Dubois et al., 2010; Katsnelson et al., 2012; Smead, 2014).

The study of swarm intelligence is deeply embedded in the biological study of self-organized behaviors in social insects (Garnier et al., 2007). Social animals choose between alternative actions (Conradt & Roper, 2005; King & Cowlishaw, 2009) that is vital if a group is to remain a cohesive unit and accrue the many advantages of group living (Krause & Ruxton, 2002). Couzin and Krause (2003) wrote: “Information transfer among individuals is likely to influence their response to stimuli, such as the positions of resources, or favorable regions within a heterogeneous environment. In aquatic habitats, for example, resources such as phytoplankton, the temperature or salinity of the water, and concentrations of dissolved gases are all known to vary in a nonuniform way, and over both small and large length scales. Individuals are therefore expected to modify their positions with respect to these properties so as to maximize resource intake and minimize physiological stress. However, this is a nontrivial task: unpredictability and local fluctuations make finding and moving up or down such environmental gradients (taxis) difficult when an individual has only local knowledge on which to base its motion. Grünbaum (1998) used computer simulation to investigate the theoretical consequences of grouping to such taxis behavior. He assumed individuals use a simple form of taxis, known as klinotaxis, whereby a moving individual modifies its probability of making a turn as a function of whether conditions are perceived to improve or deteriorate over a given time interval. Such behavior is known to facilitate taxis in even simple organisms, such as bacteria (Keller & Segel, 1971; Alt, 1980; Tranquillo, 1990). Although they are not directly detecting the gradient, individuals performing such taxes will, on average, spend more time moving in favorable directions than in unfavorable ones. By simulating groups of individuals performing this behavior under conditions in which they do not interact with one another (asocial taxis) and do interact by balancing the tendency for taxis with a simple schooling behavior (social taxis), Grünbaum (1998) demonstrated that such social interactions improve the motion of individuals up a gradient. The alignment of individuals, and thus transfer of information, when schooling, allows averaging of individual errors in gradient detection, and therefore results in reduced deviations in motion from the desired direction of travel. This information sharing within schools of fish has been likened to a ‘sensory array’ (Kils, 1986), which allows information to be gathered over a wider spatial range than would be possible for a solitary or noninteracting individual, and dampens the influence of small-scale fluctuations in the environment. The model also predicts that the benefits of such information sharing are dependent on group size. As group size is increased the efficiency of taxis shows an asymptotic increase: initially it increases steeply, but then the rate of increase reduces over time, leading to a plateau where further increases in group size have little effect on taxis accuracy.”

A large amount of work suggests that a social-insect colony is a decentralized system comprised of cooperative, autonomous units that are distributed in the environment, exhibit simple probabilistic stimulus response behavior, and only have access to local information (Deneubourg & Goss, 1989; Bonabeau et al. 1997; Theraulaz et al. 1998, 2003; Camazine et al. 2001). Without centralized control, workers are able to work together and collectively tackle tasks far beyond the abilities of any one individual. The resulting patterns produced by a colony are not explicitly coded at the individual level, but rather they emerge from myriads of simple nonlinear interactions between individuals or between individuals and their environment (Theraulaz et al., 2003). The decisive factor favoring social learning in insects may not be coloniality, as suggested by the comparison between colonial and solitary bees (Dukas, 1987; Dukas & Real, 1991), but more simply the opportunity to interact with organisms sharing similar ecological needs and constraints (Coolen et al., 2005; Worden & Papaj, 2005). Social interaction can provide a solution to a cognitive problem via two potential mechanisms: (i) individuals can aggregate information, thus augmenting their ‘collective cognition’, or (ii) interaction with conspecifics can allow individuals to follow specific ‘leaders’, those experts with information particularly relevant to the decision at hand. For repeated decisions–where individuals are able to consider the success of previous decision outcomes–the collective’s aggregated information is almost always superior (Katsikopoulos & King, 2010). Social/swarm intelligence can facilitate solving cognitive problems that go beyond the capacity of single animals (Kennedy et al., 2001; Garnier et al., 2007; Hinchey et al., 2007; Moussaid et al., 2009; Katsikopoulos & King, 2010; Krause et al., 2010). In most cases, the collective decisions and patterns arise as a result of competition between different sources of information which can be amplified in different ways (De Schutter et al., 2001). As shown in ants and honeybees, quorum sensing is used to collate individual assessments and form them into a collective decision (Pratt et al., 2002; Seeley & Visscher, 2004). Individual animals show irrational changes in preference (Latty & Beekman, 2011). The collective decisions of insect societies are immune to irrationality due to the colonies’ decentralized decision mechanism that may prevent systematic errors that would otherwise arise from the cognitive limitations of individuals (Edwards & Pratt, 2009). Research on collective behavior and group-decision-making in animals has shown that individuals in groups can outperform solitary individuals (Bonabeau et al., 1999; Camazine et al., 2001; Couzin, 2009; Krause et al., 2010). Group-living has important consequences for the performance of individuals in anti-predator contexts (Krause & Ruxton, 2002; Beauchamp, 2013). In predator detection, groups achieve higher true positives (Pulliam, 1973; Powell, 1974; Siegfried & Underhill, 1975; Kenward, 1978; Lazarus, 1979; van Schaik et al., 1983b; Cresswell, 1994; Wolf et al., 2013) and diminish the negative consequences of false positives by using a behavioral rule that does not respond to single but only to multiple other individuals (Lima, 1994; Roberts, 1997; Proctor et al., 2001; Beauchamp & Ruxton, 2007; Beauchamp, 2010; Wolf et al., 2013).

Brains, e.g. of mammals, can be considered as societies of neurons and similar adaptive network processes may be involved during cognitive processes generated by central nervous systems and social insects (Schall, 1999, 2001; De Schutter et al., 2001; Shadlen & Newsome, 2001; Fewell, 2003; Heekeren et al., 2004; Couzin, 2009). In the case of social insect societies, the insects must organize their workforce efficiently in order to survive. This organization involves making collective decisions that optimize the colony’s fitness (Sumpter & Beekman, 2003). Thus, both ants and honeybees are capable of choosing collectively the best of several possible new nest sites during migration or swarming (Seeley, 1995; Sumpter & Beekman, 2003) or turn individual and collectively gathered information into foraging success (Wilson, 1971; Hölldobler, 1976; Aron et al., 1989; Traniello, 1989; Crist & MacMahon, 1991; Mull & MacMahon, 1997; Flanagan et al., 2012; Letendre & Moses, 2013). Even bacterial colonies, swarms and films exhibit an unanticipated complexity of behaviors that can undoubtedly be characterized as based on biogenic cognition (Ben-Jacob et al., 2004, 2006; Waters & Bassler, 2005; van Duijn et al., 2006; Ben-Jacob, 2008, 2009; Ng & Bassler, 2009).

During a century of research on social learning, the focus has been largely on interactions between conspecifics. Examples of studies on learning from members of other species are rare (e.g. Seyfarth & Cheney, 1990; Coolen et al., 2003; Rainey et al., 2004; Seiler et al., 2013), yet, there is no a priori reason to treat those cases differently. In fact, where resources are shared or where generalist predators lurk, picking up information from heterospecifics may be just as valuable as from members of the same species (Leadbeater & Chittka, 2007b).

16.2 Division of labor due to self-organization

In nature, division of labor, in the broad sense of the expression, is widespread. The main evolutionary transitions, such as those from prokaryotes to eukaryotes and from unicellular to multicellular organisms, were accompanied by division of labor (Szathmáry & Maynard Smith, 1995). Within social groups, division of labor is also common. In species with biparental care, males and females frequently have different roles in raising the offspring. In hornbills, for example, breeding females seal themselves in the nest, and males must feed them during this time (Kemp & Woodcock, 1995). Group hunting (e.g., Gazda et al., 2005), sentinel behavior in group foragers such as meerkats (Manser, 1999), and specialization in either predator defense or provisioning in noisy miners (Arnold et al., 2005) are other examples of division of labor. Division of labor occurs when individual members of a group specialize by performing particular tasks toward some common goal. Division of labor is an important aspect of cooperative behavior (Robinson, 1992; Crespi, 2001; Wahl, 2002a, 200b; Arnold et al., 2005) that increases the entire group’s productivity (Michod & Roze, 2001; Lnger et al., 2004; Pruitt & Riechert, 2011). Among vertebrates, division of labor is not strictly caste based, but rather varies by age, sex and size (such behavioral differences between categories of individuals are known as ‘polyethisms’; e.g. naked mole rats Heterocephalus glaber; Lacey & Sherman, 1991, 1997; meerkats; Clutton-Brock et al., 2003; white-winged choughs Corcorax melanorhamphos; Heinsohn & Cockburn, 1994; and cichlids Neolamprologus pulcher; Bruintjes & Taborsky, 2008).

The ecological success of social insects can be mainly attributed to their division of labor (Oster & Wilson, 1978; Wilson, 1985, 1987; Hölldobler & Wilson, 1990; Myerscough & Oldroyd, 2004). Highly specialized individuals are thought to contribute to colony fitness by working more efficiently than less specialized workers (Oster & Wilson, 1978; Heinrich, 1979; Calabi & Traniello, 1989; Beshers & Fewell, 2001; Chittka & Muller, 2009; Pruitt & Riechert, 2011). The fittest populations are those that divide tasks fairly and associate in large, highly specialized groups. Generalists have a distinct advantage in small groups, but the presence of generalists lowers group fitness. A model (Wahl, 2002) suggested a tendency for populations to evolve increasingly unfair divisions of labor. This implies that an evolutionary ratchet favors disparity between the workload of specialist populations (Wahl, 2002). Temnothorax longispinosus ant colonies with more intracolonial behavioral variation as an important component of division of labor in brood care, aggression and exploration of novel objects were more productive under standardized conditions than colonies with less variation (Modlmeier & Foitzik, 2011; Modlmeier et al., 2012). Importantly, the concept of specialization is a statistical one, reflecting an individual’s tendency to perform particular tasks more often than others. The strength of this tendency may vary greatly, ranging from temporary behavioral differentiation to fixed morphological differentiation in insect species that form large societies (Robinson, 1992).

Changes in division of labor can be induced experimentally by altering colony size (Winston & Fergusson, 1985) or demography (Rösch, 1930) or by increasing the need for nest maintenance (Gordon, 1991), nest repair (O’Donnell & Jeanne, 1990), additional storage comb (Rösch, 1930; Kolmes, 1985; Fergusson & Winston, 1988), and defensive behavior (Gordon, 1991). Proximate analyses of division of labor generally are based on the concept of self-organization (Duarte et al., 2011). According to this view, division of labor emerges spontaneously at the origins of sociality, before becoming a target of natural selection, due to the interaction of individuals obeying simple behavioral rules (Bonabeau et al., 1997; Page & Mitchell, 1998). This concept has been supported by behavioral experiments showing that normally solitary harvester ant queens and halictine bees exhibit task specialization when forced to associate, i.e., paired individuals dedicated most of their time to different tasks (Page, 1997; Fewell & Page, 1999; Beshers & Fewell, 2001; Camazine et al., 2001; Helms Cahan & Fewell, 2004; Jeanson et al., 2005, 2008; Jeanson & Fewell, 2008; Holbrook et al., 2009; Helms Cahan & Gardner-Morse, 2013). The division of labor among cofounding queens results from dominance interactions in wasps (Pratte, 1989; Strassmann, 1989) and in small ant societies (Bourke, 1988b; Ito & Higashi, 1991). These aggressive interactions occur when the cofounding queens are both related (Bourke, 1988b; Pratte, 1989; Strassmann, 1989; Ito & Higashi, 1991) or unrelated (Kolmer & Heinze, 2000). In the ant Cerapachys biroi, where genetic variation within colonies is constrained by parthenogenesis, individual experience and learning is crucial for self-organized task specialization (d’Ettorre, 2007; Ravary et al., 2007). In addition, increasing evidence supports the role of genetic factors in the process of caste determination and division of labor in several species (Calderone & Page, 1988; Frumhoff & Baker, 1988; Page & Robinson, 1991; Dreller et al., 1995; O’Donnell, 1996; Julian & Fewell, 2004; Anderson et al., 2008; Crozier & Schlüns, 2008; Lo et al., 2009; Schwander et al., 2010; Libbrecht et al., 2011). For example, in species with multiply mated queens or multiple queens per colony, different patrilines and matrilines tend to differ in their tendencies to perform certain tasks, demonstrating a genetic component in response threshold (e.g. Robinson & Page, 1988; Page & Robinson, 1991; Page et al., 1998). Epigenetic DNA methylation is associated with differential gene expression in castes of the honeybee, Apis mellifera, and ants (Kucharski et al., 2008; Elango et al., 2009; Chittka & Chittka, 2010; Bonasio et al., 2012; Chittka et al., 2012; Foret et al., 2012; Smith et al., 2012; Shao et al., 2014). Worker development is associated with increased DNA methylation; larvae reared in vitro with RNAi knockdown of DNA methyl transferase 3 show greatly increased probabilities of developing into queens (Kucharski et al., 2008). DNA methylation might serve as a developmental switch to regulate the expression of many downstream genes involved in determining caste fate, including IIS-related genes (Smith et al., 2008). In addition, neural network processes have been proposed to account for the emergence of intracolony task allocation on the basis of variation in response thresholds (Lichocki et al., 2012).

It has been shown (Simpson, 2012) that at the transition from unicellularity to multicellularity, a reproductive division of labor will evolve first in a majority of cases, and that the total extent of functional differentiation will be larger if there is a reproductive division of labor. Accordingly, specialization of cells in reproductive (germ cells) and vegetative functions (soma) is an universal feature of unitary multicellular life. Conflicts between units arise when the selection pressures on some of the units favor one outcome, whereas those on other units favor another. The most basic conflict is between units of the same species or organism when selection pressure on one of the units favors the survival of its own lineage over survival of the lineage of the other unit. These conflicts with their Red Queen coevolutionary dynamics underlie the evolutionary dynamics of functional differentiation within and between organisms and their speciation.

17. Stochasticity, lotteries and insurance


Summary
Stochastic environments force organisms into risky lotteries (see chapter 5.4).On the other hand, insurance is the risk-sharing strategy of risk-averse agents that have to compete in lotteries.It is a common observation that people/foragers exhibit risk-aversion when making some choices while also exhibiting risk-preference in other cases. Swarm formation in response to predator pressure is aprototypic insurance against idiosyncratic risk. The defense-worthy fortress providing shelter from predation, a favorable microclimate, and sometimes food, is a life insurance. Social queuing processes, i.e. waiting peacefully in line to potentially acquire dominant status in the future, seem to occur in a wide variety of taxa and life histories. Queuing can be regarded as conservative bet-hedging strategy compared to the risky alternative, dispersal. In a range of queuing species, offspring that delay dispersal and wait on their natal territory either gain better quality territories or are more likely to breed than their siblings that disperse as juveniles. Assured fitness returns models argue that patterns of adult mortality combine with offspring dependence on parental care to select for group living. Assured fitness return represents a type of conservative bet-hedging strategy in a general sense, because a solitary female that survives the period of her brood’s dependence would have the highest reproductive success, but brooding females in multifemale colonies trade the possibility for maximum reproductive success for a greater mean reproductive success.

In nature, individuals have to choose from sets of risky alternatives. Stochastic environments force organisms into lotteries. The lotteries may involve either idiosyncratic risk or aggregate uncertainty or both (Robson, 1996). Idiosyncratic risk, respectively uncertainty, is risk or uncertainty to which only specific agents are exposed, in contrast to systematic or aggregate risk/uncertainty that is faced by all agents in the market. For example, the weather is a standard example of aggregate risk—a very harsh winter may kill all members of a population. In evolution, often risk is a combination of systemic component stemming primarily from the impact of unfavorable weather events and of an idiosyncratic component depending on individual characteristics and events. Cooper and Kaplan (1982) have demonstrated that when lotteries are aggregate, the optimal decision rule involves randomization. Via the law of large numbers, evolution generated a form of automatic biological insurance against idiosyncratic risk, whereas this insurance is inoperative in the same sense against aggregate uncertainty (Robson, 1996).

17.1 Swarm formation as insurance

Animals form more cohesive or larger groups, with consequent greater mutualistic benefits under greater predation risk (Seghers, 1974; Farr, 1975; Dunbar, 1988; Spieler, 2003; Krams et al., 2010). Empirical and theoretical studies suggest an antipredation-function of animals forming aggregations (Hamilton, 1971; Pulliam, 1973; McNamara & Houston, 1992; Krebs & Davies, 1993; Lima, 1995; Fréon & Misund, 1999; Coleman et al., 2004; Caro, 2005; Brierley & Cox, 2010; Ioannou et al., 2012). It has been shown that the perception of an increase in the risk of predation can induce cooperative behavior in some bird species (Krams et al., 2010) and other taxa with and without kinship (Lavalli & Herrnkind, 2009). Although most studies focused either on the prey or the predator side, the few studies that tackled both sides of the mechanisms simultaneously showed an optimization of the grouping strategies for both prey and predators (Major, 1978; Lett et al., 2004, 2014; Fryxell et al., 2007; Vaughn et al., 2010).

Attacks at high prey densities may decrease when prey has active or inducible defense behavior such as aggregation or swarming that is induced by predators as emergent behavior at high prey densities (Jeschke & Tollrian, 2005, 2007; Jeschke, 2006; Krams et al., 2010; Vucic-Pestic et al., 2010). Fish in large groups have a reduced per capita risk of predation as a result of several mechanisms, including earlier predator detection, numerical dilution of risk and predator confusion during attacks (e.g. Neill & Cullen, 1974; Magurran et al., 1985; Morgan & Godin, 1985; Godin, 1986; Pitcher & Parrish, 1993; Krause & Godin, 1995, Hoare et al., 2004). One main way that the school reduces a predator’s chance of making a successful kill is to confuse the predator as it makes its strike (Zheng et al., 2005). A predator facing a large number of fish quickly moving together has difficulty in singling out and tracking individual fish in a group. This confusion effect has been studied by many researchers (e.g. Major, 1978; Pitcher & Parrish, 1993). Neil and Cullen (1974) demonstrated by observing the hunting behavior of cephalopods and fish predators that for all species the increasing size of prey fish group decreased the success of the predators’ attacks per encounter with a prey. A much greater success rate per attack has been achieved if single preys are targeted (Zheng et al., 2005).

17.2 Fortress defense

Evolution of sociality in colonial invertebrates, social insects, and nonhuman mammals is likely to have often been driven by the adaptive nature of fortress defense. The fortress providing shelter from predation, a favorable microclimate, and sometimes food, is a life insurance (Foster & Northcott, 1994; Choe & Crespi, 1997; Queller & Strassmann, 1998; Pike & Foster, 2004; Wilson, 2008). Nest architectures obtained by simulations show that the complexity of the structures that are built by social insects is based on simple probabilistic stimulus-response behaviors but does not require sophisticated individual behavioral rules (Garnier et al., 2007). Property in the form of a private nest is an important preadaptation for eusociality (Alexander, 1974). One of the most familiar features of vertebrate or invertebrate social cooperation is a burrow, nest, hive or gall (e.g. ant, termite or bird nests, rodent burrows, beehives or aphid galls), usually made by the animals (Choe & Crespi, 1997). All of the clades known with primitively eusocial species surviving (in aculeate wasps, halictine and xylocopine bees, sponge-nesting shrimp, termopsid termites, colonial aphids and thrips, ambrosia beetles, and naked mole rats) have colonies that have built and occupied defensible nests (Wilson, 2008). The nest functions not only as a shelter for the queen and the brood, but may buffer individuals from temperature changes and enables the storage of food as an insurance against the vagaries of nature. In a sequence of events, the first stage and causative agent is the advantage of a defensible nest, especially one both expensive to make and within reach of adequate food (Nowak et al., 2010). Nowak et al. (2010) further argued: “In a few cases, unrelated individuals join forces to create the little fortresses. Unrelated colonies of Zootermopsis angusticollis, for example, fuse to form a supercolony with a single royal pair through repeated episodes of combat (Johns et al., 2009). In most cases of animal eusociality, the colony is begun by a single inseminated queen (Hymenoptera) or pair (others). In all cases, however, regardless of its manner of founding, the colony grows by the addition of offspring that serve as non-reproductive workers. Inclusive fitness theorists have pointed to the resulting close pedigree relatedness as evidence for the key role of kin selection in the origin of eusociality, but as argued here and elsewhere (Wilson & Wilson, 2007, 2008), relatedness is better explained as the consequence rather than the cause of eusociality.” Group living among kin (who might come together merely by virtue of being neighbors) will evolve more easily (West-Eberhard, 1978; Schwarz, 1988). Grouping by family can hasten the spread of eusocial alleles, but it is not the causative agent of eusociality (Nowak et al., 2010).

Sexually reproducing organisms should carry at least a twofold disadvantage with regard to the evolution of altruism since the benefits are at least multiplied by 0.5 while in clonal and parthenogenetic organisms with r = 1, altruism should arise much more easily. Clonality is often cited as a fundamental pillar of the evolution of the extreme aphid altruism that is sometimes observed (e.g., Krebs & Davies, 1993) and it has been claimed that the relatedness value of 1 often seen in aphids is the ultimate predisposition to altruism. However, the rarity of social aphid species (which represent just 1% of aphids) provides a clear demonstration that clonality alone is not sufficient to produce altruism (Stern & Foster, 1996; Pike & Foster, 2008). In a number of aphid species, kin recognition was sought, but not found (Aoki et al., 1991; Foster & Benton, 1992; Carlin et al., 1994; Miller, 1998a; 2004; Shibao, 1999). Clonal mixing may be facilitated by this inability. In a Pemphigus species, which is highly social, P. obesinymphae, the average level of clonal mixing is 41% with a range of 21–71% (Abbot et al., 2001). In P. spyrothecae, a species which is also highly social, microsatellite genotyping has been used to demonstrate that the average level of clonal contamination was 10%, albeit with a great range that varied from 0–59% (Johnson PCD et al., 2002). Abbot et al. (2001) were also able to confirm theoretical predictions by demonstrating that the invaders had (i) a vastly reduced propensity to defend and (ii) developmental rates that were accelerated such that they reached the reproductive stage more quickly than individuals of the host clone. One of the key correlates of aphid sociality is the galling habit. Whereas not all gall-forming aphids are social, all of the approximately 60 aphid species that are known to be social do form galls on a host plant at some point in their life cycle (Foster & Northcott, 1994). It thus follows that galling life must convey crucial selective predispositions to sociality, over and above its more general selective advantages such as protection from hygrothermal stress and enemies and improved nutrition (Price et al., 1987). Because galls are a rich and truly invaluable resource to the aphids which create and inhabit them, they are ideal examples of defense-worthy fortresses (Queller & Strassman, 1998). The implications that the galling habit holds for the evolution of sociality have been reviewed by Foster and Northcott (1994) and Pike and Foster (2008).

17.3 Social queuing as lottery

Traditionally mainly indirect benefits have been considered to explain cooperative behavior like cooperative breeding. The group augmentation hypothesis (Woolfenden, 1975; Rood, 1978; Brown, 1987; Kokko et al., 2001) states that if helpers in cooperatively breeding animals raise the reproductive success of the group, the benefits of living in a resulting larger group favor the evolution of helping behavior. Kingma et al. (2014) illustrated that direct benefits of group augmentation can accrue via different evolutionary mechanisms that relate closely to well-supported general concepts of group living and cooperation (Kokko et al., 2001; Dierkes et al., 2005; Bergmüller et al., 2007; Clutton-Brock, 2009a; Heg & Taborsky, 2010; Sumner et al., 2010; Kingma et al., 2011; Wong & Balshine, 2011; Zöttl et al., 2013a). Helping might increase survivorship, or the possibility of eventually obtaining reproductive dominance in that group (Reyer, 1980; Wiley & Rabenold, 1984; Courchamp et al., 1999; Kokko & Johnstone, 1999; Pen & Weissing, 2000; Griffin & West, 2002).

Queuing processes, i.e. waiting peacefully in line to acquire dominant status in the future, seem to occur in a wide variety of taxa (birds, fish, mammals, and invertebrates) and life histories (van de Pol et al., 2007; Field & Cant, 2009; Kingma et al., 2011; Sharp & Clutton-Brock, 2011; Wong & Balshine, 2011). Queue-like systems range from queues for mating opportunities (Schwagmeyer & Parker, 1987), social and breeding position in group-living and cooperative breeding species (Wiley & Rabenold, 1984; East & Hofer, 2000; Heg et al., 2005; Mitchell, 2005), positions on the lek (Kokko et al., 1998), or access to harems or colonies (Poston, 1997; Voigt & Streich, 2003) to queues for high-quality territories (Stacey & Ligon, 1991; Zack & Stutchbury, 1992; Ens et al., 1995; Ekman et al., 2001b). The queue hypothesis suggests that individuals maximize lifetime fitness by strategically waiting (queuing) for high quality breeding opportunities to become available, instead of immediately accepting a low-quality breeding opportunity (Zack & Stutchbury, 1992; Ens et al., 1995). In most cooperative vertebrates and many primitively social insects, subordinates apparently remain in or join groups as nonbreeders in expectation of inheriting a breeding position in the future (in Hymenoptera: e.g., Strassmann & Meyer, 1983; Samuel, 1987; Hughes & Strassmann, 1988; Field et al., 1999; Monnin & Ratnieks, 1999; Cant & Field, 2001; and in vertebrates: Wiley & Rabenold, 1984; Stacey & Koenig, 1990; Emlen, 1991; Creel & Waser, 1994; Poston, 1997; East & Hofer, 2001; Buston, 2003b, 2004).

Queuing can be regarded as conservative behavior compared to the risky alternative, dispersal. In a range of queuing species, offspring that delay dispersal and wait on their natal territory either gain better quality territories or are more likely to breed than their siblings that disperse as juveniles (Woolfenden & Fitzpatrick, 1984; Strickland, 1991; Walters et al., 1992; Ekman et al., 2001b; Green & Cockburn, 2001; Komdeur & Edelaar, 2001). Studies that measure bird survivorship and fecundity across a range of habitats reveal that individuals using the poorest measured habitats have lifetime fitnesses of less than one (Stacey & Ligon, 1987; Ligon & Ligon, 1990; Rabenold, 1990; Fitzpatrick et al., 1991; Komdeur, 1992; Ekman et al., 1999; Ridley & Sutherland, 2002). The decision to queue instead of leaving their natal nests and pursue other strategies may be a bet-hedging strategy (Yuan et al., 2006; Field, 2008). Subordinates can inherit dominance if they outlive those above them in the hierarchy (Strassmann & Meyer, 1983; Wiley & Rabenold, 1984; Field et al., 1999, 2006; Monnin & Peeters, 1999; Monnin & Ratnieks, 1999; Cant & Field, 2001). In many cooperative breeders, a proportion of helpers can inherit their home territory (Woolfenden, 1975; Rood, 1978, 1990; Wiley & Rabenold, 1984; Dierkes et al., 2005; Hawn et al., 2007; Field & Cant, 2009; Sumner et al., 2010; Kingma et al., 2011, 2014; Leadbeater et al., 2011; Sharp & Clutton-Brock, 2011; Marino et al., 2012). Ragsdale (1999) defined the resource inheritance, a form of future direct benefit, as the probability of inheriting valuable resources multiplied by the expected number of offspring that an individual would produce after it inherits the resources (relative to a lone breeder). With resource inheritance, stable associations will form over a greater range of conditions, thus reducing the need for a ‘social contract’. Inclusion of resource inheritance in reproductive skew theory generates predictions that are relevant to many social systems, including (under some conditions) parental facilitation, ‘lazy workers’, helping for ‘payment’, and complete skew when relatedness is zero or cooperative benefits are absent (Ragsdale, 1999). In green woodhoopoes (Hawn et al., 2007) and female dwarf mongooses (Rood, 1990), 58% and 43% of breeding positions, respectively, are obtained by inheritance. The scheme that individuals form a strict queue and move up in rank only when one of those ahead of them in the queue dies fits some species well (e.g., primitively social wasps: Field et al., 1999; Cant & Field, 2001, 2005; Sumner et al., 2002; vertebrates: Wiley & Rabenold, 1984; East & Hofer, 2001; Buston, 2003b, 2004). Alternative possibilities exist, however (Monnin & Ratnieks, 1999; Cant & Johnstone, 2000; Cant et al., 2006b). For example, in some termites, ants, mole-rats, and many primates, nonbreeders engage in vicious contests for any breeding vacancy that may arise (Pollock & Rissing, 1985; Pusey & Packer, 1987b; Myles, 1988; Clarke & Faulkes, 1997; Thorne, 1997). Even in the absence of a breeding vacancy, low rankers may jump the queue by attacking those of higher rank(e.g., groove-billed anis, Vehrencamp et al., 1986; dunnocks, Davies, 1992; hyenas, East & Hofer, 2001; many primates, Walters & Seyfarth, 1987).

Stable queues will select for the production of new recruits by helpers because the latter eventually will not only inherit a breeding position but also a group of helpers (the former recruits), and this exemplifies the benefits of group augmentation by delayed reciprocity (Wiley & Rabenold, 1984; Kokko et al., 2001a). An important feature of queuing decisions is their frequency-dependent nature, i.e. how many other individuals are queuing for the same opportunity (Shreeves & Field, 2002). Queuing is probably favored by selection because it confers a higher probability of ascending in rank than either dispersing or contesting. When ecological constraints are harsh and there is a real threat of eviction, individuals may tolerate non-breeding positions in a society purely because of their potential to realize benefits in the future (Emlen, 1991; Kokko & Johnstone, 1999; Ragsdale, 1999; Buston, 2004). In queuing systems where joining decisions are based on inheritance prospects, egalitarian groups may be less attractive to a potential joiner than despotic groups. This is because the huge payoff of becoming the sole breeder in a despotic group may more than make up for the relatively low probability of reaching that position. By contrast, a subordinate joining an egalitarian group has a relatively high chance of inheriting one of the many breeding positions, but on doing so its offspring will face competition from those of other breeders for access to group resources (Cant & English, 2006).

Analysing models according to whether or not helpers can inherit their parents’ territory, Pen and Weissing (2000) concluded that (i) territory inheritance always promotes cooperative breeding; (ii) if territories are not inherited, neither ecological constraints nor variation in life-history traits predict interspecific variation in cooperative breeding; and (iii) if territories are inherited, the mechanism of density regulation is crucial in determining which factors promote cooperative breeding. If density dependence acts on the probability to obtain a free territory or on the survival of dispersers, variation in ecological constraints cannot explain variation in cooperative breeding. Lower adult mortality favors helping, not because it reduces the availability of free territories, but because it enhances the direct benefits of helpers. If density dependence acts on fecundity, lower probability of obtaining a free territory and lower survival of dispersers promote cooperative breeding. In this case, lower adult mortality works against the evolution of helping. Pen and Weissing (2000) suggest that the difference between birds and social insects in the covariance between cooperative breeding and life-history traits is due to different mechanisms of density regulation that operate in these taxa, and explain how natural selection on habitat choice might have caused these different mechanisms to operate.

Nest inheritance can explain the presence of unrelated helpers in a classic social insect model, the primitively eusocial wasp Polistes dominulus. Leadbeater et al. (2011) found that subordinate helpers produced more direct offspring than lone breeders, some while still subordinate but most after inheriting the dominant position: about 32% of the reproduction of subordinates came from sneaking in eggs while the dominant was still alive and 68% came from inheriting the position of the dominant in the nest after the latter had died. Toyoizumi and Field (2014) analyzed queues formed by social hover wasps to inherit the dominant position in the nest. The analysis of various potential strategies of hover wasps (lone breeding, simple social queuing and social queuing with division of labor) shows that lone breeding and simple social queuing that includes extended parental care fail to maintain a viable population with realistic parameter values. On the other hand, division of labor, which extends queen lifespan, will significantly increase the productivity of a simple social queue. The numerical analysis shows that the impact of division of labor on nest productivity is slightly greater than the impact of simple social queuing itself. Thus, division of labor might be one of the main benefits of social queues. If helping roles within groups are assigned through a lottery rather than being genetically determined, maximum degrees of “altruism” can evolve in groups of nonrelatives of any size (Gadagkar, 1991a; Wilson, 2001; Avilés et al., 2004).

Frequency-dependent queuing processes have been used to explain the logic of delayed breeding (Ens et al., 1995), territory choice (Kokko & Sutherland, 1998; Kokko et al., 2001b; Pen & Weissing, 2001), reproductive skew (Kokko & Johnstone, 1999), and cooperative breeding (Pen & Weissing, 2000; Kokko & Ekman, 2002). Natal philopatry characterizes the social dynamics in some communally breeding birds (Brown, 1987; Mumme, 1997), carnivores (Moehlman, 1986; Rood, 1986), cetaceans (Connor, 2000), primates (Pusey & Packer, 1987b), proboscideans (Archie et al., 2006b), rodents (Michener, 1983; Solomon, 2003; Armitage, 2007; Lacey & Sherman, 2007; Ebensperger & Hayes, 2008; Lucia et al., 2008), or mammals in general (Waser, 1988; Lefebvre et al., 2003). Experiments that manipulated resource levels in western bluebirds and carrion crows (Dickinson & McGowan, 2005; Baglione et al., 2006) indicate that predictable access to resources promotes offspring philopatry and helping. In evolutionary equilibrium, the competition, and thereby queuing time, for high-quality breeding positions increases up to a point at which it pays to accept low-quality breeding positions at a young age (Ens et al., 1995). Queuing models that take into account even the smallest individual quality differences, which are probably plentiful in nature, result in individuals using queuing strategies completely different from those in models that assume qualitatively equal competitors (van de Pol et al., 2007). Under most ecological scenarios, queuing is obviously more favored than dispersing. Since there are benefits to be gained from remaining in the natal territory, larger and socially dominant siblings within broods are more likely to stay and to expel siblings from natal territories (Strickland, 1991; Ekman et al., 2002).

17.4 Parental care and assured fitness returns

In many taxa, the survival of offspring is dependent on parental care, which is defined as any trait that enhances the fitness of offspring and originated/is maintained for this function (Smiseth et al., 2012). Parental care is common across animal taxa and increases offspring survival and/or quality in a range of species (Clutton-Brock, 1991). Parental care tends to increase the current reproductive success of the parent, but providing care is physically costly and decreases the amount of resources an individual can invest into future reproduction (Williams, 1966a; Trivers, 1972). A variable environment that affects adult or egg death rates can either increase or decrease the fitness of parental care relative to that in a constant environment, depending on the specific costs of care. Ecological factors such as harsh environments, ephemeral food sources or predation pressure are broadly accepted as evolutionary drivers of parental care (Wong et al., 2013). Increasing parental care across different life-history stages can increase fitness gains in variable environments. Wilson (1975) and Clutton-Brock (1991) suggested that care is most likely to evolve when environmental conditions are harsh and competition for resources is intense, as these are the conditions under which the benefits of care are likely to be large. Klug and Bonsall (2010) showed that parental care can evolve from an ancestral state of no care under a range of combinations of ecological conditions and life histories (e.g. egg, juvenile, and adult mortality rates, adult reproductive rate, egg maturation rate, and the duration of the juvenile stage). Costly investment in care is expected to affect the overall fitness benefits, the fitness optimum and rate of evolution of parental care. The authors compared the evolution of parental care in a constant versus a variable environment. They found that in a variable environment, the selection of parental care depends on the interaction between environmental variability, the life-history traits affected by such variability, and the specific costs of care (Bonsall & Klug, 2011). For example, environmental variability reduces selection for parental care when the costs of care are associated with both reduced parental survival and reproductive rate, but favors parental care if the only cost of care is a reduced parental survival rate. Whereas recent theoretical developments support the idea that ecological agents of selection in combination with preexisting life histories are important, they also revealed that ecological agents on their own are usually not sufficient for the emergence of parental care (Klug & Bonsall, 2010; Klug et al., 2012), leaving scope for other important factors (Wong et al., 2013). One of them is the social environment, which results from interactions between the two parents (Smiseth & Moore, 2004), between parents and offspring (Mas et al., 2009) or among siblings (Ohba et al., 2006). Such social interactions are indeed known to shape the benefits/costs ratio of care and, hence, possibly to influence the strength of natural selection on parental care once a basic level of care has evolved (Royle et al., 2002; Smiseth et al., 2012). Extended parental care (Stubblefield & Charnov, 1986; Queller, 1994b; Queller & Strassmann, 1998) has been associated with the evolution of eusociality in Hymenoptera. The role of maternal behavior in the evolution of eusociality was supported by wasp brain gene expression that, in workers, was more similar to that in foundresses, which show maternal care, than to that in queens and gynes, which do not (Toth et al., 2007). Insulin-related genes were among the differentially regulated genes, suggesting that the evolution of eusociality involved major nutritional and reproductive pathways (Toth et al., 2007).

One factor that may promote group living is the idea of assured/delayed fitness returns (Gadagkar, 1990c). The key preadaptation for eusociality in the social Hymenoptera is progressive provisioning, a behavior that in solitary species arises by individual direct selection (Wilson, 2008). Most nonsocial wasps and bees are mass provisioners, sealing each egg into its own cell containing all of the food required to reach maturity. Other species instead provision progressively, feeding their offspring only gradually as they develop and usually provisioning more than one offspring simultaneously (Field, 2005). In order to obtain a return on its reproductive investments, an individual progressive provisioner that nests solitarily must survive until its offspring are no longer dependent on it for their survival. If the individual dies before its offspring have become independent, the offspring will also die and the individual’s fitness will be zero. In contrast, an individual that contributes only partially to raising young in a social group will not have wasted its investments when it dies, provided the surviving adults can complete rearing the offspring through to independence. This safeguard reduces the fitness cost of death, and therefore also allows individuals to engage in more risky, but potentially more rewarding, strategies (Clark & Dukas, 1994). The longer the brood is dependent on the continued presence of the adult for survival, the greater the potential advantages of this insurance. In progressive provisioning species, which have an extended period of brood dependence, assured fitness returns may therefore be especially important (Field, 2005). Assured fitness returns models argue that patterns of adult mortality combine with offspring dependence on parental care to select for group living (Queller 1989, 1994b, 1996; Strassmann & Queller, 1989; Gadagkar, 1990c, 1991; Bull & Schwarz, 2001; Field & Brace, 2004). Studies of progressive provisioning species (those species that must feed offspring repeatedly throughout larval development) such as paper wasps and hover wasps (Vespidae), and allodapine bees (Apidae) support assured fitness returns models (Queller, 1989; Gadagkar, 1990c; Bull & Schwarz, 1996, 1997; Schwarz et al., 1997, 1998; Field et al., 2000; Hogendoorn et al., 2001). There are well-supported examples of assured fitness returns in social groups with dedicated workers, such as the hover wasp Liostenogaster flavolineata (Field et al., 2000), the paper wasp Polistes dominulus (Shreeves et al., 2003), but also in species without workers/helpers such as the apoid wasp Microstigmus nigrophthalmus (Lucas & Field, 2011) and the social spider Anelosimus studiosus (Jones et al., 2007).

In a mass provisioner, females provide all necessary food for offspring development before laying an egg (Wilson, 1971; Michener, 1974). Although offspring could potentially reach maturity even if the parent died immediately after oviposition, they may be dependent on other forms of parental care, such as defense against nest predation and parasitism (Queller, 1994a; Eickwort et al., 1996; Kukuk et al., 1998; Forbes et al., 2002b; Smith et al., 2003). Nests of solitary females are unattended for a greater portion of each day than are nests with multiple females. One benefit of shared nests may be increased vigilance against predators and parasites (Lin & Michener, 1972; Alexander, 1974; Abrams & Eickwort, 1981; Rehan et al., 2011). In the mass provisioner sweat bee Megalopta genalis, ant predation selects for group living not because the brood is safer when defended by more than one bee, but because having two or more bees cohabiting increases the likelihood that at least one bee will remain to defend the nest (Smith et al., 2003). Group-nesting females have lower per capita reproductive output than solitary females (Soucy et al., 2003), a paradox noted already by Michener (1964) for many hymenopterans. Assured fitness returns models represent a type of conservative bet-hedging strategy in a general sense, because a solitary female that survives the period of her brood’s dependence would have the highest reproductive success, but brooding females in multifemale colonies trade the possibility for maximum reproductive success for a greater mean reproductive success (Soucy et al., 2003; Jones & Riechert, 2008; Rehan et al., 2011).

19. Prosocial intent and anthropocentric arguments


Summary
Sociobiology is burdened by a multitude of teleological, anthropocentric and moralizing concepts. Intentional wording like altruism, spite, selfish, cheating, and sacrifice abound. George C. Williams argued vigorously against the use of terms “burdened with value judgment and emotional flavour,” such as altruism, and proposed a value-neutral terminology.

Evolution is far-sighted (Altenberg, 2005; Heininger, 2015) but it has no foresight. 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). The literature on social interactions, however, is replete with teleological arguments. A good deal of concepts in sociobiology are anthropocentric post hoc statements, biased by theory, misguided by moral prejudices: they judge the intentions of the agents by the final outcome irrespective of the underlying processes. Earlier (Heininger, 2012), I have emphasized that a lively style of writing cannot do without teleological wording. However, this wording should clarify issues but not distort and misrepresent them. Anthropocentric moral judgments should altogether be banned from discussions of animal social interactions.

Studying the structure of an atom is not personal, and neither is studying, for example, night vision in mammals. Studying altruism can be personal, however, because we all want to understand the origins of goodness (Dugatkin, 2007). Sociobiology is burdened by a multitude of anthropocentric and moralizing concepts. In a discussion of social adaptations, George C. Williams argued vigorously against the use of terms “burdened with value judgment and emotional flavour,” such as altruism, and proposed a value-neutral terminology (Williams & Williams, 1957, p. 32–33). “Altruist” and “cooperator” were to be replaced with “social donor” or, simply, “donor.” Noncooperators–now routinely labeled as selfish or “cheaters”–were “nondonors.” (Lyon, 2011). I fully support this attitude and argue that the moralizing concepts have blocked, or at least distorted, the awareness for evolutionary patterns, patterns that are general and have easily been recognized in other areas of evolutionary biology that are less biased by value judgments.

20. Kin selection and inclusive fitness: the fall of Hamilton


…that an opinion has been widely held is no evidence whatever that it is not utterly absurd….
Bertrand Russell (1929)

Inclusive fitness theory, often called kin selection theory, is both mathematically and biologically incorrect.
Edward O. Wilson, 2012

Summary
From a systems biology perspective, Hamilton’s rule is simplistic, biased by observation selection, static, and parochial. Both Einstein’s relativity theory and Hamilton’s rule are hypothetico-deductive models. For this type of models applies what Einstein said: “No amount of experimentation can ever prove me right; a single experiment can prove me wrong.” Hamilton’s rule and its predictions have been disproved repeatedly. Several of the basic, often implicit, assumptions of Hamilton’s rule are flawed: linearity, egalitarian society, deterministic reproductive success in stochastic environments, haplodiploidy as cause of eusociality. An alternative theory of eusociality was advanced 40 years ago (Alexander, 1974; Michener & Brothers, 1974): The worker caste has arisen by individual selection on mothers resulting in their control of the activities of their female offspring. The queen is somehow able to keep them in the nest and diminish their reproductivity, and in a sense is parasitic upon her daughters. The theory was virtually ignored. 
In response to the increasingly restricted scope of the 
kin selection theory, advocates of the theory tried to broaden the definitions of the behavior/outcome of altruism and inclusive fitness, and use the highly speculative greenbeard argument as an argumentative red herring. That the plausible alternative theory was widely ignored but the highly speculative greenbeard argument is often discussed is a telling witness to the biased scientific climate in which Hamilton’s rule flourished.

From a systems biology perspective, Hamilton’s rule is simplistic, biased by observation selection, static, and parochial. Hamilton’s rule is hypothetico-deductive. The hypothetico-deductive model or method proceeds by formulating a hypothesis in a form that could conceivably be falsified by a test on observable data. A test that could and does run contrary to predictions of the hypothesis is taken as a falsification of the hypothesis. This method can never absolutely verify (prove the truth of) the hypothesis but can only falsify it (Hempel, 1966). This is what Einstein meant when he said, “No amount of experimentation can ever prove me right; a single experiment can prove me wrong.” Hamilton’s rule and its predictions have been disproved repeatedly, e.g. disproval of the haplodiploidy hypothesis, disproval that inclusive fitness is the reason for eusociality (in fact, it is force and chemical manipulation). Kin selection theory in sociobiology is the result of silo thinking, silo education, and silo research. Those using silo thinking and analysis seek to simplify, to reduce complexity (Waldman, 2007). Advocates of the inclusive fitness theory prefer to live in a state of self-imposed blindness maintaining that the tail of the elephant is a snake (see chapter 2). For example Wild et al. (2010) wrote: “The inclusive-fitness approach assumes within-individual selection is negligible (Grafen, 2006), which is justified by the huge empirical success of the theory (Krebs & Davies, 1993).” The sentence was written after within-individual selection had been demonstrated conclusively for 35 years (see chapter 4.1.1). Moreover, the insights of complexity theory identifying cooperation as emergent behavior of complex systems, and neurobiology emphasizing the context-dependency and pleiotropy of social behavior, have been widely ignored.

According to West et al. (2007a), “the importance of Hamilton’s work cannot be overstated–it is one of the few truly fundamental advances since Darwin in our understanding of natural selection.” The advocates of the theory maintain that there is no doubt that kin selection theory has been very successful in explaining a wide range of phenomena and that many empirical studies support their theory. But often the connection that is made between data and theory is superficial, and as argued in chapter 4.2.3.2 is spurious due to a common “confounding factor”. Behavioral studies demonstrate that individuals in small-scale societies preferentially aid close kin over more distant kin and nonkin (e.g., Chagnon & Bugos, 1979; Chagnon, 1981; Hawkes, 1983; Betzig & Turke, 1986; Hames, 1987; Betzig, 1988; Flinn, 1988; Gurven et al., 2000; Patton, 2005). Such nepotistic biases are often cited as evidence that indirect fitness payoffs (Hamilton, 1964; Maynard Smith, 1964) have shaped human social interactions. Allen-Arave et al. (2008) considered the roles of kin selection and reciprocal altruism in maintaining nepotistic food transfers on an Ache reservation in northeastern Paraguay. Households do not primarily direct aid to related households that receive larger comparative marginal gains from food intake as would be predicted under kin selection theory. Instead, (i) food transfers favor households characterized by lower relative net energy production values irrespective of kinship ties, (ii) households display significant positive correlations in amounts exchanged with each other, suggesting contingency in food transfers, and (iii) kinship interacts with these positive correlations in amounts households exchange with each other, indicating even stronger contingency in sharing among related households than among unrelated households. While kin are preferred recipients of food aid, food distributions favor kin that have given more to the distributing household in the past rather than kin that would benefit more from the aid. Such discrimination among kin accords better with reciprocal altruism theory than with kin selection theory (Allen-Arave et al., 2008).

“For testing the usefulness of inclusive fitness theory it is not enough to obtain data on genetic relatedness and then look for correlations with social behavior. Instead one has to perform an inclusive fitness type calculation for the scenario that is being considered and then measure each quantity that appears in the inclusive fitness formula. Such a test has never been performed” (Nowak et al., 2010). Edward O. Wilson, famous for his work on insect societies and sociobiology and once a forceful advocate of kin selection, now argues that kinship plays a minor part in the evolution of ant, bee, termite and other social insect colonies (Wilson, 2005; Wilson & Hölldobler, 2005; Wilson & Wilson, 2007; Nowak et al., 2010). More important, he says, are the ecological factors that make social living so successful. Cautioning against an excessive concentration on the role of genetic relatedness in driving social evolution, Lin and Michener (1972) drew attention to a large number of social insect species where sterility is absent or incomplete and suggested the possibility that individuals in such groups may be selected to come together for mutual benefit (Gadagkar & Bonner, 1994). This theory of mutualism has often been dismissed as incapable, by definition, of explaining the evolution of a sterile worker caste because the term mutualism suggests that both or all participants benefit. In contrast, the sterile worker caste and the fertile queen are not usually thought of as benefiting equally from the associations (e.g., Crozier, 1977; Itô, 1989). The insight (since more than 25 years) that the help of the workers is enforced (remember, punishment can select for any behavior [Boyd & Richerson, 1992]) should have led to a rethinking of this argument but only few were able to do this.

The flawed track on which evolution theory embarked was not initiated by Hamilton. In fact, it dates back to Darwin with his assumption of constant environments. Yet, in 1876 in a letter to Moritz Wagner he wrote: “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”. This late insight, however, was widely ignored. Plausible as it is, Hamilton’s rule was instrumental in paving this flawed track and stifling any attempt to look for alternative explanations. Apart from being “a decades-long distraction in the field that is theoretically unsound, unnecessarily focused on genetic relatedness and poorly supported by the empirical evidence” (Bourke, 2011a), two highly contentious issues originate from this hypothesis: the individual selection-group selection controversy and the kin selection-group selection controversy.

The theory of kin selection was criticized in two studies, one published in 1998 (Alonso, 1998) and another in 2002 in PNAS (Alonso & Schuck-Paim, 2002). Alonso and Schuck-Paim argued that the behaviors which kin selection attempts to explain are not altruistic (in pure Darwinian terms) because: (i) they may directly favor the performer as an individual aiming to maximize its progeny (so the behaviors can be explained as ordinary individual selection); (ii) these behaviors benefit the group (so they can be explained as group selection); or (iii) they are by-products of a developmental system of many “individuals” performing different tasks (like a colony of bees, or the cells of multicellular organisms, which are the focus of selection). They also argued that the genes involved in sex ratio conflicts could be treated as “parasites” of (already established) social colonies, not as their “promoters”, and, therefore the sex ratio in colonies would be irrelevant to the transition to eusociality. Those papers were mostly ignored until they were re-edited by Martin Nowak, Corina Tarnita, and Edward O. Wilson. These latter authors argue that “Inclusive fitness theory is not a simplification over the standard approach. It is an alternative accounting method, but one that works only in a very limited domain. Whenever inclusive fitness does work, the results are identical to those of the standard approach. Inclusive fitness theory is an unnecessary detour, which does not provide additional insight or information.” A mathematically meaningful approach to inclusive fitness cannot be performed for the majority of evolutionary processes (Nowak et al., 2010), and the linear regression method (Hamilton, 1970; Queller, 1992c; Gardner et al., 2011) does not provide meaningful insights and cannot make empirical predictions (Allen et al., 2013; Wilson & Hölldobler, 2014). Hamilton’s rule is a linear regression to generally non-linear data (Damore & Gore, 2012). Van Dyken et al. (2011) wrote: “If doubling the number of altruists in a group always exactly doubles the total benefit of altruism, then social interactions are said to be linear. However, if the total benefit is more than doubled (or less than doubled) or if it is doubled only when altruists are at a specific frequency within the group, then social interactions are nonlinear. A disadvantage of this approach is that the number of possible nonlinear fitness functions is vast, making it difficult to extract general predictions.” The last sentence provides in a nutshell the dilemma of mathematical models: the trade-off between the degree of abstraction from reality and the mathematical tractability. The development of systems of equations, however sophisticated, may well have very little to tell us about the social world. Nonlinearity challenges the universalism of the Newtonian conception at the level of the real world which we inhabit and experience (Byrne & Callaghan, 2014).

In conventional quantitative accounts of causality, change in the value of an effect is proportionate to changes in causal elements. The simplest expression of this is through a linear equation (Byrne & Callaghan, 2014). In a world of nonlinearity of biological processes and social interactions (Avilés, 1999; Avilés et al., 2002; Chuang et al., 2010; smith et al., 2010) caused by a variety of nonlinear processes such as density- and frequency-dependent selection (Mueller et al., 1991; de Mazancourt & Dieckmann, 2004), diminishing returns (Orr, 2005; Barrick et al., 2009; MacLean et al., 2010b; Chou et al., 2011; Khan et al., 2011; Flynn et al., 2013), pleiotropic trade-offs (see chapter 9.3), plastic genotype-phenotype maps (Müller, 2007; Gjuvsland et al., 2013; Badyaev & Walsh, 2014), epistasis (MacLean et al., 2010b; Flynn et al., 2013; Anholt & Mackay, 2015) and degeneracy (Edelman & Gally, 2001; Whitacre & Bender, 2010; Heininger, 2013) it is simplistic to think that such a complex trait like social behavior should (in addition to ecological cost-benefit ratio) be regulated by linear genetic kinship. For instance, epistasis dominates the genetic architecture of complex traits, including behaviors (Yamamoto et al., 2009; Huang et al., 2012; Swarup et al., 2012, 2013). Wright (1931, 1969) considered epistasis to be ‘ubiquitous’ and stated that “The inadequacy of any evolutionary theory that treats genes as if they had constant effects, favourable or unfavourable, irrespective of the rest of the genome, seems clear“ (Wright, 1969, p. 88). He emphasized that the “…existence [of epistasis] must be taken as a major premise in any serious discussion of population genetics and evolution” (Wright, 1969, p. 105). Degeneracy is the ability of elements that are structurally different or of molecular tools 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) proposed that degeneracy may act both as a source of robustness and evolvability in biological systems (Whitacre & Bender, 2010). Due to the degeneracy of molecular tools in addition to the stochasticity of the genotype-phenotype mapping no certain molecular process can be assigned to a phenotypic effect (Heininger, 2013). Taking the nonlinearity of social interactions (Avilés, 1999; Avilés et al., 2002; Chuang et al., 2010; smith et al., 2010; Byrne & Callaghan, 2014) into account, inclusive fitness in a nonlinear public goods game points in the wrong direction (van Veelen, 2009). In addition, while a two-player situation still allows for (adjusted) formula’s that do use relatedness, a slightly less simple example shows that with groups larger than two, relatedness can be the wrong population characteristic to look at. This implies that the prediction of the model cannot be given in a formula with costs, benefits and relatedness only (van Veelen, 2009).

Hamilton’s rule was (and many mathematical models still are) shaped by an egalitarian worldview in which autonomous individuals “decide” to forego their reproduction and help their kin to ensure the representation of their genes in the next generation. In nature, however, the vast majority of cooperative systems are characterized by asymmetric conflicts between dominants and subordinates over limited resources/reproductive opportunities. The “decision” of subordinates to help is not self-determined but enforced by despotic dominants and the prevailing environmental conditions that limit the subordinates’ options for independent reproduction. As soon as the ecological conditions favor their own reproductive activity they “vote with their feet and hope for the best” (Betzig, 2004).

Nowak et al. (2010) wrote: “For many models we find that cooperators are favoured over defectors for weak selection, if a condition holds that is of the form (Nowak, 2006a; Ohtsuki et al., 2006; Traulsen & Nowak, 2006; Taylor et al., 2007; Antal et al., 2009; Tarnita et al., 2009a): ‘something’ > c/b. This result is a straightforward consequence of the linearity introduced by weak selection (Tarnita et al., 2009b) and has nothing to do with inclusive fitness considerations.” As Nowak et al. (2011) pointed out, even in simple models, if cost and benefit are parameters of individual actions then Hamilton’s rule almost never holds (Cavalli-Sforza & Feldman, 1978; Karlin & Matessi, 1983; Nowak et al., 2010; smith et al., 2010; Damore & Gore, 2012). Since reproduction per definition produces kin-structured populations (Blouin, 2003) this ‘something’ could be reproduction. Another ‘something’ may be delayed dispersal (see chapter 4.2.3). The vast majority of multicellular organisms develop clonally via ‘staying together’ after mitotic reproduction. Evolutionary theory predicts that cells’ staying together provides several key advantages over multicellular construction via cells ‘coming together’, a prediction that has been recently corroborated experimentally in Saccharomyces cerevisiae (Pentz et al., 2014). On the other hand, Goodnight (2005) showed that what Hamilton called relatedness is more generally interpreted as the proportion for variance among groups, and that many processes in addition to relatedness can increase the variance among groups.

An alternative theory of eusociality was advanced 40 years ago (Alexander, 1974; Michener & Brothers, 1974): The worker caste has arisen by individual selection on mothers resulting in their control of the activities of their female offspring. The influence that even a solitary mother hymenopteron exercises on her immature progeny by feeding and protecting them continues, in eusocial species, past maturation of the daughters. The queen is somehow able to keep them in the nest and diminish their reproductivity, and in a sense is parasitic upon her daughters (Michener & Brothers, 1974). The theories were virtually ignored (as of 02-20-2015, according to Google Scholar 150 citations for Michener & Brothers [1974] vs. 12,151 for Hamilton [1964]).

20.1 Attempts to save Hamilton's rule

The “survival kit” of Hamilton’s rule is: (i) ignore all types of cooperation that are not compatible with your framework (e.g. cooperation among nonkin, mutualisms, symbioses representing the vast majority of cooperative behaviors); (ii) call acts of fitness transfer “altruistic” (e.g. “reciprocal altruism”, “enforced altruism”, “coin-flipping altruism”) that are not “altruistic”; (iii) ignore or downplay the fact that game theoretic models are indifferent to the degree of relatedness between the cooperators; (iv) ignore the fact that dominance hierarchies in eusociality have both an “altruistic” (worker) and exploitative/parasitic (queen) side making Hamilton’s rule a zero-sum game; (v) ignore the fact that reproductive success in uncertain, unpredictable environments is highly stochastic and future inclusive fitness gains should be heavily discounted.

When a theory comes under critique, the range of conditions under which it applies often has to be restricted (e.g., see Doebeli & Hauert, 2006; Nowak et al., 2010). Conversely, the advocates of the kin selection theory try to broaden the definitions of the outcome/behavior that the theory putatively describes, presumably in the attempt to maintain the theory any meaningful area of scope. I will illustrate the latter attempts by means of three examples:

1. Altruism is defined as self-sacrifice with no apparent personal reward (Bryan & Test, 1967; Bar-Tal, 1976; Hoffman, 1981). In addition, altruism is commonly expected to be voluntary (Unger, 1991; Smith et al., 1995; Ziemek, 2006; Carpenter & Myers, 2010; Linardi & McConnell, 2011). The insight that helping in eusocial colonies is coerced by queen pheromones and worker policing (Ratnieks, 1988; Ratnieks & Visscher, 1989; Liebig et al., 1999; Ratnieks & Wenseleers, 2005; Smith AA et al., 2009), even in a clonal ant without genetic conflicts (Hartmann et al., 2003), led scientists to create the semantic monster “enforced altruism” (e.g. Wenseleers & Ratnieks, 2006b; Ratnieks & Wenseleers, 2008; Ratnieks & Helanterä, 2009). In a similar vein of thought, a putative gene for queen pheromone production that must be expressed by termite queens to suppress worker reproduction (Korb et al., 2009) was dubbed “altruism gene” (Thompson et al., 2013). Obviously, enforced altruism is not voluntary but with benefit to the “giver”, i.e. the avoidance of punishment. Punishment is such a strong evolutionary force that it allows the evolution of anything (Boyd & Richerson, 1992). Within the conceptual framework of “enforced altruism”, the tithes payed to feudal lords and payments of tribute of former times although enforced, were acts of altruism. Obviously for this type of “coerced altruism” no kinship ties are necessary. Within this logic, robbery is also “enforced altruism”: the enforced transfer of assets that lower the fitness of the “benefactor” and increase the fitness of the “beneficiary”. Even prey-predator dyads can be characterized as “enforced altruism”. All of these “transfers” are enforced and no one, except some adherents of the kin selection theory, would call them altruistic. Random fates in the lotteries of life (see chapter 5.4) have been termed “coin-flipping altruism” by Cooper and Kaplan (1982). According to this logic every participant of a lottery, by buying a lottery ticket, commits an act of altruism towards the eventual winner(s) of the lottery. Likewise, the clients of insurance companies in which the insured event does not occur would act altruistically versus the ones in which the insured event occurs. And are gamblers at the casino and/or stock traders (Statman, 2002; Gao and Lin, 2011; Liao, 2013) when they lose altruists towards the winners? In the same vein of thought the winner of a lottery should be deemed to be selfish. Likewise, the term “reciprocal altruism” is misleading because this type of interaction is not altruistic but mutually beneficial (e.g. West et al., 2007a, 2007c; Douglas, 2008; Davies et al., 2012).

2. Another example concerns recently extended definitions of inclusive fitness. Since current concepts of fitness put much emphasis on the representation of the individual’s genes in the next generation, expected fitness is defined by reproduction rather than survival of the individual (the latter is only evolutionarily relevant in the tautological sense of “survive to reproduce”). The definition of fitness as expected number of offspring has a one-generation time-scale (Sober, 2001). In principle, however, there is no a priori limit on the size of the time frame over which the concept of fitness may have to be stretched (Beatty & Finsen, 1989; Sober, 2001; Abrams, 2009). Grafen (1999) defines “reproductive value” as “a measure of the contribution made by an individual to the gene pool in distant generations.” Long-term concepts of fitness (Thoday1953; Cooper, 1984; Beatty & Finsen, 1989; Sober, 2001; Abrams, 2009; McNamara et al., 2011) suggested that fitness should be defined as the probability of leaving descendants in the very long run. Traditionally, inclusive fitness theory partitioned fitness components into direct fitness effects of the focal individual’s own reproduction and indirect fitness effects due to reproduction of the focal individual’s non-offspring relatives (Hamilton, 1964; Price, 2007). “Parents helping children is not an example of kin selection, but rather straightforward selection-maximizing direct fitness” (Rand & Nowak, 2013). The basic idea of inclusive fitness theory was purportedly enunciated by JBS Haldane in a pub when he quipped that he would sacrifice himself by jumping into a river to save two brothers or eight cousins, a view he only expressed in print at a much later date (Haldane, 1955). Recently, however, this traditional concept was extended by some authors to the focal individual’s own reproduction and parental care. They considered parental care a prime example for an altruistic trait that evolved to enhance the fitness of the recipients of care (offspring) at the expense to the donor of care (parents) (Ekman et al., 1994; Royle et al., 2012; Smiseth et al., 2012; Wong et al., 2013). “Parental care is beneficial to the offspring because it increases their direct fitness. From the perspective of the parent (or a parental care gene), the offspring fitness benefit is an indirect fitness benefit because the fitness of the genetically related recipient of care (i.e. the offspring) is enhanced, not that of the donor of care (i.e. the parent). Similarly, parents may pay a direct fitness cost of care in terms of their fecundity” (Wong et al., 2013).

The issue of parent-offspring conflict has been analysed in detail by Alexander (1974). He argued that any behavior by the parent which can prevent offspring selfishness will be selected for. But in addition, he and also Williams (1966b) have suggested that this kind of selfish behavior cannot even spread because although the offspring may temporarily gain by being selfish as a juvenile, this gain will be more than offset by the loss in reproduction incurred when the juvenile becomes a parent. This loss in reproduction is due to the fact that as a parent the selfish individual can expect to produce broods with an increased proportion of selfish members. Thus, even though as a juvenile a selfish individual may gain, as a parent it will on average produce fewer descendents such that in the long run it loses. Hence, selfish behavior will be selected against (Blick, 1977).

In resource-limited habitats (the vast majority of habitats; Heininger, 2012), quality of offspring is more important than quantity and individuals that invest into more competitive individuals are selectively favored (Heininger, 2013). Extended parental investment is an important factor that has been suggested to facilitate family formation (Brown, 1987; Ekman et al., 2001a; Ekman, 2006) and increases the reproductive success of parents and survival of offspring, both direct fitness benefits (Royle et al., 2012). For many species of birds and mammals, much of the variance in lifetime reproductive success among females can be attributed to differences in the survival of their offspring until recruitment (Clutton-Brock, 1988).

3. Another quite popular rescue strategy for a theory under pressure is the advancement of highly speculative arguments that are backed by no evidence whatsoever. The first to speculate on a tag for altruism was Hamilton (1964). He conceived what he called a supergene, able to produce (i) a distinctive phenotypic trait, (ii) the faculty to recognize the trait in others, and (iii) the propensity to direct benefits toward bearers of that trait, even though this entails a fitness cost. Soon afterwards, Dawkins (1976) described Hamilton’s thought experiment by using as phenotypic trait the catchy example of a greenbeard. The supergene was now termed ‘‘greenbeard gene,’’ in part to acknowledge its inherent unlikelihood. ‘‘Too good to be true,’’ were Dawkins’ words: for the gene would have to be able to program for 3 effects, namely the feature, its recognition, and the altruistic propensity. The greenbeard argument has been used as an argumentative red herring. The ‘greenbeard effect’ has often been dismissed as implausible because it is unlikely that a single gene can code for altruism and a recognizable tag (Hamilton, 1964; Dawkins, 1976; Blaustein, 1983; West & Gardner, 2010). The second reason greenbeard traits have been thought implausible is that they can be “cheated”. In a population of greenbeard altruists, a mutant capable of producing a greenbeard effect but without the costly tendency to be altruistic, would spread (Guilford, 1988; Okasha, 2002; Henrich, 2004; Gardner & West, 2010). Consequently, this model works only if we allow a mutational constraint. Moreover, Henrich (2004) argued that the common practice should be unacceptable of simply assuming that individuals know the strategies of other individuals and assort accordingly, without providing any justification (theoretical or empirical) as to why this should be so (e.g., Wilson & Dugatkin, 1997). This is tantamount to assuming the answer (Henrich, 2004).

According to a suggestion already made by Fisher (1918) most complex traits such as social behavior are products of a large number of loci with individually small effects (Lynch & Lande, 1993). This premise has stood up to a substantial body of empirical work (Falconer, 1989). For example, mutations generally have small effects on social tendencies in animals, and the resulting effects of these mutations on social tendencies have been considered to be too small for selection to act strongly (Sinervo & Lively, 1996; Sinervo & Calsbeek, 2006; Sinervo et al., 2006; Ross-Gillespie et al., 2007; Wild & Traulsen, 2007). In budding yeast, non-sexual cell aggregation has a polygenic molecular architecture (Li et al., 2013). The complex genetic programming of cooperation even in a relatively “primitive” amoeba (Santorelli et al., 2008) (providing manifold opportunities to cheat) implies that any genetic kin recognition system would require a highly complex network to ensure cooperation with kin. Recent evidence shows that the distribution of allelic effects of quantitative traits is exponential (Mackay, 2001). A few loci with large effects (major genes) influence most of the genetic variation and an increasingly large number of loci with increasingly smaller effects (minor genes) influence the remaining variation. Numerous genes with large effects on behavior have been identified by mutation (Sokolowski, 2001), Mendelian analysis of behavioral variants (Sokolowski, 2001), QTL mapping (Anholt & Mackay, 2004) and the identification of differences in RNA or protein expression between behavioral variants (Insel & Young, 2000). These are all candidate genes for natural variation in behavior (Fitzpatrick et al., 2005).

Repeatedly, the demonstration of greenbeard genes has been claimed (Haig, 1996; Keller & Ross, 1998; Krieger & Ross, 2002; Queller et al., 2003; Summers & Crespi, 2005; Sinervo et al., 2006; Smukalla et al., 2008; West & Gardner, 2010). Taking into account the polygenic nature of cell adhesion systems, the selective data collection results in an observation selection, sensu Bostrom (2003, see chapter 2). In fact, a variety of the reported “greenbeard effects” are dependent on polygenic prokaryotic and eukaryotic cell adhesion and self/non-self recognition systems (Haig, 1996; Queller et al., 2003; Summers & Crespi, 2005; Smukalla et al., 2008). Another one, reported by Keller, Krieger and Ross (1998, 2002), rather than controlling an altruistic greenbeard effect, is a locus that encodes a pheromone-binding protein and may be closely involved in maintaining monogyny and polygyny, and genetic separation of subpopulations (Grafen, 1998). That a highly speculative argument such as the greenbeard has been widely discussed (as of 20-02-2015 according to Google Scholar 1600 citations) while more plausible alternative models have been almost ignored (e.g. 150 citations for Michener & Brothers [1974]) offers eloquent testimony to the biased scientific climate in which the kin selection theory flourishes.

4. Aldous Huxley is said to have stated (Preston, 1981): “The tragedy of science is that frequently a beautiful hypothesis is slain by an ugly fact.” A huge amount of “ugly” data so far was not able to “slain” Hamilton’s rule. The adherents of this “beautiful hypothesis” go at lengths to rescue their favorite topic of scientific investigation, an undertaking that results in twisted logic and misrepresentation.

In its last consequence, kin selection theory is a fascistoid concept, emphasizing the cohesive value of genetic homogeneity for a population tied together by parochial altruism. Nature holds many examples of kin-structured societies that solve their conflicts in an Orwellian society by policing and oppression (Whitfield, 2002, see chapter 9). “Worker policing is a mechanism by which a society resolves its conflicts,” says Ratnieks. “I think it’s the best example of conflict resolution in nature.” (Whitfield, 2002). Parochialism results in hostility toward individuals not of one’s kin or own ethnic, racial, or other group (Choi & Bowles, 2007). Parochial altruism promotes group conflict and may coevolve with warfare (Choi & Bowles, 2007). Consistently, it has been shown that kin and tag-based selection are a theoretical basis for parochialism, war, and other attempts to damage nonkin, because those behaviors lead to enhanced survival of kin (Hamilton, 1964; Bernhard et al., 2006; Choi & Bowles, 2007; Efferson et al., 2008). On the other hand, mutualism causes partners to become increasingly dependent on each other as a basis for peaceful coexistence in societies (Clutton-Brock, 2002). In promoting peaceful coexistence, mutualism would be antagonized by kin selection (Zahavi, 1995; Clutton-Brock, 2002).

Andrew Bourke’s (2011a) rhetorically intended statement “If the analysis of these authors [i.e., Nowak et al. 2010, KH] is correct, inclusive fitness theory has been a decades-long distraction in the field that is theoretically unsound, unnecessarily focused on genetic relatedness and poorly supported by the empirical evidence” has been revealed in this work as a factual statement. It hit the nail on the head.

Occam’s razor, one of the fundamental tenets of modern science, was formulated in the late Middle Ages and states that “Entities should not be multiplied beyond necessity”. According to Sir William of Occam's notion of “parsimony of explanations”, given two models, the simpler one should be preferred because simplicity is desirable in itself. In fact, the focus on genetic relatedness is unnecessary to explain the evolution of cooperation in stochastic environments and should be abandoned altogether.

21. Does altruism exist?


Summary
The claim that altruism does not exist has a long tradition in philosophical political, economic and biological thought. In humans, the most widely assumed autonomous motivation for altruism is thought to be empathy, which has also been proposed for other mammals. Behind the “veil of ignorance”, the fair lottery-social insurance interplay may be the common denominator that regulates such diverse phenomena as stochastic cell fate decisions and acts of “altruistic” sacrifice. Oxytocin may be the neuro-endocrinological agent linking altruism and psychosomatic health.In this way, altruism might have an intrinsic positive component, calling for reconsideration of the emphasis placed on its cost.

In sociobiology, altruism is defined (West et al., 2007c): (i) with respect to the lifetime consequences of a behavior; (ii) on absolute fitness effects (i.e. does it increase or decrease the actor’s fitness, and not relative to just some subset of the population). A fundamental flaw in the sociobiological definition of altruism is that it does not define how the fitness effects of the alleged altruistic act occur. This leads to strange concepts such as “enforced altruism” and “coin-flipping altruism” (see chapter 20.1). On the other hand, if a cooperative behavior was costly in the short term, but provided some long-term (future) benefit, which outweighed that, it would be mutually beneficial and not altruistic. Determining whether and how a cooperative behavior provides short- or long-term direct fitness benefits remains a major problem (Clutton-Brock, 2002; Griffin & West, 2002). It is presumably for this reason that Hamilton (1996, p. 263) thought that reciprocal altruism was misnamed, and that he and others have used alternative terms such as ‘reciprocity’ (Alexander, 1974), or ‘reciprocal cooperation’ (Axelrod & Hamilton, 1981). The literature often attributes non-human primate altruism and cooperation to kin selection, thus calling human cooperation with non-relatives a ‘huge anomaly’ in the animal kingdom (Fehr & Fischbacher, 2003; Gintis et al., 2003; Boyd, 2006; Melis & Semmann, 2010).

The claim that altruism does not exist has a long tradition in philosophical political, economic and biological thought (Wilson, 2015). Those who challenge the existence of altruism do not deny that there are seemingly altruistic acts but question whether they are based upon altruistic motives (Wilson, 2015). In humans, the most widely assumed autonomous motivation for altruism is thought to be empathy (Batson et al., 1988, 2002; Batson, 1991, 2012; Batson & Shaw, 1991a; Batson & Moran, 1999; Van Lange, 2008; Rumble et al., 2010; Lozada et al., 2011; Silk & House, 2011), which has also been proposed for other mammals (de Waal, 1996, 2008). Empathy is the capacity to (i) be affected by and share the emotional state of another (e.g. emotional contagion), (ii) assess the reasons for the other’s state and/or (iii) identify with the other, adopting his or her perspective (Eisenberg & Fabes, 1998; Hinde, 2002; de Waal, 2008; de Waal & Suchak, 2010). Research indicates that empathy is a multicomponent process that includes affect sharing, cognitive perspective taking, and cognitive appraisal (Decety & Jackson, 2004; Lamm et al., 2007; Hein & Singer, 2008; Olsson & Ochsner, 2008). Experiments where empathic concern was induced showed that high empathic conditions increased altruistic responses (Van Lange, 2008). Studies suggested that the experience of empathy has been shown to motivate prosocial behaviors, such as volunteering and donating to charities (Brooks, 2006; Dovidio et al., 2006). Willingness to help others has been correlated with brain activation patterns similar to those activated during empathic states (e.g. Singer & Lamm, 2009; Lutz et al., 2008). Empathy is biased the way one would predict from evolutionary theories of cooperation (i.e. by kinship, social closeness and reciprocation) (Batson & Shaw, 1991b; de Waal & Suchak, 2010).

A simple form of empathy—not retaliating after being punished for involuntary defection—is a prerequisite for evolutionarily stable cooperation. Furthermore, the stability of this, empathic retaliator, strategy increases with the number of opportunities for cooperative exchanges in the life of an average individual (Fishman, 2006). Not all altruistic behavior requires empathy, though. When animals alert others to an outside threat, sacrifice themselves by stinging an intruder or vocally attract others to discovered food, biologists may speak of altruism or cooperation, yet such behavior is unlikely to be based on empathy with the beneficiary. The traditional example of the dangerous rescue of another’s child seems to go beyond cooperation to demand strong altruism, that is, valuing another’s welfare. However, even this stronger sort of altruism alone will not produce stable cooperation in social dilemmas; reciprocity is also necessary (Levine, 1998; Sethi & Somanathan, 2001; Danielson, 2002). Indeed, these behaviors are probably inborn responses to certain stimuli performed with little consideration for the exact situation of the recipients. The fair lottery-social insurance interplay may be the common denominator that regulates such diverse phenomena as stochastic cell fate decisions (see chapter 5.4) and acts of “altruistic” sacrifice. The role of empathy is limited to so-called directed altruism, defined as helping or comforting behavior directed at an individual in need, pain or distress. It has been proposed that empathy relies on automatically activated state-matching that produces shared representations and similar emotions (Preston & de Waal, 2002; Decety & Jackson, 2006). Empathy may promote prosocial behavior, whereas a sense of fairness may play a role as a stabilizer, but not as a promoter of prosocial behavior in non-human animals. Yamamoto and Takimoto (2012) argued that prosocial behavior motivated by sympathetic concerns can survive only with a sense of fairness, the inhibitory system for unnecessarily excessive expression of prosocial behavior. Without a sense of fairness, empathic animals might be exploited by free-riders, which might lead to the extinction of cooperation. Therefore, the interplay of a sense of fairness and empathy are both important to maintaining prosocial behavior and cooperation (Silk & House, 2011). However, these concepts do not seem to account for the pro-social behavior observed when players know that reputation building is not possible (Hilbe & Sigmund, 2010). The evidence that altruism and social support can be beneficial for health (Brown et al., 2005; Cohen & Janicki-Deverts, 2009; Lozada et al., 2011; Thoits, 2011) could partly explain its frequent incidence among unrelated individuals and in situations not involving reciprocity. In this way, altruism might have an intrinsic positive component, calling for reconsideration of the emphasis placed on its cost.

Oxytocin promotes emotional and cognitive aspects of empathy, by exogenous administration as well as at the gene level (Domes et al., 2007b; Rodrigues et al., 2009; DeDreu, 2012; Wu et al., 2012). Empathy towards strangers triggers oxytocin release and subsequent generosity (Barraza & Zak, 2009). On the other hand, consistent with a potential to disrupt social collaboration (see chapter 12), administering exogenous testosterone decreases a variety of empathic processes (Hermans et al., 2006; van Honk et al., 2011; Ronay & Carney, 2013). Oxytocin has potent and long-term physiological antistress effects (Uvnäs-Moberg, 1998) and has been implicated in both immune competence, mental and somatic health, and well-being (Uvnäs-Moberg, 1998; Neumann, 2009; Tom & Assinder, 2010; IsHak et al., 2011; Norman et al., 2012; Smith & Wang, 2012). Thus, oxytocin may be the common neuro-endocrinological agent linking altruism and psychosomatic health.

22. Conclusions


At its 125th anniversary the Science journal asked 125 questions that scientists should have a good shot at answering over the next 25 years, or they should at least know how to go about answering them (Kennedy & Norman, 2005). Questions 93 and 94 were:

93 How Did Cooperative Behavior Evolve?

94 How Will Big Pictures Emerge From a Sea of Biological Data?

I think, with this work I provide some clues how the answers to questions 93 and 94 may look like. The answer to question 93 is given on the preceding pages. I think that the answer to question 94 is implicitly given in my 4 recent papers, including this one. All of these papers (Heininger, 2012, 2013, 2015) that contain a total of more than 18,500 references, deal with evolutionary/ecological conundrums and controversies using a systems biology approach.

As Thomas Kuhn (1970) said, science is puzzle solving. Each of the millions of publications in the biomedical literature can be considered as a piece in the puzzle of the Big Picture. There is a good chance to drown in this sea of biological data. Where could be a lighthouse that may lead the castaway in this huge sea and guide her/him to safe ground? Fortunately, nature provides clues in which direction to move. During its almost 4 billion years, evolution has left traces that can be detected. Darwin's “descent with modification” has left a trail of genetic, physiological and procedural continuity, that can be used like Ariadne’s string for the way out of the labyrinth. It has been advocated previously that a deeper understanding of biological and pathological processes can only be achieved by “unearthing” the “fossil record” of the genome (Runnegar, 1986; Buss, 1987). But this approach, at least to my knowledge, never has been followed rigorously, possibly because it requires the integration of virtually the complete biomedical sciences into a coherent concept. Concerning question 94, my work is based on information retrieved from more than 500,000 papers and a plethora of books from evolutionary biology, ecology, microbiology, epidemiology, paleontology, molecular biology, botany, endocrinology, immunology, and neurobiology that were read over approx. 25 years. Its early stages date back to my occupation with Alzheimer’s disease (Heininger, 1999a, b, 2000a, b), when I realized that metabolically stressed neurons use stress response tools that evolved in microorganisms (Storz & Hengge-Aronis, 2000) and are rooted deep in evolutionary time (Heininger, 2001). A fundamental feature of the “start from the scratch” approach is that it should be an open-ended process, not guided by any preconceived opinion. Previously, holistic concepts have been based on a priori plausibility assumptions. All evolutionary theories were informed by population genetics, the dominant paradigm of evolutionary biology. But population genetics “ is not an empirically sufficient theory” (Lewontin, 1974 p. 267) and failed gloriously to provide consistent and generally accepted answers to the conundrums of evolutionary biology: aging, sexual reproduction, the level of selection, and social behavior. My evidence-based hypotheses to these conundrums and controversies represent a unique, in their scope unprecedented approach that provides revolutionary (sensu Kuhn) paradigm shifts (see Heininger 2012, 2013, 2015, this work). Importantly, these papers sketch a coherent and consistent Big Picture of ecological and evolutionary processes, defining evolution as cybernetic automaton orchestrated by Ross Ashby’s “Law of Requisite Variety”. And, as corollary, give explanations for other biomedical phenomena: the Cambrian explosion, the different bauplans of modular and unitary organisms, Francis Crick’s Central Dogma as closed circle with stochastic environmental-genetic feedback, the relationship between sexual reproduction and cancerogenesis (cancer as evolutionary cost of evolvability), apoptosis as fair lottery, and the rebuttal of Motoo Kimura’s Neutral Theory.

23. Abbreviations


AVP: arginine vasopressin
CA: corpora allata
CBP: conspecific brood parasitism
CRF: corticotropin-releasing factor
DA: dopamine
DIFs: differentiation-inducing factors
GC: glucocorticosteroids
HC: hydrocarbons
HPA: hypothalamic-pituitary-adrenal
IGF: insulin-like growth factor
JH: juvenile hormone
IIS: insulin/IGF-like signaling
MHC: major histocompatibility complex
OT: oxytocin
OTR: oxytocin receptor
PHB: poly-3-hydroxybutyrate
QTL: quantitative trait loci
V1aR: vasopressin 1a receptor

24. References


Abbot P, Withgott JH, Moran NA (2001) Genetic conflict and conditional altruism in social aphid colonies. Proc Natl Acad Sci USA 98: 12068–12071.

Abbot P, Abe J, Alcock J, Alizon S, Alpedrinha, JAC, Andersson M, et al. (2011) Inclusive fitness theory and eusociality. Nature 471: E1–E4.

Abbott DH (1987) Behaviourally mediated suppression of reproduction in female primates. J Zool Lond 213: 455–470.

Abrahamson WG, Anderson SS, McCrea KD (1991) Clonal integration: nutrient sharing between sister ramets of Solidago altissima (Compositae). Am J Bot 78: 1508–1514.

Abrams J, Eickwort GC (1981) Nest switching and guarding by the communal sweat bee Agapostemon virescens (Hymenoptera, Halictidae). Insect Soc 28: 105–116.

Abrams JM (2002) Competition and compensation: coupled to death in development and cancer. Cell 110: 403–406.

Abrams M (2009) The unity of fitness. Philos Sci 76: 750–761.

Abrams PA (1986) Character displacement and niche shift analyzed using consumer-resource models of competition. Theor Popul Biol 29: 107–160.

Abramson G, Kuperman M (2001) Social games in a social network. Phys Rev E 63: 030901(R).

Acher R, Chauvet J, Chauvet MT (1995) Man and the chimaera. Selective versus neutral oxytocin evolution. Adv Exp Med Biol 395: 615–627.

Ackermann M, Stecher B, Freed NE, Songhet P, Hardt WD, Doebeli M (2008) Self-destructive cooperation mediated by phenotypic noise. Nature 454: 987–990.

Adamo SA, Schildberger K, Loher W (1994) The role of the corpora allata in the adult male cricket (Grillus campestris and Gryllus bimaculatus) in the development and expression of its agonistic behavior. J Insect Physiol 40: 439–446.

Adams ES (1991) Nest-mate recognition based on heritable odors in the termite Microcerotermes arboreus. Proc Natl Acad Sci USA 88: 2031–2034.

Adams J, Rothman ED, Kerr WE, Paulino ZL (1977) Estimation of the number of sex alleles and queen matings from diploid male frequencies in a population of Apis mellifera. Genetics 86: 583–596.

Adams MR, Kaplan JR, Koritnik DR (1985) Psychosocial influences on ovarian endocrine and ovulatory function in Macaca fascicularis. Physiol Behav 35: 935–940.

Addicott JF (1985) Competition in mutualistic systems. In: Boucher DH, ed. The biology of mutualism. London, UK: Croom Helm. pp 217–247.

Addicott JF (1996) Cheaters in yucca/moth mutualism. Nature 380: 114–115.

Addicott JF, Bao T (1999) Limiting the costs of mutualism: multiple modes of interaction between yuccas and yucca moths. Proc R Soc Lond B 266: 197–202.

Adkins-Regan E (2005) Hormones and animal social behavior. Princeton, NJ: Princeton University Press.

Adkins-Regan E, Krakauer A (2000) Removal of adult males from the rearing environment increases preference for same-sex partners in the zebra finch. Anim Behav 60: 47–53.

Adl SM, Simpson AG, Farmer MA, Andersen RA, Anderson OR, Barta JR, et al. (2005) The new higher level classification of eukaryotes with emphasis on the taxonomy of protists. J Eukaryot Microbiol 52: 399–451.

Afkhami ME, Rudgers JA, Stachowicz JJ (2014) Multiple mutualist effects: conflict and synergy in multispecies mutualisms. Ecology 95: 833–844.

Agnarsson I (2006) A revision of the new world eximius lineage of Anelosimus (Araneae, Theridiidae) and a phylogenetic analysis using worldwide exemplars. Zool J Linn Soc 146: 453–593.

Agrahari M, Gadagkar R (2003) Juvenile hormone accelerates ovarian development and does not affect age polyethism in the primitively eusocial wasp, Ropalidia marginata. J Insect Physiol 49: 217–222.

Agrell J, Wolff JO, Ylonen H (1998) Counter-strategies to infanticide in mammals: costs and consequences. Oikos 83: 507–517.

Aguirre JD, Marshall DJ (2012) Does genetic diversity reduce sibling competition? Evolution 66: 94–102.

Ahmed A, Karlapalem K (2014) Inequity aversion and the evolution of cooperation. Phys Rev E 89: 022802.

Aikio S, Pakkasmaa S (2003) Relatedness and competitive asymmetry – implications for growth and population dynamics. Oikos 100: 283–290.

Akre RD, Garnett JF, Mac Donald JF, Greene A, Landolt P (1976) Behaviour and colony development of Vespula pensylvanica and V. atropilosa (Hymenoptera: Vespidae). J Kans Entomol Soc 49: 63–84.

Aktipis CA, Fernandez-Duque E (2011) Parental investment without kin recognition: simple conditional rules for parent–offspring behavior. Behav Ecol Sociobiol 65: 1079–1091.

Alaux C, Savarit F, Jaisson P, Hefetz A (2004) Does the queen win it all? Queen–worker conflict over male production in the bumblebee, Bombus terrestris. Naturwissenschaften 91: 400-403.

Alaux C, Boutot M, Jaisson P, Hefetz A (2007) Reproductive plasticity in bumblebee workers (Bombus terrestris)—reversion from fertility to sterility under queen influence. Behav Ecol Sociobiol 62: 213–222.

Alaux C, Sinha S, Hasadsri L, Hunt GJ, Guzmán-Novoa E, DeGrandi-Hoffman G, et al. (2009) Honey bee aggression supports a link between gene regulation and behavioral evolution. Proc Natl Acad Sci USA 106: 15400–15405.

Alberts SC (1999) Paternal kin discrimination in wild baboons. Proc R Soc B Biol Sci 266: 1501–1506.

Alberts SC,Watts HE, Altmann J (2003) Queuing and queue-jumping: long term patterns of reproductive skew in male savannah baboons. Anim Behav 65: 821–840.

Alcantara F, Monk M (1974) Signal propagation during aggregation in the slime mould D. discoideum. J Gen Microbiol 85: 321–334.

Alcock J (2005) Animal behavior: An evolutionary approach. 8th edn. Sunderland, MA: Sinauer Associates.

Alexander RD (1974) The evolution of social behavior. Annu Rev Ecol Syst 5: 325–383.

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, Borgia G (1978) Group selection, altruism, and the levels of organization of life. Annu Rev Ecol Syst 9: 449–474.

Alexander RD, Noonan KM, Crespi BJ (1991) The evolution of eusociality. In: Sherman PW, Jarvis JUM, Alexander RD, eds. The biology of the naked mole rat. Princeton, NJ: Princeton University Press. pp 3–44.

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 (1971) Size and shape. London, UK: Edward Arnold.

Alison MR, Sarraf CE (1992) Apoptosis: a gene-directed programme of cell death. J R Coll Physicians Lond 26: 25–35.

Alizon S, Taylor P (2008) Empty sites can promote altruistic behavior. Evolution 62: 1335–1344.

Almeida O, Gozdowska M, Kulczykowska E, Oliveira RF (2012) Brain levels of arginine-vasotocin and isotocin in dominant and subordinate males of a cichlid fish. Horm Behav 61: 212–217.

Allard RW, Adams J (1969) Population studies in predominantly self-pollinating species. XIII. Intergenotypic competition and competition structure in barley and wheat. Am Nat 103: 621–645.

Allee WC (1931) Animal aggregations. Chicago, IL: University of Chicago Press.

Allee WC, Emerson AE, Park O, Park T, Schmidt KP (1949) Principles of animal ecology. Philadelphia: Saunders.

Allen B, Nowak MA, Wilson EO (2013) Limitations of inclusive fitness. Proc Natl Acad Sci USA 110: 20135–20139.

Allen C, Garmestani A, Havlicek T, Marquet P, Peterson G, Restrepo C, et al. (2006) Patterns in body mass distributions: sifting among alternative hypotheses. Ecol Lett 9: 630–643.

Allen-Arave W, Gurven M, Hill K (2008) Reciprocal altruism, rather than kin selection, maintains nepotistic food transfers on an Ache reservation. Evol Hum Behav 29: 305–318.

Allsopp TE, Fazakerley JK (2000) Altruistic cell suicide and the specialized case of the virus-infected nervous system. Trends Neurosci 23: 284–290.

Alonso W (1998) The role of Kin Selection theory on the explanation of biological altruism: A critical review. J Comp Biol 3: 1–14.

Alonso WJ, Schuck-Paim C (2002) Sex-ratio conflicts, kin selection, and the evolution of altruism. Proc Natl Acad Sci USA 99: 6843–6847.

Alpedrinha J, West SA, Gardner A (2013) Haplodiploidy and the evolution of eusociality: worker reproduction. Am Nat 182: 421–438.

Alpert P (1991) Nitrogen sharing among ramets increases clonal growth in Fragaria chiloensis. Ecology 72: 69–80.

Alpert P (1999) Clonal integration in Fragaria chiloensis differs between populations: ramets from grassland are selfish. Oecologia 120: 69–76.

Alpert P, Holzapfel C, Slominski C (2003) Differences in performance between genotypes of Fragaria chiloensis with different degrees of resource sharing. J Ecol 91: 27–35.

Alpert P, Mooney HA (1986) Resource sharing among ramets in the clonal herb, Fragaria chiloensis. Oecologia 70: 227–233.

Alpert P, Stuefer JF (1997) Division of labour in clonal plants. In: de Kroon H, van Groenendael J, eds. The ecology and evolution of clonal plants. Leiden, the Netherlands: Backhuys Publishers. pp 137–154.

Alt W (1980) Biased random walk models for chemotaxis and related diffusion approximations. J Math Biol 9: 147–177.

Altenberg L (2005) Evolvability suppression to stabilize far-sighted adaptations. Artif Life 11: 427–443.

Altizer S, Nunn CL, Thrall PH, Gittleman JL, Antonovics J, Cunningham AA, et al. (2003) Social organization and parasite risk in mammals: integrating theory and empirical studies. Annu Rev Ecol Evol Syst 34: 517–547.

Altmann SA (1974) Baboons, space, time, and energy. Am Zoologist 14: 221–248.

Altuvia S, Weinstein-Fischer D, Zhang A, Postow L, Storz G (1997) A small, stable RNA induced by oxidative stress: role as a pleiotropic regulator and antimutator. Cell 90: 43–53.

Alvard MS (2003) The adaptive nature of culture. Evol Anthropol 12: 136–149.

Amar KO, Chadwick NE, Rinkevich B (2008) Coral kin aggregations exhibit mixed allogeneic reactions and enhanced fitness during early ontogeny. BMC Evol Biol 8: 126.

Amdam GV, Seehuus SC (2006) Order, disorder, death: lessons from a superorganism. In: Vande Woude GF, Klein G, eds. Advances in cancer research, Vol. 95. San Diego, CA: Academic Press. pp 31–60.

Amdam GV, Csondes A, Fondrk MK, Page RE (2006) Complex social behaviour derived from maternal reproductive traits. Nature 439: 76–78.

Ament SA, Corona M, Pollock HS, Robinson GE (2008) Insulin signaling is involved in the regulation of worker division of labor in honey bee colonies. Proc Natl Acad Sci USA 105: 4226–4231.

Amsalem E, Hefetz A (2010) The appeasement effect of sterility signaling in dominance contests among Bombus terrestris workers. Behav Ecol Sociobiol 64: 1685–1694.

Amsalem E, Teal P, Grozinger CM, Hefetz A (2014) Precocene-I inhibits juvenile hormone biosynthesis, ovarian activation, aggression and alters sterility signal production in bumble bee (Bombus terrestris) workers. J Exp Biol 217: 3178–3185.

Anacker AMJ, Beery AK (2013) Life in groups: the roles of oxytocin in mammalian sociality. Front Behav Neurosci 7: 185.

Andalo C, Goldringer I, Godelle B (2001) Inter- and intragenotypic competition under elevated carbon dioxide in Arabidopsis thaliana. Ecology 82: 157–164.

Anderson AJ, Dawes EA (1990) Occurrence, metabolism, metabolic role, in industrial uses of bacterial polyhydroxyalkanoates. Microbiol Rev 54: 450–472.

Anderson KE, Linksvayer TA, Smith CR (2008) The causes and consequences of genetic caste determination in ants (Hymenoptera: Formicidae). Myrmecol News 11: 119–132.

Andersson M (1984) The evolution of eusociality. Annu Rev Ecol Syst 15: 165–189.

Andras P, Roberts G, Lazarus J (2003) Environmental risk, cooperation and communication complexity. In: Alonso EKD, ed. Adaptive agents and multi-agent systems. Berlin, Germany: Springer-Verlag. pp 49–65.

Andras P, Lazarus, J (2005) Cooperation, risk and the evolution of teamwork. In: Gold N, ed. Teamwork: multi-disciplinary perspectives. Basingstoke, UK: Palgrave Macmillan. pp 56–77.

Andras P, Lazarus J, Roberts G (2007) Environmental adversity and uncertainty favour cooperation. BMC Evol Biol 7: 240.

Anetzberger C, Pirch T, Jung K (2009) Heterogeneity in quorum sensing-regulated bioluminescence of Vibrio harveyi. Mol Microbiol 73: 267–277.

Anetzberger C, Schell U, Jung K (2012) Single cell analysis of Vibrio harveyi uncovers functional heterogeneity in response to quorum sensing signals. BMC Microbiol 12: 209.

Anholt BR (1994) Cannibalism and early instar survival in a larval damselfly. Oecologia 99: 60–65.

Anholt RRH, Mackay TFC (2004) Quantitative genetic analyses of complex behaviours in Drosophila. Nat Rev Genet 5: 838–849.

Anholt RRH, Mackay TFC (2012) Genetics of aggression. Annu Rev Genet 46: 145–164.

Anholt RRH, Mackay TFC (2015) Dissecting the genetic architecture of behavior in Drosophila melanogaster. Curr Opin Behav Sci 2: 1–7.

Annesley SJ, Fisher PR (2009) Dictyostelium discoideum—a model for many reasons. Mol Cell Biochem 329: 73–91.

Anstett MC, Gibernau M, Hossaert-McKey M (1998) Partial avoidance of female inflorescences of a dioecious fig by their mutualistic pollinator wasps. Proc R Soc Lond B 265: 45–50.

Antal T, Ohtsuki H, Wakeley J, Taylor PD, Nowak MA (2009) Evolution of cooperation by phenotypic similarity. Proc Natl Acad Sci USA 106: 8597–8600.

Anthony CD (2003) Kinship influences cannibalism in the wolf spider, Pardosa milvina. J Insect Behav 16: 23–36.

Antolin MF, Strobeck C (1985) The population genetics of somatic mutations in plants. Am Nat 126: 52–62.

Aoki K, Wakano JY, Feldman MW (2005) The emergence of social learning in a temporally changing environment: a theoretical model. Curr Anthropol 46: 334–340.

Aoki S (1977) Colophina clematis (Homoptera, Pemphigidae), an aphid species with soldiers. Kontyu 45: 276–282.

Aoki S, Kurosu U, Stern DL (1991) Aphid soldiers discriminate between soldiers and non-soldiers, rather than between kin and non-kin, in Ceratoglyphina bambusae. Anim Behav 42: 865–866.

Apel K, Hirt H (2004) Reactive oxygen species: metabolism, oxidative stress, and signal transduction. Annu Rev Plant Biol 55: 373–399.

Araki T, Maeda Y (1995) Cell-cycle progression during the development of Dictyostelium discoideum and its relation to the subsequent cell-sorting in multicellular structures. Dev Growth Differ 37: 479–485.

Araki T, Maeda Y (1998) Mutual relation between the cell-cycle progression and prespore differentiation in Dictyostelium development. Zool Sci 15: 77–84.

Araújo MS, Bolnick DI, Layman CA (2011) The ecological causes of individual specialization. Ecol Lett 14: 948–958.

Arbilly M, Motro U, Feldman MW, Lotem A (2010) Co-evolution of learning complexity and social foraging strategies. J Theor Biol 267: 573–581.

Archetti M, Scheuring I (2011) Coexistence of cooperation and defection in public goods games. Evolution 65: 1140–1148.

Archetti M, Scheuring I, Hoffman M, Frederickson ME, Pierce NE, Yu DW (2011) Economic game theory for mutualism and cooperation. Ecol Lett 14: 1300–1312.

Archetti M, Scheuring I (2013) Trading public goods stabilizes interspecific mutualism. J Theor Biol 318: 58–67.

Archie EA, Morrison TA, Foley CA, Moss CJ, Alberts SC (2006a) Dominance rank relationships among wild female African elephants, Loxodonta africana. Anim Behav 71: 117–127.

Archie EA, Moss CJ, Alberts SC (2006b) The ties that bind: genetic relatedness predicts the fission and fusion of social groups in wild African elephants. Proc R Soc Lond B 273: 513–522.

Ardanaz N, Pagano PJ (2006) Hydrogen peroxide as a paracrine vascular mediator: regulation and signaling leading to dysfunction. Exp Biol Med (Maywood) 231: 237–251.

Argyres AZ, Schmitt J (1992) Neighbor relatedness and competitive performance in Impatiens capensis (Balsaminaceae): a test of the resource partitioning hypothesis. Am J Bot 79: 181–185.

Arias E, Cuervo AM (2011) Chaperone-mediated autophagy in protein quality control. Curr Opin Cell Biol 23: 184–189.

Arletti R, Benelli A, Bertolini A (1992) Oxytocin involvement in male and female sexual behavior. Ann NY Acad Sci 652: 180–193.

Armitage KB, Schwartz OA (2000) Social enhancement of fitness in yellow-bellied marmots. Proc Natl Acad Sci USA 97: 12149–12152.

Armitage KB (2007) Evolution of sociability in marmots: it begins with hibernation. In: Wolff JO, Sherman PW, eds. Rodent societies: an ecological and evolutionary perspective. Chicago, IL: University of Chicago Press. pp 356–367.

Armstrong RA, McGehee R (1980) Competitive exclusion. Am Nat 115: 151–170.

Arnold C, Taborsky B (2010) Social experience in early ontogeny has lasting effects on social skills in cooperatively breeding cichlids. Anim Behav 79: 621–630.

Arnold G, Le Conte Y, Trouiller J, Hervet H, Chappe B, Masson C (1994) Inhibition of worker honeybee ovaries development by a mixture of fatty acid esters from larvae. C R Acad Sci Ser III 317: 511–515.

Arnold KE, Owens IPF (1998) Cooperative breeding in birds: a comparative test of the life history hypothesis. Proc R Soc Lond Ser B 265: 739–745.

Arnold KE, Owens IPF (1999) Cooperative breeding in birds: the role of ecology. Behav Ecol 10: 465–471.

Arnold KE, Owens IPF, Goldizen AW (2005) Division of labour within cooperatively breeding groups. Behaviour 142: 1577–1590.

Arnold W (1990a) The evolution of marmot sociality. I. Why disperse late? Behav Ecol Sociobiol 27: 229–237.

Arnold W (1990b) The evolution of marmot sociality. II. Costs and benefits of joint hibernation. Behav Ecol Sociobiol 27: 239–246.

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: 123–127.

Arnqvist G, Edvardsson M, Friberg U, Nilsson T (2000) Sexual conflict promotes speciation in insects. Proc Natl Acad Sci USA 97: 10460–10464.

Arnqvist G, Rowe L (2005) Sexual conflict. Princeton, NJ: Princeton University Press.

Aron S, Pasteels JMM, Deneubourg JLL (1989) Trail laying behavior during exploratory recruitment in the Argentine ant, Iridomyrmex humile (mayr). Biol Behav 14: 207–217.

Arquier N, Vigne P, Duplan E, Hsu T, Therond PP, Frelin C, D’Angelo G (2006) Analysis of the hypoxia-sensing pathway in Drosophila melanogaster. Biochem J 393: 471–480.

Arrow KJ, Cropper M, Gollier C, Groom B, Heal GM, Newell RG, et al. (2012) How should benefits and costs be discounted in an intergenerational context? The views of an expert panel. Washington, DC: Workshop held at Resources for the Future in September 2011.

Asberg M, Scalling D, Trakeman-Bendz I, Wagner A (1987) A psychology of suicide, impulsivity, and related phenomena. In: Meltzer HY, ed. Psychopharmacology: Third generation of progress. New York, NY: Raven Press. pp 655−688.

Askew RR (1971) Parasitic insects. New York, NY: American Elsevier.

Assaneo F, Coutinho RM, Lin Y, Mantilla C, Lutscher F (2013) Dynamics and coexistence in a system with intraguild mutualism. Ecol Complexity 14: 64–74.

Atkinson L, Teschendorf G, Adams ES (2008) Lack of evidence for nepotism by workers tending queens of the polygynous termite Nasutitermes corniger. Behav Ecol Sociobiol 62: 805–812.

Atzmony D, Zahavi A, Nanjundiah V (1997) Altruistic behaviour in Dictyostelium discoideum explained on the basis of individual selection. Curr Sci 72: 142–145.

Austad SN, Rabenold KN (1985) Reproductive enhancement by helpers and an experimental inquiry into its mechanism in the bicoloured wren. Behav Ecol Sociobiol 17: 19–27.

Avery SV (2006) Microbial cell individuality and the underlying sources of heterogeneity. Nat Rev Microbiol 4: 577–587.

Avilés L (1986) Sex ratio bias and possible group selection in the social spider Anelosimus eximius. Am Nat 128: 1–12.

Avilés L (1993) Interdemic selection and the sex ratio: a social spider perspective. Am Nat 142: 320–345.

Avilés L (1997) Causes and consequences of cooperation and permanent sociality in spiders. In: Choe JC, Crespi BJ, eds. The evolution of social behavior in insects and arachnids. Cambridge, UK: Cambridge University Press. pp 476–498.

Avilés L (1999) Cooperation and non-linear dynamics: An ecological perspective on the evolution of sociality. Evol Ecol Res 1: 459–477.

Avilés L (2002) Solving the freeloaders paradox: genetic associations and frequency dependent selection in the evolution of cooperation among nonrelatives. Proc Natl Acad Sci USA 99: 14268–14273.

Avilés L, Tufiño P (1998) Colony size and individual fitness in the social spider Anelosimus eximius. Am Nat 152: 403–418.

Avilés L, Abbot P, Cutter AD (2002) Population ecology, nonlinear dynamics, and social evolution. I. Associations among nonrelatives. Am Nat 159: 115–127.

Avilés L, Fletcher JA, Cutter AD (2004) The kin composition of social groups: trading group size for degree of altruism. Am Nat 164: 132–144.

Aumann RJ, Maschler M (1995) Repeated games with incomplete information. Cambridge, MA: MIT Press.

Axelrod R (1984) The evolution of cooperation New York, NY: Basic Books.

Axelrod R, Hamilton WD (1981) The evolution of cooperation. Science 211: 1390–1396.

Ayalew L, Murphy BE (1986) In vitro demonstration of in utero larval development in an oviparous parasitic nematode: Haemonchus contortus. Parasitology 93: 371–381.

Ayroles JF, Carbone MA, Stone EA, Jordan KW, Lyman RF, Magwire MM, et al. (2009) Systems genetics of complex traits in Drosophila melanogaster. Nat Genet 41: 299–307.

Azevedo SV, Hartfelder K (2008) The insulin signaling pathway in honey bee (Apis mellifera) caste development—differential expression of insulin-like peptides and insulin receptors in queen and worker larvae. J Insect Physiol 54: 1064–1071.

Azevedo SV, Caranton OAM, de Oliveira TL, Hartfelder K (2011) Differential expression of hypoxia pathway genes in honey bee (Apis mellifera L.) caste development. J Insect Physiol 57: 38–45.

Baba T, Schneewind O (1996) Target cell specificity of a bacteriocin molecule: a C-terminal signal directs lysostaphin to the cell wall of Staphylococcus aureus. EMBO J 15: 4789–4797.

Baba T, Schneewind O (1998) Instruments of microbial warfare: bacteriocin synthesis, toxicity and immunity. Trends Microbiol 6: 66–71.

Babis M, Holman L, Fenske R, Thomas ML, Baer B (2014) Cuticular lipids correlate with age and insemination status in queen honeybees. Insect Soc 61: 337–345.

Bachmann H, Fischlechner M, Rabbers I, Barfa N, dos Santos FB, Molenaar D, Teusink B (2013) Availability of public goods shapes the evolution of competing metabolic strategies. Proc Natl Acad Sci USA 110: 14302–14307.

Backx AG, Guzmán-Novoa E, Thompson GJ (2012) Factors affecting ovary activation in honey bee workers: a meta-analysis. Insect Soc 59: 381–388.

Badyaev AV, Walsh JB (2014) Epigenetic processes and genetic architecture in character origination and evolution. In: Charmantier A, Garant D, Kruuk LEB, eds. Quantitative Genetics in the Wild. Oxford, UK: Oxford University Press. pp 177–189.

Baer B, Schmid-Hempel P (1999) Experimental variation in polyandry affects parasite loads and fitness in a bumble-bee. Nature 397: 151–154.

Baffy G, Loscalzo J (2014) Complexity and network dynamics in physiological adaptation: An integrated view. Physiol Behav 131: 49–56.

Baglione V, Marcos JM, Canestrari D, Ekman J (2002) Direct fitness benefits of group living in a complex cooperative society of carrion crows, Corvus corone corone. Anim Behav 64: 887–893.

Baglione V, Canestrari D, Marcos JM, Ekman J (2003) Kin selection in cooperative alliances of carrion crows. Science 300: 1947–1949.

Baglione V, Canestrari D, Marcos JM, Ekman J (2006) Experimentally increased food resources in the natal territory promote offspring philopatry and helping in cooperatively breeding carrion crows. Proc R Soc B Biol Sci 273: 1529–1535.

Bagnères AG, Lorenzi MC, Dusticier G, Turillazzi S, Clément JL (1996) Chemical usurpation of a nest by paper wasp parasites. Science 272: 889–892.

Bagnères AG, Lorenzi MC (2010) Chemical deception/mimicry using cuticular hydrocarbons. In: Blomquist GJ, Bagnères AG, eds. Insect hydrocarbons: biology, biochemistry and chemical ecology. Cambridge, UK: Cambridge University Press. pp 282–324.

Bagowski CP, Ferrell J, James E (2001) Bistability in the JNK cascade. Curr Biol 11: 1176–1182.

Bagowski CP, Besser J, Frey CR, Ferrell JE Jr (2003) The JNK cascade as a biochemical switch in mammalian cells: ultrasensitive and all-or-none responses. Curr Biol 13: 315–320.

Baier A, Wittek B, Brembs B (2002) Drosophila as a new model organism for neurobiology of aggression? J Exp Biol 205: 1233–1240.

Baker AC (2001) Reef corals bleach to survive change. Nature 411: 765–766.

Baker NE (2011) Cell competition. Curr Biol 21: R11–R15.

Baker PJ, Funk SM, Harris S, White PCL (2000) Flexible spatial organization of urban foxes, Vulpes vulpes, before and during an outbreak of sarcoptic mange. Anim Behav 59: 127–146.

Balas MT, Adams ES (1996) The dissolution of cooperative groups: mechanisms of queen mortality in incipient fire ant colonies. Behav Ecol Sociobiol 38: 391–399.

Balázsi G, van Oudenaarden A, Collins JJ (2011) Cellular decision making and biological noise: from microbes to mammals. Cell 144: 910–925.

Baldauf SL (2003) The deep roots of eukaryotes. Science 300: 1703–1706.

Bales KL, Perkeybile AM (2012) Developmental experiences and the oxytocin receptor system. Horm Behav 61: 313–319.

Baliadi Y, Yoshiga T, Kondo E (2001) Development of endotokia matricida and emergence of originating infective juveniles of steinernematid and heterorhabditid nematodes. Jpn J Nematol 31: 26–35.

Ballinger PT, Wilcox NT (1997) Decisions, error and heterogeneity. Econ J 107: 1090–1005.

Bamford CV, d'Mello A, Nobbs AH, Dutton LC, Vickerman MM, Jenkinson HF (2009) Streptococcus gordonii modulates Candida albicans biofilm formation through intergeneric communication. Infect Immun 77: 3696–3704.

Banin E, Vasil ML, Greenberg EP (2005) Iron and Pseudomonas aeruginosa biofilm formation. Proc Natl Acad Sci USA 102: 11076–11081.

Barabasi AL, Gulbahce N, Loscalzo J (2011) Network medicine: a network-based approach to human disease. Nat Rev Genet 12: 56–68.

Baranger M (2000) Chaos, complexity, and entropy: a physics talk for non-physicists. Cambridge, MA: New England Complex Systems Institute.

Baras E (1999) Sibling cannibalism among juvenile vundu under controlled conditions. I. Cannibalistic behaviour, prey selection and prey size selectivity. J Fish Biol 54: 82–105.

Barbieri M, Bonafe M, Franceschi C, Paolisso G (2003) Insulin/IGF-I-signaling pathway: an evolutionarily conserved mechanism of longevity from yeast to humans. Am J Physiol Endocrinol Metab 285: 1064–1071.

Barnard CJ (1984) Producers and scroungers: strategies of exploitation and parasitism. New York, NY: Chapman and Hall.

Barnard CJ (1990) Kin recognition: problems, prospects, and the evolution of discrimination systems. Adv Study Behav 19: 29–81.

Barnard CJ (1991) Kinship and social behaviour: the trouble with relatives. Trends Ecol Evol 6: 310–311.

Barnard C (1999) Still having trouble with relatives. Trends Ecol Evol 14: 448.

Barnard CJ, Burk TE (1979) Dominance hierarchies and the evolution of ‘individual recognition’. JTheor Biol 81: 65–73.

Barnard CJ, Sibly RM (1981) Producers and scroungers–a general model and its application to captive flocks of house sparrows. Anim Behav 29: 543–550.

Barnard CJ, Hurst JL, Aldhous PGM (1991) Of mice and kin: the functional significance of kin bias in social behaviour. Biol Rev 66: 379–430.

Barnett OW, Fagan RW, Booker JM (1991) Hostility and stress as mediators of aggression in violent men. J Fam Violence 6: 217–241.

Barocas A, Ilany A, Koren L, Kam M, Geffen E (2011) Variance in centrality within rock hyrax social networks predicts adult longevity. PLoS ONE 6: e22375.

Barraza J, Zak PJ (2009) Empathy towards strangers triggers oxytocin release and subsequent generosity. Ann NY Acad Sci 1167: 182–189.

Barrett J, Abbott DH, George LM (1993) Sensory cues and the suppression of reproduction in subordinate female marmoset monkeys, Callithrix jacchus. J Reprod Fertil 97: 301–310.

Barrett L, Henzi SP, Weingrill T, Lycett JE, Hill RA (1999) Market forces predict grooming reciprocity in female baboons. Proc R Soc B 266: 665–670.

Barrick JE, Yu DS, Yoon SH, Jeong H, Oh TK, Schneider D, et al. (2009) Genome evolution and adaptation in a long-term experiment with Escherichia coli. Nature 461: 1243–1247.

Barron AB, Oldroyd BP (2001) Social regulation of ovary activation in’anarchistic’honey-bees (Apis mellifera). Behav Ecol Sociobiol 49: 214–219.

Barron AB, Oldroyd BP, Ratnieks FLW (2001) Worker reproduction in honey-bees (Apis) and the anarchic syndrome: a review. Behav Ecol Sociobiol 50: 199–208.

Barron AB, Robinson GE (2008) The utility of behavioral models and modules in molecular analyses of social behavior. Genes Brain Behav 7: 257–265.

Bar-Tal D (1976) Prosocial behavior: Theory and research. New York, NY: Wiley.

Bartlett J (1987) Filial cannibalism in burying beetles. Behav Ecol Sociobiol 21: 179–183.

Bartlett MS (1960) Stochastic population models. London, UK: Methuen.

Barton HA, Johnson Z, Cox CD, Vasil AI, Vasil ML (1996) Ferric uptake regulator mutants of Pseudomonas aeruginosa with distinct alterations in the iron-dependent repression of exotoxin A and siderophores in aerobic and microaerobic environments. Mol Microbiol 21: 1001–1017.

Barton NH (1990) Pleiotropic models of quantitative variation. Genetics 124: 773–782.

Barton NH, Post RJ (1986) Sibling competition and the advantage of mixed families. J Theor Biol 120: 381–387.

Bartz SH, Hölldobler B (1982) Colony founding in Myrmecocystus mimicus Wheeler (Hymenoptera: Formicidae) and the evolution of foundress associations. Behav Ecol Sociobiol 10: 137–147.

Bascompte J, Jordano P, Olesen JM (2006) Asymmetric coevolutionary networks facilitate biodiversity maintenance. Science 312: 431–433.

Bassler BL, Losick R (2006) Bacterially speaking. Cell 125: 237–246.

Bastian ML, Sponberg AC, Suomi SJ, Higley JD (2003) Long-term effects of infant rearing condition on the acquisition of dominance rank in juvenile and adult rhesus macaques (Macaca mulatta). Dev Psychobiol 42: 44–51.

Bastolla U, Fortuna MA, Pascual-García A, Ferrera A, Luque B, Bascompte J (2009) The architecture of mutualistic networks minimizes competition and increases biodiversity. Nature 458: 1018–1020.

Batra SWT (1966) Nests and social behaviour of halictine bees of India. J Entomol 28: 375–393

Batson CD (1991) The altruism question: toward a social-psychological answer. Hillsdale, NJ: Erlbaum.

Batson CD (2012) The empathy-altruism hypothesis. In: Decety J, ed. Empathy: from bench to bedside. Cambridge, MA: MIT Press. pp 41–54.

Batson CD, Dyck JL, Brandt JR, Batson JG, Powell AL, McMaster MR, Griffitt C (1988) Five studies testing two new egoistic alternatives to the empathy-altruism hypothesis. J Pers Soc Psychol 55: 52–77.

Batson CD, Shaw LL (1991a) Evidence for altruism: toward a pluralism of prosocial motives. Psychol Inq 2: 107–122.

Batson CD, Shaw LL (1991b) Encouraging words concerning the evidence for altruism. Psychol Inq 2: 159–168.

Batson CD, Moran T (1999) Empathy-induced altruism in a Prisoner’s Dilemma. Eur J Soc Psychol 29: 909–924.

Batson CD, Ahmad N, Lishner DA, Tsang J (2009) Empathy and altruism. In: Snyder CR, Lopez SJ, eds. Oxford handbook of positive psychology, 2nd edn. Oxford, UK: Oxford University Press. pp 417–426.

Battesti A, Majdalani N, Gottesman S (2011) The RpoS-mediated general stress response in Escherichia coli. Annu Rev Microbiol 65: 189–213.

Baumgartner T, Heinrichs M, Vonlanthen A, Fischbacher U, Fehr E (2008) Oxytocin shapes the neural circuitry of trust and trust adaptation in humans. Neuron 58: 639–650.

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.

Beauchamp G (2010) Determinants of false alarms in staging flocks of semipalmated sandpipers. Behav Ecol 21: 584–587.

Beauchamp G (2013) Social predation: how group living benefits predators and prey. New York, NY: Academic Press.

Beauchamp G, Ruxton GD (2007) False alarms and the evolution of antipredator vigilance. Anim Behav 74: 1199–1206.

Beauchamp GK, Yamazaki K, Boyse EA (1985) The chemosensory recognition of genetic individuality. Sci Am 253: 86–92.

Beauchamp GK, Yamazaki K (2003) Chemical signalling in mice. Biochem Soc Trans 31: 147–151.

Beckert J, Lutter M (2013) Why the poor play the lottery: sociological approaches to explaining class-based lottery play. Sociology 47: 1152–1170.

Bednarz JC (1988) Cooperative hunting Harris’ hawks (Parabuteo unicinctus). Science 239: 1525–1527.

Bednekoff PA (1996) Risk-sensitive foraging, fitness, and life histories: where does reproduction fit into the big picture? Am Zool 36: 471–483.

Beecher MD (1991) Success and failures of parent-offspring recognition in animals. In: Hepper PG, ed. Kin Recognition. Cambridge, UK: Cambridge University Press. pp 95–124.

Beekman M, Ratnieks FLW (2003) Power over reproduction in the social Hymenoptera. Phil Trans R Soc Lond B 358: 1741–1753.

Beekman M, Oldroyd BP (2003) Different policing rates of eggs laid by queenright and queenless anarchistic honey-bee workers (Apis mellifera L.). Behav Ecol Sociobiol 54: 480–484.

Beekman M, Komdeur J, Ratnieks FL (2003) Reproductive conflicts in social animals: who has power? Trends Ecol Evol 18: 277–282.

Beekman M, Oldroyd BP (2008) When workers disunite: intraspecific parasitism by eusocial bees. Annu Rev Entomol 53: 19–37.

Be'er A, Zhang HP, Florin EL, Payne SM, Ben-Jacob E, Swinney HL (2009) Deadly competition between sibling bacterial colonies. Proc Natl Acad Sci USA 106: 428–433.

Be'er A, Ariel G, Kalisman O, Helman Y, Sirota-Madi A, Zhang HP, Florin EL, Payne SM, Ben-Jacob E, Swinney HL (2010) Lethal protein produced in response to competition between sibling bacterial colonies. Proc Natl Acad Sci USA 107: 6258–6263.

Beggs KT, Glendining KA, Marechal NM, Vergoz V, Nakamura I, Slessor KN, Mercer AR (2007) Queen pheromone modulates brain dopamine function in worker honey bees. Proc Natl Acad Sci USA 104: 2460–2464.

Beggs KT, Mercer AR (2009) Dopamine receptor activation by honey bee queen pheromone. Curr Biol 19: 1206–1209.

Bejerano-Sagie M, Xavier KB (2007) The role of small RNAs in quorum sensing. Curr Opin Microbiol 10: 189–198.

Bekoff M, Wells MC (1982) The behavioral ecology of coyotes: social organization, rearing patterns, space use, and resource defense. Z Tierpsychol 60: 281–305.

Bell G (1985) Two theories of sex and variation. Experientia 41: 1235–1245.

Bell G, Koufopanou V (1991) The architecture of the life cycle in small organisms. Phil Trans R Soc Lond Ser B 332: 81–89.

Bell MBV (2007) Cooperative begging in banded mongoose pups. Curr Biol 17: 717–721.

Bell MBV, Cant MA, Borgeaud C, Thavarajah N, Samson J, Clutton-Brock TH (2014) Suppressing subordinate reproduction provides benefits to dominants in cooperative societies of meerkats. Nat Commun 5: 4499.

Beloin C, Ghigo JM (2005) Finding gene-expression patterns in bacterial biofilms. Trends Microbiol 13: 16–19.

Beloin C, Roux A, Ghigo JM (2008) Escherichia coli biofilms. Curr Top Microbiol Immunol 322: 249–289.

Belozertseva IV, Bespalov AY (1999) Effects of NMDA receptor channel blockade on aggression in isolated male mice. Aggressive Behav 25: 381–396.

Benabentos R, Hirose S, Sucgang R, Curk T, Katoh M, Ostrowski EA, et al. (2009) Polymorphic members of the lag gene family mediate kin discrimination in Dictyostelium. Curr Biol 19: 567–572.

Bendesky A, Bargmann CI (2011) Genetic contributions to behavioural diversity at the gene–environment interface. Nat Rev Genet 12: 809–820.

Bendor J, Kramer RM, Stout S (1991) When in doubt. Cooperation in a noisy prisoner’s dilemma. J Conflict Resolut 35: 691–719.

Ben-Jacob E (2003) Bacterial self-organization: Co-enhancement of complexification and adaptability in a dynamic environment. Phil Trans R Soc London Ser A 361: 1283–1312.

Ben-Jacob E (2008) Social behavior of bacteria: from physics to complex organization. The European Physical Journal B 65: 315–322.

Ben-Jacob E (2009) Learning from bacteria about natural information processing. Ann NY Acad Sci 1178: 78–90.

Ben-Jacob E, Schochet O, Tenenbaum A, Cohen I, Czirók A, Vicsek T (1994) Generic modelling of cooperative growth patterns in bacterial colonies. Nature 368: 46–49.

Ben-Jacob E, Cohen I, Levine H (2000) Cooperative self-organization of microorganism. Adv Phys 49: 395–554.

Ben-Jacob E, Becker I, Shapira Y, Levine H (2004) Bacterial linguistic communication and social intelligence. Trends Microbiol 12: 366–372.

Ben-Jacob E, Shapira Y, Tauber A (2006) Seeking the foundations of cognition in bacteria. Physica A 359: 495–524.

Ben-Jacob E, Schultz D (2010) Bacteria determine fate by playing dice with controlled odds. Proc Natl Acad Sci USA 107: 13197–13198.

Benmayor R, Buckling A, Bonsall MB, Brockhurst MA, Hodgson DJ (2008) The interactive effects of parasites, disturbance, and productivity on experimental adaptive radiations. Evolution 62: 467–477.

Bennett B (1989) Nestmate recognition systems in a monogynous-polygynous species pair of ants (Hymenoptera: Formicidae). I. Worker and queen derived cues. Sociobiology 16: 121–139.

Bennett NC, Faulkes CG, Molteno AJ (1996) Reproductive suppression in subordinate, non-breeding female Damaraland mole-rats: two components to a lifetime of socially induced infertility. Proc R Soc Lond Ser B Biol Sci 263: 1599–1603.

Benskin CMH, Mann NI, Lachlan RF, Slater PJB (2002) Social learning directs feeding preferences in the zebra finch, Taeniopygia guttata. Anim Behav 64: 823–828.

Bentley PJ (2009) Methods for improving simulations of biological systems: systemic computation and fractal proteins. J R Soc Interface 6(Suppl 4): S451–S466.

Benton TG, Grant A (2000) Evolutionary fitness in ecology: comparing measures of fitness in stochastic, density-dependent environments. Evol Ecol Res 2: 769–789.

Berec L, Boukal DS, Berec M (2001) Linking the Allee effect, sexual reproduction, and temperature-dependent sex determination via spatial dynamics. Am Nat 157: 217–230.

Berg A, Lindberg T, Källebrink KG (1992) Hatching success of lapwings on farmland: Differences between habitats and colonies of different sizes. J Anim Ecol 61: 469–476.

Berg KH, Biørnstad TJ, Johnsborg O, Håvarstein LS (2012) Properties and biological role of streptococcal fratricins. Appl Environ Microbiol 78: 3515–3522.

Berger J (1987) Reproductive fates of dispersers in a harem-dwelling ungulate: the wild horse. In: Chepko-Sade BD, Halpin ZT, eds. Mammalian dispersal patterns. Chicago, IL: University of Chicago Press. pp 41–54.

Bergman A, Feldman MW (1995) On the evolution of learning: representation of a stochastic environment. Theor Popul Biol 48: 251–276.

Bergmüller R, Johnstone RA, Russell AF, Bshary R (2007) Integrating cooperative breeding into theoretical concepts of cooperation. Behav Process 76: 61–72.

Bergstrom BT, Lachmann M (2003) The red king effect: when the slowest runner wins the coevolutionary race. Proc Natl Acad Sci USA 100: 593–598.

Bergstrom TC (2002) Evolution of social behavior: individual and group selection. J Econ Perspect 16: 67–88.

Bernasconi G, Keller L (1996) Reproductive conflicts in cooperative associations of fire ant queens. Proc R Soc Lond B 263: 509–513.

Bernasconi G, Strassmann JE (1999) Cooperation among unrelated individuals: the ant foundress case. Trends Ecol Evol 14: 477–482.

Berner RA (2006) GEOCARBSULF: A combined model for Phanerozoic atmospheric O2 and CO2. Geochim Cosmochim Acta 70: 5653–5664.

Bernhard H, Fischbacher U, Fehr E (2006) Parochial altruism in humans. Nature 442: 912–915.

Bernstein IS, Ehardt C (1986) The influence of kinship and socialization on aggressive behaviour in rhesus monkeys (Macaca mulatta). Anim Behav 34: 739–747.

Berryman AA, ed. (1988) Dynamics of forest insect populations: patterns, causes, implications. New York, NY: Plenum Press.

Berryman AA, Dennis B, Raffa KF, Stenseth NC (1985) Evolution of optimal group attack, with particular reference to bark beetles (Coleoptera, Scolytidae). Ecology 66: 898–903.

Bertin A, Hausberger M, Henry L, Richard-Yris MA (2009) Adult:young ratio influences song acquisition in female European starlings (Sturnus vulgaris) J Comp Psychol 123: 195–203.

Bertness MD (1989) Intraspecific competition and facilitation in a northern acorn barnacle population. Ecology 70: 257–268.

Bertness MD, Shumway SW (1993) Competition and facilitation in marsh plants. Am Nat 142: 718–724.

Bertness MD, Callaway RM (1994) Positive interactions in communities. Trends Ecol Evol 9: 191–193.

Bertness MD, Hacker SD (1994) Physical stress and positive associations among marsh plants. Am Nat 144: 363–372.

Bertness MD, Yeh SM (1994) Cooperative and competitive interactions in the recruitment of marsh elders. Ecology 75: 2416–2429.

Bertness MD, Leonard GH (1997) The role of positive interactions in communities: lessons from intertidal habitats. Ecology 78: 1976–1989.

Beshers SN, Fewell JH (2001) Models of division of labor in social insects. Annu Rev Entomol 46: 413–440.

Best MA, Thorpe JP (1985) Autoradiographic study of feeding and the colonial transport of metabolites in the marine bryozoan Membranipora membranacea. Mar Biol 84: 295–300.

Bester-Meredith JK, Marler CA (2007) Social experience during development and female offspring aggression in Peromyscus mice. Ethology 113: 889–900.

Betzig L (1988) Redistribution: Equity or exploitation? In: Betzig L, Borgerhoff Mulder M, Turke P, eds. Human reproductive behavior: A Darwinian perspective. Cambridge, UK: Cambridge University Press. pp 49−63.

Betzig L (2004) Where’s the beef? It’s less about cooperation, more about conflict. Behav Brain Sci 27: 561–562.

Betzig L, Turke PW (1986) Food sharing on Ifaluk. Curr Anthropol 27: 397−400.

Bever JD (2002) Negative feedback within a mutualism: host-specific growth of mycorrhizal fungi reduced plant benefit. Proc R Soc B 269: 2595–2601.

Bever JD, Richardson SC, Lawrence BM, Holmes J, Watson M (2009) Preferential allocation to beneficial symbiont with spatial structure maintains mycorrhizal mutualism. Ecol Lett 12: 13–21.

Beye M, Neumann P, Moritz RFA (1997) Nestmate recognition and the genetic gestalt in the mound-building ant Formica polyctena. Insect Soc 44: 49–58.

Beye M, Neumann P, Chapuisat M, Pamilo P, Moritz RFA (1998) Nestmate recognition and the genetic relatedness of nests in the ant Formica pratensis. Behav Ecol Sociobiol 43: 67–72.

Beye M, Gattermeier I, Hasselmann M, Gempe T, Schioett M, et al. (2006) Exceptionally high levels of recombination across the honey bee genome. Genome Res 16: 1339–1344.

Bhadra A, Mitra A, Deshpande S, Chandrasekhar K, Naik D, Hefetz A, Gadagkar R (2010) Regulation of reproduction in the primitively eusocial wasp Ropalidia marginata: on the trail of the queen pheromone. J Chem Ecol 36: 424–431.

Bhardwaj V (2013) Characterization of the role of MrpC in Myxococcus xanthus developmental cell fate determination. Doctoral dissertation. Marburg, Germany: Philipps-Universität Marburg.

Bhaumik B, Mathur M (2003) A cooperation and competition based simple cell receptive field model and study of feed-forward linear and nonlinear contributions to orientation selectivity. J Comput Neurosci 14: 211–227.

Bhola PD, Simon SM (2009) Determinism and divergence of apoptosis susceptibility in mammalian cells. J Cell Sci 122: 4296–4302.

Bielsky IF, Young LJ (2004) Oxytocin, vasopressin, and social recognition in mammals. Peptides 25: 1565–1574.

Bielsky IF, Hu SB, Ren X, Terwilliger EF, Young LJ (2005) The V1a vasopressin receptor is necessary and sufficient for normal social recognition: a gene replacement study. Neuron 47: 503–513.

Bierbaum G, Götz F, Peschel A, Kupke T, van de Kamp M, Sahl HG (1996) The biosynthesis of the lantibiotics epidermin, gallidermin, Pep5 and epilancin K7. Antonie Van Leeuwenhoek 69: 119–127.

Biernaskie JM (2011) Evidence for competition and cooperation among climbing plants. Proc R Soc B: Biol Sci 278: 1989–1996.

Biggar SR, Crabtree GR (2001) Cell signaling can direct either binary or graded transcriptional responses. EMBO J 20: 3167–3176.

Bijma P, Wade MJ (2008) The joint effects of kin, multilevel selection and indirect genetic effects on response to genetic selection. J Evol Biol 21: 1175–1188.

Bilde T, Maklakov AA, Taylor PW, Lubin Y (2002) State dependent decisions in nest selection by a web-building spider. Anim Behav 64: 447–452.

Binmore K (1992) Fun and games. Lexington, MA: D.C. Heath.

Binmore K (1994a) Game theory and the social contract. I. Playing fair. Cambridge, MA: MIT Press.

Binmore K (1994b) Game theory and the social contract. II. Just playing. Cambridge, MA: MIT Press.

Birmingham AL, Hoover SE, Winston ML, Ydenberg RC (2004) Drifting bumble bee (Hymenoptera: Apidae) workers in commercial greenhouses may be social parasites. Can J Zool 82: 1843–1853.

Birmingham AL, Winston ML (2004) Orientation and drifting behavior of bumblebees (Hymenoptera: Apidae) in commercial tomato greenhouses. Can J Zool 82: 52–59.

Björkman O (1981) Responses to different quantum flux densities. In: Lange OL, Nobel PS, Osmond CB, Ziegler H, eds. Physiological plant ecology. I. Responses to the physical environment. Berlin, Germany: Springer-Verlag. pp 57–107.

Björkstrand E, Eriksson M, Uvnäs-Moberg K (1996) Evidence of a peripheral and a central effect of oxytocin on pancreatic hormone release in rats. Neuroendocrinology 63: 377–383.

Blacher P, Yagound B, Lecoutey E, Devienne P, Chameron S, Châline N (2013). Drifting behaviour as an alternative reproductive strategy for social insect workers. Proc R Soc B Biol Sci 280: 20131888.

Blackman S (2004) Spite: Evolution finally gets nasty. The Scientist 18: 14–15.

Blaffer Hrdy S (1979) Infanticide among animals: a review, clarification and examination of the implication for the reproductive strategies of females. Ethol Sociobiol 1: 13–40.

Blalock G, Just DR, Simon DH (2007) Hitting the jackpot or hitting the skids: entertainment, poverty, and the demand for state lotteries. Am J Econ Sociol 66: 545–570.

Blanchard RJ, Wall PM, Blanchard DC (2003) Problems in the study of rodent aggression. Horm Behav 44: 161–170.

Blat Y, Eisenbach M (1995) Tar-dependent and -independent pattern formation by Salmonella typhimurium. J Bacteriol 177: 1683–1691.

Blaustein AR (1983) Kin recognition mechanisms: phenotypic matching or recognition alleles? Am Nat 121: 749–754.

Blavatskyy PR (2007) Stochastic expected utility theory. J Risk Uncertainty 34: 259–286.

Blick J (1977) Selection for traits which lower individual reproduction. J Theor Biol 67: 597–601.

Bloch G, Borst DW, Huang ZY, Robinson GE, Cnaani J, Hefetz A (2000a) Juvenile hormone titers, juvenile hormone biosynthesis, ovarian development and social environment in Bombus terrestris. J Insect Physiol 46: 47–57.

Bloch G, Simon T, Robinson GE, Hefetz A (2000b) Brain biogenic amines and reproductive dominance in bumble bees (Bombus terrestris). J Comp Physiol A 186: 261–268.

Blomquist GJ, Bagnères AG, eds. (2010) Insect hydrocarbons: biology, biochemistry and chemical ecology. Cambridge, UK: Cambridge University Press.

Blouin MS (2003) DNA-based methods for pedigree reconstruction and kinship analysis in natural populations. Trends Ecol Evol 18: 503–511.

Blüthgen N, Menzel F, Hovestadt T, Fiala B, Blüthgen N (2007) Specialization, constraints, and conflicting interests in mutualistic networks. Curr Biol 17: 341–346.

Bobbioni-Harsch E, Frütiger S, Hughes G, Panico M, Etienne A, Zappacosta F, et al. (1995) Physiological concentrations of oxytocin powerfully stimulate insulin secretionin vitro. Endocrine 3: 55–59.

Bobe R, Leakey MG (2009) Ecology of Plio-Pleistocene mammals in the Omo-Turkana basin and the emergence of Homo. In: Grine FE, Fleagle JG, Leakey RE, eds. The first humans: Origins and early evolution of the genus Homo. New York, NY: Springer. pp 173–184.

Bode NWF, Faria JJ, Franks DW, Krause J, Wood AJ (2010a) How perceived threat increases synchronization in collectively moving animal groups. Proc R Soc B Biol Sci 277: 3065–3070.

Bode NWF, Franks DW, Wood AJ (2010b) Making noise: emergent stochasticity in collective motion. J Theor Biol 267: 292–299.

Boehm T (2006) Co-evolution of a primordial peptide-presentation system and cellular immunity. Nat Rev Immunol 6: 79–84.

Boehm T (2013) Sensory biology: a whiff of genome. Nature 496: 304–305.

Boehm T, Zufall F (2006) MHC peptides and the sensory evaluation of genotype. Trends Neurosci 29: 100–107.

Boerlijst M, Hogeweg P (1991) Self-structuring and selection: Spiral waves as a substrate for prebiotic evolution. In: Langton CG, Taylor C, Farmer JD, Rasmussen S, ed. Artificial Life, vol. 2. Redwood City, CA: Addison-Wesley. pp 255–276.

Boesen JJB, Dieteren N, Bal E, Lohman PHM, Simons JWIM (1992) A possible factor in genetic instability of cancer cells: stress-induced secreted proteins lead to decrease in replication fidelity. Carcinogenesis 13: 2407–2413.

Bohannan BJM, Lenski RE (2000) The relative importance of competition and predation varies with productivity in a model community. Am Nat 156: 329–340.

Böhm R, Rockenbach B (2013) The inter-group comparison – intra-group cooperation hypothesis: comparisons between groups increase efficiency in public goods provision. PLoS ONE 8: e56152.

Bohn KM, Moss CF, Wilkinson GS (2009) Pup guarding by greater spear-nosed bats. Behav Ecol Sociobiol 63: 1693–1703.

Bohnet I, Greig F, Herrmann B, Zeckhauser R (2008) Betrayal aversion: Evidence from Brazil, China, Oman, Switzerland, Turkey, and the United States. Am Econ Rev 98: 294–310.

Boland CR, Heinsohn R, Cockburn A (1997) Deception by helpers in cooperatively breeding white-winged choughs and its experimental manipulation. Behav Ecol Sociobiol 41: 251–256.

Boles BR, Singh PK (2008) Endogenous oxidative stress produces diversity and adaptability in biofilm communities. Proc Natl Acad Sci USA 105: 12503–12508.

Bolnick DI (2001) Intraspecific competition favours niche width expansion in Drosophila melanogaster. Nature 410: 463–466.

Bolnick DI (2004) Can intraspecific competition drive disruptive selection? An experimental test in natural populations of sticklebacks. Evolution 58: 608–618.

Bolton GE, Brandts J, Ockenfels A (2005) Fair procedures: Evidence from games involving lotteries. Econ J 115: 1054–1076.

Bonabeau E, Theraulaz G, Deneubourg JL, Aron S, Camazine S (1997) Self-organization in social insects. Trends Ecol Evol 12: 188–193.

Bonabeau E, Dorigo M, Theraulaz G (1999) Swarm intelligence: from natural to artificial systems. Oxford, UK: Oxford University Press.

Bonadonna F, Sanz-Aguilar A (2012) Kin recognition and inbreeding avoidance in wild birds: the first evidence for individual kin-related odour recognition. Anim Behav 84: 509–513.

Bonasio R, Li Q, Lian J, Mutti NS, Jin L, Zhao H, et al. (2012) Genome-wide and caste-specific DNA methylomes of the ants Camponotus floridanus and Harpegnathos saltator. Curr Biol 22: 1755–1764.

Bondar T, Medzhitov R (2010) p53-mediated hematopoietic stem and progenitor cell competition. Cell Stem Cell 6: 309–322.

Bonner JT (1982) Evolutionary strategies and developmental constraints in the cellular slime-molds. Am Nat 119: 530–552.

Bonner JT (1998) The origins of multicellularity. Integr Biol Issues News Rev 1: 27–36.

Bonner JT (2000) First signals: the evolution of multicellular development. Princeton, NJ: Princeton University Press.

Bono JM, Crespi BJ (2006) Costs and benefits of joint-nesting in Australian Acacia thrips. Insect Soc 53: 489–495.

Bono JM, Crespi BJ (2008) Cofoundress relatedness and group productivity in colonies of social Dunatothrips (Insecta: Thysanoptera) on Australian Acacia. Behav Ecol Sociobiol 62: 1489–1498.

Bonsall MB (2010) Parasite replication and the evolutionary epidemiology of parasite virulence. PLoS ONE 5: e12440.

Bonsall MB, Klug H (2011) The evolution of parental care in stochastic environments. J Evol Biol 24: 645–655.

Bonsall MB; Wright AE (2012) Altruism and the evolution of resource generalism and specialism. Ecol Evol 2: 515–524.

Boomsma JJ (1991) Adaptive colony sex ratios in primitively eusocial bees. Trends Ecol Evol 6: 92–95.

Boomsma JJ (2007) Kin selection versus sexual selection: why the ends do not meet. Curr Biol 17: R673–R683.

Boomsma JJ (2009) Lifetime monogamy and the evolution of eusociality. Phil Trans R Soc Lond B 364: 3191–3207.

Boomsma JJ (2013) Beyond promiscuity: mate-choice commitments in social breeding. Phil Trans R Soc B 368: 20120050.

Boomsma JJ, Grafen A (1990) Intraspecific variation in ant sex ratios and the Trivers-Hare hypothesis. Evolution 44: 1026–1034.

Boomsma JJ, Grafen A (1991) Colony-level sex ratio selection in the eusocial Hymenoptera. J Evol Biol 4: 383–407.

Boomsma JJ, Ratnieks FLW (1996) Paternity in eusocial Hymenoptera. Phil Trans R Soc Lond B 351: 947–975.

Boomsma JJ, Fjerdingstad EJ, Frydenberg J (1999) Multiple paternity, relatedness and genetic diversity in Acromyrmex leaf-cutter ants. Proc R Soc Lond Ser B 266: 249–254.

Boomsma JJ, Beekman M, Cornwallis CK, Griffin AS, Holman L, Hughes WHO, et al. (2011) Only full-sibling families evolved eusociality. Nature 471: E4–E5.

Boomsma JJ, d’Ettorre P (2013) Nice to kin and nasty to non-kin: revisiting Hamilton's early insights on eusociality. Biol Lett 9: 20130444.

Booth DJ (1995) Juvenile groups in a coral-reef damselfish: Density-dependent effects on individual fitness and population demography. Ecology 76: 91–106.

Booth KK, Katz LS (2000) Role of the vomeronasal organ in neonatal offspring recognition in sheep. Biol Reprod 63: 953–958.

Bordy EM, Bumby AJ, Catuneanu O, Eriksson PG (2009) Possible trace fossils of putative termite origin in the Lower Jurassic (Karoo Supergroup) of South Africa and Lesotho. S Afric J Sci 105: 356–362.

Borenstein E, Feldman MW, Aoki K (2008) Evolution of learning in fluctuating environments: When selection favors both social and exploratory individual learning. Evolution 62: 568–602.

Bornholdt S, Rohlf T (2000) Topological evolution of dynamical networks: Global criticality from local dynamics. Phys Rev Lett 84: 6114.

Bornstein G, Erev I, Rosen O (1990) Intergroup competition as a structural solution to social dilemmas. Soc Behav 5: 247–260.

Bornstein P, Sage EH (2002) Matricellular proteins: extracellular modulators of cell function. Curr Opin Cell Biol 14: 608–616.

Bos PA, Panksepp J, Bluthé RM, Honk JV (2012) Acute effects of steroid hormones and neuropeptides on human social-emotional behavior: a review of single administration studies. Front Neuroendocrinol 33: 17–35.

Bosch O, Nguyen N, Sun D (2013) Addressing the critical need for ‘new ways of thinking’ in managing complex issues in a socially responsible way. Business Syst Rev 2: 48–70.

Bosch OJ, Meddle SL, Beiderbeck DI, Douglas AJ, Neumann ID (2005) Brain oxytocin correlates with maternal aggression: link to anxiety. J Neurosci 25: 6807–6815.

Bostrom N (2003) Mysteries of self-locating belief and anthropic reasoning. Harvard Rev Philos 11: 59–73.

Botta AL, Santacecilia A, Ercole C, Cacchio P, Del Gallo M (2013) In vitro and in vivo inoculation of four endophytic bacteria on Lycopersicon esculentum. New Biotechnol http://dx.doi.org/10.1016/j.nbt.2013.01.001.< /p>

Boucher DH, ed. (1985) The biology of mutualism: ecology and evolution. New York, NY: Oxford University Press.

Boucher DH, James S, Keeler KH (1982) The ecology of mutualism. Annu Rev Ecol Syst 13: 315–347.

Boulay R, Hooper-Bui LM, Woodring J (2001) Oviposition and oogenesis in virgin fire ant females Solenopsis invicta are associated with a high level of dopamine in the brain. Phys Ent 26: 294–299.

Boulay R, Katzav-Gozansky T, Hefetz A, Lenoir A (2004) Odour convergence and tolerance between nestmates through trophallaxis and grooming in the ant Camponotus fellah (Dalla Torre). Insect Soc 51: 55–61.

Boulay R, Cerda X, Fertin A, Ichinose K, Lenoir A (2009) Brood development into sexual females depends on the presence of a queen but not on temperature in an ant dispersing by colony fission, Aphaenogaster senilis. Ecol Entomol 34: 595–602.

Boulding K (1978) Ecodynamics. Beverly Hills, CA: Sage.

Bourke AFG (1988a) Worker reproduction in the higher eusocial Hymenoptera. Q Rev Biol 63: 291–311.

Bourke AFG (1988b) Dominance orders, worker reproduction, and queen-worker conflict in the slave-making ant Harpagoxenus sublaevis. Behav Ecol Sociobiol 23: 323–333.

Bourke AF (2011a) The validity and value of inclusive fitness theory. Proc R Soc B Biol Sci 278: 3313–3320.

Bourke AFG (2011b) Principles of social evolution. Oxford, UK: Oxford University Press.

Bourke AFG, Heinze J (1994) The ecology of communal breeding: the case of multiple-queen leptothoracine ants. Phil Trans R Soc Lond Ser B 345: 359–372.

Bourke AFG, Franks NR (1995) Social evolution in ants. Princeton, NJ: Princeton University Press.

Bouwma AM, Nordheim EV, Jeanne RL (2006) Per-capita productivity in a social wasp: no evidence for a negative effect of colony size. Insect Soc 53: 412–419.

Bowen MT, McGregor IS (2014) Oxytocin and vasopressin modulate the social response to threat: a preclinical study. Int J Neuropsychopharmacol 17: 1621–1633.

Bowles S (2006) Group competition, reproductive leveling, and the evolution of human altruism. Science 314: 1569–1572.

Bowles S (2009) Did warfare among ancestral hunter–gatherers affect the evolution of human social behaviors? Science 324: 1293–1298.

Bowles S, Gintis H (2004) The evolution of strong reciprocity: cooperation in heterogeneous populations. Theor Popul Biol 65: 17–28.

Bowles S, Gintis H (2011) A cooperative species: human reciprocity and its evolution. Princeton, NJ: Princeton University Press.

Bowman LA, Dilley SR, Keverne EB (1978) Suppression of oestrogen-induced LH surges by social subordination in talapoin monkeys. Nature 275: 56–58.

Boyd R (1989) Mistakes allow evolutionary stability in the repeated Prisoner’s Dilemma game. J Theor Biol 136: 47–56.

Boyd R (1992) The evolution of reciprocity when conditions vary. In: Harcourt AH, de Waal FBM, eds. Coalitions and alliances in humans and other animals. Oxford, UK: Oxford University Press. pp 473−489.

Boyd R (2006) The puzzle of human sociality. Science 314: 1555–1556.

Boyd R, Richerson PJ (1988) The evolution of reciprocity in sizable groups. J Theor Biol 132: 337–356.

Boyd R, Richerson PJ (1989) The evolution of indirect reciprocity. Social Networks 11: 213–236.

Boyd R, Richerson P (1990) Group selection among alternative evolutionarily stable strategies. J Theor Biol 145: 331–342.

Boyd R, Richerson PJ (1992) Punishment allows the evolution of cooperation (or anything else) in sizable groups. Ethol Sociobiol 13: 171–195.

Boyd R, Richerson PJ (2002) Group beneficial norms can spread rapidly in structured populations. J Theor Biol 215: 287–296.

Boyd R, Gintis H, Bowles S, Richerson PJ (2003) The evolution of altruistic punishment. Proc Natl Acad Sci USA 100: 3531–3535.

Boyd R, Gintis H, Bowles S (2010) Coordinated punishment of defectors sustains cooperation and can proliferate when rare. Science 328: 617–620.

Boydston EE, Morelli TL, Holekamp KE (2001) Sex differences in territorial behavior exhibited by the spotted hyena (Hyaenidae, Crocuta crocuta). Ethology 107: 369–385.

Boza G, Scheuring I (2004) Environmental heterogeneity and the evolution of mutualism. Ecol Complex 1: 329–339.

Bradbury JW, Vehrencamp SL (1976) Social organization and foraging in emballonurid bats II. A model for the determination of group size. Behav Ecol Sociobiol 2: 383–404.

Bradley TJ, Briscoe AD, Brady SG, Contreras HL, Danforth BN, Dudley R, et al. (2009) Episodes in insect evolution. Integr Comp Biol 49: 590–606.

Brady SG, Sipes S, Pearson A, Danforth BN (2006) Recent and simultaneous origins of eusociality in halictid bees. Proc R Soc Lond B Biol Sci 273: 1643–1649.

Braendle C, Flatt T (2006) A role for genetic accommodation in evolution? Bioessays 28: 868–873.

Brain C (1992) Deaths in a desert baboon troop. Int J Primatol 13: 593–599.

Branch GM, Barkai A (1988) Interspecific behaviour and its reciprocal interaction with evolution, population dynamics and community structure. In: Chelazzi G, Vannini M, ed. Behavioral adaptation to intertidal life. New York, NY: Plenum Press. pp 225–254.

Branchi I, D’Andrea I, Fiore M, Di Fausto V, Aloe L, Alleva E (2006) Early social enrichment shapes social behavior and nerve growth factor and brain-derived neurotrophic factor levels in the adult mouse brain. Biol Psychiat 60: 690–696.

Branchi I, D’Andrea I, Gracci F, Santucci D, Alleva E (2009) Birth spacing in the mouse communal nest shapes adult emotional and social behavior. Physiol Behav 96: 532–539.

Branda SS, González-Pastor JE, Ben-Yehuda S, Losick R, Kolter R (2001) Fruiting body formation by Bacillus subtilis. Proc Natl Acad Sci USA 98: 11621–11626.

Brandstaetter AS, Endler A, Kleineidam CJ (2008) Nestmate recognition in ants is possible without tactile interaction. Naturwissenschaften 95: 601–608.

Brandt H, Sigmund K (2004) The logic of reprobation: assessment and action rules for indirect reciprocation. J Theor Biol 231: 475–486.

Brandt M, van Wilgenburg E, Sulc R, Shea KJ, Tsutsui ND (2009) The scent of supercolonies: the discovery, synthesis and behavioural verification of ant colony recognition cues. BMC Biol 7: 71.

Brännström A, Dieckmann U (2005) Evolutionary dynamics of altruism and cheating among social amoebas. Proc R Soc B 272: 1609–1616.

Brauchli K, Killingback T, Doebeli M (1999) Evolution of cooperation in spatially structured populations. J Theor Biol 200: 405–417.

Bräuer J, Call J, Tomasello M (2009) Are apes inequity averse? New data on the token-exchange paradigm. Am J Primatol 71: 175–181.

Braun V (1997) Avoidance of iron toxicity through regulation of bacterial iron transport. Biol Chem 378: 779–786.

Breden FJ, Wade MJ (1989) Selection within and between kin groups of the imported willow leaf beetle. Am Nat 134: 35–50.

Breder Jr CM (1967) On the survival value of fish schools. Zoologica 52: 25–40.

Breed MD (1983) Correlations between juvenile hormone and aggressive behavior in worker honeybees. Insect Soc 30: 482–495.

Breed MD, Gamboa GJ (1977) Behavioral control of workers by queens in primitively eusocial bees. Science 195: 694–695.

Breed MD, Smith TA, Torres A (1992) Guard honey bees: role in nestmate recognition and replacement. Ann Entomol Soc Am 85: 633–637.

Breed MD, Welch CK, Cruz R (1994) Kin discrimination within honey bee (Apis mellifera) colonies: an analysis of the evidence. Behav Process 33: 25–40.

Breivik J, Gaudernack G (1999) Genomic instability, DNA methylation, and natural selection in colorectal carcinogenesis. Semin Cancer Biol 9: 245–254.

Brembs B, Lorenzetti FD, Reyes FD, Baxter DA, Byrne JH (2002) Operant reward learning in Aplysia: neuronal correlates and mechanisms. Science 296: 1706–1709.

Brent CS, Vargo EL (2003) Changes in juvenile hormone biosynthetic rate and whole body content in maturing virgin queens of Solenopsis invicta. J Insect Physiol 49: 967–974.

Briand F, Yodzis P (1982) The phylogenetic distribution of obligate mutualism: evidence of limiting similarity and global instability. Oikos 39: 273–275.

Brierley AS, Cox MJ (2010) Shapes of krill swarms and fish schools emerge as aggregation members avoid predators and access oxygen. Curr Biol 20: 1758–1762.

Briggman KL, Abarbanel HD, Kristan WB Jr (2005) Optical imaging of neuronal populations during decision-making. Science 307: 896–901.

Brinner RE, Clotfelter CT (1975) An economic appraisal of state lotteries. Natl Tax J 23: 395–404.

Brock DA, Read S, Bozhchenko A, Queller DC, Strassmann JE (2013) Social amoeba farmers carry defensive symbionts to protect and privatize their crops. Nat Commun 4: 2385.

Brockhurst MA, Rainey PB, Buckling A (2004) The effect of spatial heterogeneity and parasites on the evolution of host diversity. Proc R Soc Lond B 271: 107–111.

Brockhurst MA, Buckling A, Rainey PB (2005) The effect of a bacteriophage on diversification of the opportunistic bacterial pathogen, Pseudomonas aeruginosa. Proc R Soc Lond B 272: 1385–1391.

Brockhurst MA, Hochberg ME, Bell T, Buckling A (2006) Character displacement promotes cooperation in bacterial biofilms. Curr Biol 16: 2030–2034.

Brockhurst MA, Buckling A, Gardner A (2007) Cooperation peaks at intermediate disturbance. Curr Biol 17: 761–765.

Brockhurst MA, Buckling A, Racey D, Gardner A (2008) Resource supply and the evolution of public-goods cooperation in bacteria. BMC Biol 6: 20.

Brockhurst MA, Habets MG, Libberton B, Buckling A, Gardner A (2010) Ecological drivers of the evolution of public-goods cooperation in bacteria. Ecology 91: 334–340.

Brockhurst MA, Chapman T, King KC, Mank JE, Paterson S, Hurst GDD (2014) Running with the Red Queen: the role of biotic conflicts in evolution. Proc R Soc B 281: 20141382.

Brockmann HJ, Barnard CJ (1979) Kleptoparasitism in birds. Anim Behav 27: 487–514.

Brockmann HJ, Grafen A, Dawkins R (1979) Evolutionarily stable nesting strategy in a digger wasp. J Theor Biol 77: 473–496.

Brockmann JH (1997) Cooperative breeding in wasps and vertebrates: The role of ecological constraints. In: Choe JC, Crespi BJ, eds. The evolution of social behavior in insects and arachnids. Cambridge, UK: Cambridge University Press. pp 347–371.

Broman KW, Murray JC, Sheffield VC, White RL, Weber JL (1998) Comprehensive human genetic maps: individual and sex-spesific variation in recombination. Am J Hum Genet 63: 861–869.

Bronstein JL (1994) Our current understanding of mutualism. Q Rev Biol 69: 31–51.

Bronstein JL (2001) The exploitation of mutualisms. Ecol Lett 4: 277–287.

Bronstein JL (2009) The evolution of facilitation and mutualism. J Ecol 97: 1160–1170.

Bronstein JL, Alarcón R, Geber M (2006) Tansley review: evolution of insect⁄plant mutualisms. New Phytol 172: 412–428.

Brook BW, Bradshaw CJA (2006) Strength of evidence for density dependence in abundance time series of 1198 species. Ecology 87: 1445–1451.

Brooker RW, Callaghan TV (1998) The balance between positive and negative plant interactions and its relationship to environmental gradients: a model. Oikos 81: 196–207.

Brookfield JFY (1998) Quorum sensing and group selection. Evolution 52: 1263–1269.

Brookman JJ, Jermyn KA, Kay RR (1987) Nature and distribution of the morphogen DIF in the Dictyostelium slug. Development 100: 119–124.

Brooks AC (2006) Who really cares. New York, NY: Basic Books.

Brooks JL, Dodson SI (1965) Predation, body size, and composition of plankton. Science 150: 28–35.

Brosnan SF (2011) A hypothesis of the co-evolution of cooperation and responses to inequity. Front Neurosci 5: 43.

Brosnan SF, de Waal FBM (2003) Monkeys reject unequal pay. Nature 425: 297–299.

Brosnan SF, Bshary R (2010) Cooperation and deception: from evolution to mechanisms. Phil Trans R Soc B 365: 2593–2598.

Brosnan SF, Talbot C, Ahlgren M, Lambeth SP, Schapiro SJ (2010a) Mechanisms underlying responses to inequitable outcomes in chimpanzees, Pan troglodytes. Anim Behav 79: 1229–1237.

Brosnan SF, Salwiczek L, Bshary R (2010b) The interplay of cognition and cooperation. Phil Trans R Soc B 365: 2699–2710.

Brosnan SF, de Waal FB (2014) Evolution of responses to (un) fairness. Science 346: 1251776.

Brothers DJ, Michener CD (1974) Interactions in colonies of primitively eusocial bees. III. Ethometry of division of labor in Lasioglossum zephyrum (Hymenoptera: Halictidae). J Comp Physiol 90: 129–168.

Broughton S, Partridge L (2009) Insulin/IGF-like signalling, the central nervous system and aging. Biochem J 418: 1–12.

Brown BE (1979) Mammalian social odors: a critical review. Adv Study Behav 10: 103–160.

Brown C, Laland KN (2003) Social learning in fishes: a review. Fish Fisheries 4: 280–288.

Brown C, Davidson T, Laland K (2003) Environmental enrichment and prior experience of live prey improve foraging behaviour in hatchery-reared Atlantic salmon. J Fish Biol 63: 187–196.

Brown C, Jones F, Braithwaite VA (2007) Correlation between boldness and body mass in natural populations of the poeciliid Brachyrhaphis episcopi. J Fish Biol 71: 1590–1601.

Brown CJ, Todd KM, Rosenzweig RF (1998) Multiple duplications of yeast hexose transport genes in response to selection in a glucose-limited environment. Mol Biol Evol 15: 931–942.

Brown CR, Stutchbury BJ, Walsh PD (1990) Choice of colony size in birds. Trends Ecol Evol 5: 398–403.

Brown CR, Brown MB (1996) Coloniality in the cliff swallow: the effect of group size on social behavior. Chicago, IL: University of Chicago Press.

Brown CT, Fishwick LK, Chokshi BM, Cuff MA, Jackson JM 4th, Oglesby T, et al. (2011) Whole-genome sequencing and phenotypic analysis of Bacillus subtilis mutants following evolution under conditions of relaxed selection for sporulation. Appl Environ Microbiol 77: 6867–6877.

Brown GL, Goodwin FK, Ballenger JC, Goyer PF, Major LF (1979) Aggression in humans correlates with cerebrospinal fluid amine metabolites. Psychiatry Res 1: 131–139.

Brown GP, Weatherhead PJ (2004) Sexual abstinence and the cost of reproduction in adult male water snakes, Nerodia sipedon. Oikos 104: 269–276.

Brown JH(1981) Two decades of homage to Santa Rosalia: toward a general theory of diversity. Am Zool 21: 877–888.

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, Gillooly JF, Allen AP, Savage VM, West GB (2004) Toward a metabolic theory of ecology. Ecology 81: 1771–1789.

Brown JL (1969) Territorial behavior and population regulation in birds. Wilson Bull 81: 293–329.

Brown JL (1974) Alternative routes to sociality in jays–with a theory for the evolution of altruism and communal breeding. Am Zool 14: 63–80.

Brown JL (1976) The evolution of behavior. New York, NY: Norton.

Brown JL (1983) Cooperation – a biologists dilemma. Adv Study Behav 13: 1–37.

Brown JL (1987) Helping and communal breeding in birds: ecology and evolution. Princeton, NJ: Princeton University Press.

Brown JL (1997) A theory of mate choice based on heterozygosity. Behav Ecol 8: 60–65.

Brown JL, Brown ER (1990) Mexican jays: uncooperative breeding. In: Stacey PB, Koenig WD, eds. Cooperative breeding in birds. Cambridge, UK: Cambridge University Press. pp 267–288.

Brown JL, Eklund A (1994) Kin recognition and the major histocompatibility complex: an integrative review. Am Nat 143: 435–461.

Brown MJF, Schmid-Hempel P (2003) The evolution of female multiple mating in social hymenoptera. Evolution 57: 2067–2081.

Brown MJF, Bonhoeffer S (2003) On the evolution of claustral colony founding in ants. Evol Ecol Res 5: 305–313.

Brown SP, Le Chat L, Taddei F (2007) Evolution of virulence: triggering host inflammation allows invading pathogens to exclude competitors. Ecol Lett 11: 44–51.

Brown WM, Consedine NS, Magai C (2005) Altruism relates to health in an ethnically diverse sample of older adults. J Gerontol Psychol Sci 60B:143–152.

Bruder CE, Piotrowski A, Gijsbers AA, Andersson R, Erickson S, Diaz de Ståhl T, et al. (2008) Phenotypically concordant and discordant monozygotic twins display different DNA copy-number-variation profiles. Am J Hum Genet 82: 763–771.

Bruhn TO, Sutton SW, Plotsky PM, Vale WW (1986) Central administration of corticotropin-releasing factor modulates oxytocin secretion in the rat. Endocrinology 119: 1558–1563.

Bruintjes R, Taborsky M (2008) Helpers in a cooperative breeder pay a high price to stay: effects of demand, helper size and sex. Anim Behav 75: 1843–1850.

Brüning JC, Gautam D, Burks DJ, Gillette J, Schubert M, Orban PC, et al. (2000) Role of brain insulin receptor in control of body weight and reproduction. Science 289: 2122–2125.

Brunner E, Heinze J (2009) Worker dominance and policing in the ant Temnothorax unifasciatus. Insectes Soc 56: 397–404.

Brunner HG, van Driel MA (2004) From syndrome families to functional genomics. Nat Rev Genet 5: 545–551.

Bry C, Basset E, Rognon X, Bonamy F (1992) Analysis of sibling cannibalism among pike, Esox lucius, juveniles reared under semi-natural conditions. Environ Biol Fish 35: 75–84.

Bryan JH, Test MA (1967) Models and helping: Naturalistic studies in aiding behavior. J Pers Soc Psychol 6: 400–407.

Bshary R (2002) Biting cleaner fish use altruism to deceive image-scoring client reef fish. Proc Biol Sci 269: 2087–2093.

Bshary R, Grutter AS (2002) Experimental evidence that partner choice is a driving force in the payoff distribution among cooperators or mutualists: the cleaner fish case. Ecol Lett 5: 130–136.

Bshary R & Noë R (2003) Biological markets: the ubiquitous influence of partner choice on the dynamics of cleaner fish-client fish interactions. In: Hammerstein P, ed. Genetic and cultural evolution of cooperation. Cambridge, MA: MIT Press. pp 167–184.

Bshary R, Grutter AB (2006) Image scoring and cooperation in a cleaner fish mutualism. Nature 441: 975–978

Bshary R, Oliveira RF, Grutter AS (2011) Short term variation in the level of cooperation in the cleaner wrasse Labroides dimidiatus: implications for the role of potential stressors. Ethology 117: 246–253.

Buchanan TW, Bagley SL, Stansfield RB, Preston SD (2012) The empathic, physiological resonance of stress. Soc Neurosci 7: 191–201.

Buchanan TW, Preston SD (2014) Stress leads to prosocial action in immediate need situations. Front Behav Neurosci 8: http://dx.doi.org/10.3389/fnbeh.2014.00005.

Buchberger A, Bukau B, Sommer T (2010) Protein quality control in the cytosol and the endoplasmic reticulum: brothers in arms. Mol Cell 40: 238–252.

Büchi L, Vuilleumier S (2012) Dispersal strategies, few dominating or many coexisting: the effect of environmental spatial structure and multiple sources of mortality. PLoS ONE 7: e34733.

Buchwald R, Breed MD (2005) Nestmate recognition cues in a stingless bee, Trigona fulviventris. Anim Behav 70: 1331–1337.

Buckle GR (1982) Queen-worker behavior and nestmate interactions in young colonies of Lasioglossum zephyrum. Insect Soc 29: 125–137.

Buckling A, Rainey PB (2002) The role of parasites in sympatric and allopatric host diversification. Nature 420: 496–499.

Buczkowski G, Kumar R, Suib SL, Silverman J (2005) Diet-related modification of cuticular hydrocarbon profiles of the Argentine ant, Linepithema humile, diminishes intercolony aggression. J Chem Ecol 31: 829–843.

Budding AE, Ingham CJ, Bitter W, Vandenbroucke-Grauls CM, Schneeberger PM (2009) The Dienes phenomenon: competition and territoriality in swarming Proteus mirabilis. J Bacteriol 191: 3892–3900.

Budrene EO, Berg HC (1991) Complex patterns formed by motile cells of Escherichia coli. Nature 349: 630–633.

Buitenhuis B, Hedegaard J, Janss L, Sørensen P (2009) Differentially expressed genes for aggressive pecking behaviour in laying hens. BMC Genomics 10: 544.

Bukau B, Weissman J, Horwich A (2006) Molecular chaperones and protein quality control. Cell 125: 443–451.

Bull JJ, Shine R (1979) Iteroparous animals that skip opportunities for reproduction. Am Nat 114: 296–303.

Bull JJ, Rice WR (1991) Distinguishing mechanisms for the evolution of cooperation. J Theor Biol 149: 63–74.

Bull NJ, Schwarz MP (1996) The habitat saturation hypothesis and sociality in an allodapine bee: cooperative nesting is not “making the best of a bad situation”. Behav Ecol Sociobiol 39: 267–274.

Bull NJ, Schwarz MP (1997) Rearing of non-descendent offspring in an allodapine bee, Exoneura bicolor Smith (Hymenoptera: Apidae: Xylocopinae): a preferred strategy or queen coercion? Aust J Entomol 36: 391–394.

Bull NJ, Schwarz MP (2001) Brood insurance via protogyny: a source of female-biased sex allocation. Proc R Soc Lond B 268: 1869–1874.

Bulmer MG (1980) The sib competition model for the maintenance of sex and recombination. J Theor Biol 82: 335–345.

Burke T, Davies NB, Bruford MW, Hatchwell BJ (1989) Parental care and mating behaviour of polyandrous dunnocks Prunella modularis related to paternity by DNA fingerprinting. Nature 338: 249–251.

Burkhart JM, Hrdy SB, van Schaik CP (2009) Cooperative breeding and human cognitive evolution. Evol Anthropol 18: 175–186.

Burks DJ, Font de Mora J, Schubert M, Withers DJ, Myers MG, Towery HH, et al. (2000) IRS-2 pathways integrate female reproduction and energy homeostasis. Nature 407: 377–382.

Burn CC, Mason GJ (2008) Effects of cage-cleaning frequency on laboratory rat reproduction, cannibalism, and welfare. Appl Anim Behav Sci 114: 235–247.

Burnet FM (1957) A modification of Jerne’s theory of antibody production using the concept of clonal selection. Aust J Sci 20: 67–69.

Burnet FM (1971) "Self-recognition'' in colonial marine forms and flowering plants in relation to the evolution of immunity. Nature 232: 230–235.

Burt A, Bell G (1987) Mammalian chiasma frequencies as a test of two theories of recombination. Nature 326: 803–805.

Burt A, Trivers R (2006) Genes in conflict: the biology of selfish genetic elements. Cambridge, MA: Belknap Press.

Burtsev M, Turchin P (2006) Evolution of cooperative strategies from first principles. Nature 440: 1041–1044.

Buser P (2002) Silk attraction: base of group cohesion and collective behaviours in social spiders. C R Biol 325: 1153–1157.

Buskirk R (1981) Sociality in the Arachnida. In: Hermann HR, ed. Social insects, vol. 2. New York, NY: Academic Press. pp 282–367.

Buss LW (1981) Group living, competition, and the evolution of cooperation in a sessile invertebrate. Science 213: 1012–1014.

Buss LW (1982) Somatic cell parasitism and the evolution of somatic tissue compatibility. Proc Natl Acad Sci USA 79: 5337–5341.

Buss LW (1983) Evolution, development, and the units of selection. Proc Natl Acad Sci USA 80: 1387–1391.

Buss LW (1987) The evolution of individuality. Princeton, NJ: Princeton University Press.

Buston P (2003a) Forcible eviction and prevention of recruitment in the clown anemonefish. Behav Ecol 14: 576–582.

Buston PM (2003b) Social hierarchies: size and growth modifications in clownfish. Nature 424: 145–146.

Buston PM (2004) Territory inheritance in clownfish. Proc R Soc Lond Ser B Biol Sci 271(Suppl 4): S252–S254.

Buston PM, Bogdanowicz SM, Wong A, Harrison RG (2007) Are clownfish groups composed of close relatives? An analysis of microsatellite DNA variation in Amphiprion percula. Mol Ecol 16: 3671–3678.

Butler CG, Callow RK, Johnston NC (1959) Extraction and purification of ‘queen substance’ from queen bees. Nature 184: 1871.

Butterfield NJ (2009) Oxygen, animals and oceanic ventilation: An alternative view. Geobiology 7: 1–7.

Buttery NJ, Rozen DE, Wolf JB, Thompson CRL (2009) Quantification of social behavior in D. discoideum reveals complex fixed and facultative strategies. Curr Biol 19: 1373–1377.

Buttery NJ, Thompson CRL, Wolf JB (2010) Complex genotype interactions influence social fitness during the developmental phase of the social amoeba Dictyostelium discoideum. J Evol Biol 23: 1664–1671.

Bygott JD, Bertram BCR, Hanby JP (1979) Male lions in large coalitions gain reproductive advantages. Nature 282: 838–840.

Byrne D, Callaghan G (2014) Complexity theory and the social sciences. New York, NY: Routledge.

Byrne RW (1996) Machiavellian intelligence. Evol Anthropol 5: 172–180.

Byrne RW, Bates LA (2007) Sociality, evolution, and cognition. Curr Biol 17: R714–R723.

Cahan S, Helms KR, Rissing S (1998) An abrupt transition in colony founding behaviour in the ant Messor pergandei. Anim Behav 55: 1583–1594.

Cain ML, Bowman WD, Hacker SD (2011) Ecology. Suntherland, MA: Sinauer Associates.

Calabi P, Traniello JFA (1989) Social organization in the ant Pheidole dentata: physical and temporal caste ratios lack ecological correlates. Behav Ecol Sociobiol 24: 69–78.

Caldera EJ, Holway DA (2004) Evidence that queens do not influence nestmate recognition in Argentine ants. Insect Soc 51: 109–112.

Calderone NW, Page RE Jr (1988) Genotypic variability in age polyethism and task specialization in the honey bee Apis mellifera (Hymenoptera: Apidae). Behav Ecol Sociobiol 22: 17–25.

Caldwell HK, Young WS III (2006) Oxytocin and vasopressin: genetics and behavioral implications. In: Lim R, ed. Handbook on neurochemistry and molecular neurobiology, edn 3. New York, NY: Springer. pp 573–607.

Caldwell HK, Lee H-J, Macbeth AH, Young WS (2008) Vasopressin: behavioural roles of an “original” neuropeptide. Prog Neurobiol 84: 1–24.

Caligari PDS (1980) Competitive interactions in Drosophila melanogaster. I. Monocultures. Heredity 45: 219–231.

Callaway RM, Walker LR (1997) Competition and facilitation: a synthetic approach to interactions in plant communities. Ecology 78: 1958–1965.

Callaway RM, Brooker RW, Choler P, Kikvidze Z, Lortie CJ, et al. (2002) Positive interactions among alpine plants increase with stress. Nature 417: 844–847.

Camazine S, Deneubourg JL, Franks NR, Sneyd J, Theraulaz G, Bonabeau E (2001) Self-organization in biological systems. Princeton, NJ: Princeton University Press.

Camerer CF (1989) An experimental test of several generalized utility theories. J Risk Uncertainty 2: 61–104.

Cameron DD, Leake JR, Read DJ (2006) Mutualistic mycorrhiza in orchids: evidence from plant-fungus carbon and nitrogen transfers in the green-leaved terrestrial orchid Goodyera repens. New Phytol 171: 405–416.

Cameron EZ, Setsaas TH, Linklater WL (2009) Social bonds between unrelated females increase reproductive success in feral horses. Proc Natl Acad Sci USA 106: 13850–13853.

Cameron SA, Mardulyn P (2001) Multiple molecular data sets suggest independent origins of highly eusocial behavior in bees (Hymenoptera:Apinae). Syst Biol 50: 194–214.

Camiletti AL, Percival-Smith A, Thompson GJ (2013) Honey bee queen mandibular pheromone inhibits ovary development and fecundity in a fruit fly. Entomol Exp Appl 147: 262–268.

Camiletti AL, Awde DN, Thompson GJ (2014) How flies respond to honey bee pheromone: the role of the foraging gene on reproductive response to queen mandibular pheromone. Naturwissenschaften 101: 25–31.

Campagna C, Bisioli C, Quintana F, Perez F, Vila A (1992) Group breeding in sea lions: pups survive better in colonies. Anim Behav 43: 541–548.

Campbell A (2008) Attachment, aggression and affiliation: the role of oxytocin in female social behavior. Biol Psychol 77: 1–10.

Candolin U (2003) The use of multiple cues in mate choice. Biol Rev 78: 575–595.

Cant MA (2006) A tale of two theories: parent-offspring conflict and reproductive skew. Anim Behav 71: 255–263.

Cant MA (2010) The role of threats in animal cooperation. Proc R Soc Lond Ser B 278: 170–178.

Cant MA (2012) Suppression of social conflict and evolutionary transitions to cooperation. Am Nat 179: 293–301.

Cant MA, Johnstone RA (2000) Power struggles, dominance testing, and reproductive skew. Am Nat 155: 406–17.

Cant MA, Field J (2001) Helping effort and future fitness in cooperative animal societies. Proc R Soc Lond B 268: 1959–1964.

Cant MA, Field J (2005) Helping effort in a dominance hierarchy. Behav Ecol 16: 708–715.

Cant MA, English S (2006) Stable group size in cooperative breeders: the role of inheritance and reproductive skew. Behav Ecol 17: 560–568.

Cant MA, English S, Reeve HK, Field J (2006a) Escalated conflict in a social hierarchy. Proc R Soc B Biol Sci 273: 2977–2984.

Cant MA, Llop JB, Field J (2006) Individual variation in social aggression and the probability of inheritance: theory and a field test. Am Nat 167: 837–852.

Cant MA, Hodge SJ, Bell MBV, Gilchrist JS, Nichols HJ (2010) Reproductive control via eviction (but not the threat of eviction) in banded mongooses. Proc R Soc B 277: 2219–2226.

Cao C, Brown MR (2001) Localization of an insulin-like peptide in brains of two flies. Cell Tissue Res 304: 317–321.

Cáp M, Váchová L, Palková Z (2012) Reactive oxygen species in the signaling and adaptation of multicellular microbial communities. Oxid Med Cell Longev 2012: 976753. doi:10.1155/2012/976753.

Caporale LH (2009) Putting together the pieces: evolutionary mechanisms at work within genomes. BioEssays 31: 700–702.

Caraco T (1980) On foraging time allocation in a stochastic environment. Ecology 61: 119–128.

Caraco T (1981) Risk-sensitivity and foraging groups. Ecology 62: 527–531.

Caraco T, Wolf LL (1975) Ecological determinants of groups sizes of foraging lions. Am Nat 109: 343–352.

Caraco T, Martindale S, Whittam TS (1980) An empirical demonstration of risk-sensitive foraging preferences. Anim Behav 28: 820–830.

Caraco T, Pulliam HR (1984) Sociality and survivorship in animals exposed to predation. In: Price PW, Slobodchikoff CN, Gaud WS, eds. A new ecology: novel approaches to interactive systems. New York, NY: Wiley. pp 279–309.

Caraco T, Kelly CK (1991) On the adaptive value of physiological integraton in colonal plants. Ecology 72: 81–93.

Caraco T, Uetz GW, Gillespie RG, Giraldeau LA (1995) Resource consumption variance within and among individuals: on coloniality in spiders. Ecology 76: 196–205.

Carbone C, DuToit JT, Gordon IJ (1997) Feeding success in African wild dogs: does kleptoparasitism by spotted hyenas influence hunting group size? J Anim Ecol 66: 318–326.

Carbone C, Frame L, Frame G, Malcolm J, Fanshawe J, FitzGibbon C, et al. (2005) Feeding success of African wild dogs (Lycaon pictus) in the Serengeti: the effects of group size and kleptoparasitism. J Zool 266: 153–161.

Cardinal S, Straka J, Danforth BN (2010) Comprehensive phylogeny of apid bees reveals the evolutionary origins and antiquity of cleptoparasitism. Proc Natl Acad Sci USA 107: 16207–16211.

Cardoso SC, Paitio JR, Oliveira RF, Bshary R, Soares MC (2015) Arginine vasotocin reduces levels of cooperative behaviour in a cleaner fish. Physiol Behav 139: 314–320.

Carlin NF, Hölldobler B (1983) Nestmate and kin recognition in interspecific mixed colonies of ants. Science 222: 1027–1029.

Carlin NF, Hölldobler B (1986) The kin recognition system of carpenter ants (Camponotus spp.). I. Hierachical cues in small colonies. Behav Ecol Sociobiol 19: 123–134.

Carlin NF, Hölldobler B (1987) The kin recognition system of carpenter ants (Camponotus spp.) II. Larger colonies. Behav Ecol Sociobiol 20: 209–217.

Carlin NF, Reeve HK, Cover SP (1993) Kin discrimination and division of labour among matrilines in the polygynous carpenter ant, Camponotus planatus. In: Keller L, ed. Queen number and sociality in insects. Oxford, UK: Oxford University Press. pp. 362–401.

Carlin NF, Gladstein DS, Berry AJ, Pierce NE (1994) Absence of kin discrimination behavior in a soldier-producing aphid, Ceratovacuna japonica (Hemiptera: Pemphigidae; Cerataphidini). J NY Entomol Soc 102: 287–298.

Carlsson F, Daruvala D, Johansson-Stenman O (2005) Are people inequality-averse, or just risk-averse? Economica 72: 375–396.

Carneiro RL (1970) A theory of the origin of the state. Science 169: 733–738.

Caro T (2005) Antipredator defenses in birds and mammals. Chicago, IL: University of Chicago Press.

Carpenter J, Myers CK (2010) Why volunteer? Evidence on the role of altruism, image, and incentives. J Public Econ 94: 911–920.

Carr PH (2004) Does God play dice? Insights from the fractal geometry of nature. Zygon 39: 933–940.

Carroll SB, Gates J, Keys DN, Paddock SW, Panganiban GE, Selegue JE, Williams JA (1994) Pattern formation and eyespot determination in butterfly wings. Science 265: 109–114.

Carter CS (1998) Neuroendocrine perspectives on social attachment and love. Psychoneuroendocrinology 23: 779–818.

Cascales E, Buchanan SK, Duché D, Kleanthous C, Lloubès R, Postle K, et al. (2007) Colicin biology. Microbiol Mol Biol R 71: 158–229.

Cash K, McKee M, Wrona F (1993) Short- and long-term consequences of grouping and group foraging in the free-living flatworm Dugesia tigrina. J Anim Ecol 62: 529–535.

Cashing DH, Harden-Jones FR (1968) Why do fish school? Nature 218: 918–920.

Castillo DI, Switz GT, Foster KR, Queller DC, Strassmann JE (2005) A cost to chimerism in Dictyostelium discoideum on natural substrates. Evol Ecol Res 7: 263–271.

Castillo DI, Queller DC, Strassmann JE (2011) Cell condition, competition, and chimerism in the social amoeba Dictyostelium discoideum. Ethol Ecol Evol 23: 262–273.

Cavaliere M, Poyatos JF (2013) Plasticity facilitates sustainable growth in the commons. J R Soc Interface 10: 20121006.

Cavalier-Smith T (1998) A revised six-kingdom system of life. Biol Rev Camb Phil Soc 73: 203–266.

Cavalli-Sforza LL, Feldman MW (1978) Darwinian selection and ‘‘altruism’’. Theor Popul Biol 14: 268–280.

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: 53–59.

Chadefaux T, Helbing D (2010) How wealth accumulation can promote cooperation. PLoS ONE 5: e13471.

Chagnon NA (1981) Terminological kinship, genealogical relatedness, and village fissioning among the Yanomamo Indians. In: Alexander RD, Tinkle DW, eds. Natural Selection and social behavior. New York, NY: Chiron Press. pp 490−508.

Chagnon NA, Bugos PE (1979) Kin selection and conflict: An analysis of a Yanomamo ax fight. In: Chagnon NA, Irons W, eds. Evolutionary biology and human social behavior: An anthropological perspective. North Scituate, MA: Duxbury Press. pp 213−238.

Châline N, Martin SJ, Ratnieks FL (2005) Absence of nepotism toward imprisoned young queens during swarming in the honey bee. Behav Ecol 16: 403–409.

Chambers I, Colby D, Robertson M, Nichols J, Lee S, Tweedie S, Smith A (2003) Functional expression cloning of Nanog, a pluripotency sustaining factor in embryonic stem cells. Cell 113: 643–655.

Chambers I, Silva J, Colby D, Nichols J, Nijmeijer B, Robertson M, et al. (2007) Nanog safeguards pluripotency and mediates germline development. Nature 450: 1230–1234.

Chambers JR, Sauer K (2013) Small RNAs and their role in biofilm formation. Trends Microbiol 21: 39–49.

Champagne FA (2008) Epigenetic mechanisms and the transgenerational effects of maternal care. Front Neuroendocrinol 29: 386–397.

Chang HH, Hemberg M, Barahona M, Ingber DE, Huang S (2008) Transcriptome-wide noise controls lineage choice in mammalian progenitor cells. Nature 453: 544–547.

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: 2974–2978.

Changeux JP, Danchin A (1976) Selective stabilisation of developing synapses as a mechanism for the specification of neuronal networks. Nature 264: 705–712.

Chao L, Levin BR (1981) Structured habitats and the evolution of anticompetitor toxins in bacteria. Proc Natl Acad Sci USA 78: 6324–6328.

Chapelle G, Peck LS (1999) Polar gigantism dictated by oxygen availability. Nature 399: 114–115.

Chapin FS III, Bloom AJ, Field CB, Waring RH (1987) Plant responses to multiple environmental factors. BioScience 37: 49–57.

Chapman CA, White FJ, Wrangham RW (1994) Party size in chimpanzees and bonobos: a reevaluation of theory based on two similarly forested sites. In: Wrangham RW, McGrew WC, de Waal FBM, Heltne PG, eds. Chimpanzee cultures. Cambridge, MA: Harvard University Press. pp 41–57.

Chapman CA, Chapman LJ (2000) Determinants of group size in primates: the importance of travel costs. In: Boinski S, Garber PA, eds. On the move: how and why animals travel in groups. Chicago, IL: University of Chicago Press. pp 24–42.

Chapman DF, Robson MJ, Snaydon RW (1992) Physiological integration in the clonal perennial herb Trifolium repens L. Oecologia 89: 338–347.

Chapman NC, Beekman M, Oldroyd BP (2010) Worker reproductive parasitism and drift in the western honeybee Apis mellifera. Behav Ecol Sociobiol 64: 419–427.

Chapman RF (1998) The insects. Structure and function. 4th edn. Cambridge, UK: Cambridge University Press.

Chapman T (2006) Evolutionary conflicts of interest between males and females. Curr Biol 16: R744–R754.

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: 241–244.

Chapman T, Partridge L (1996) Sexual conflict as fuel for evolution. Nature 381: 189–190.

Chapuisat M (2009) Social evolution: the smell of cheating. Curr Biol 19: R196–R198.

Chapuisat M, Keller L (1999) Testing kin selection with sex allocation data in eusocial Hymenoptera. Heredity 82: 473–478.

Charlesworth B (1978) Some models of the evolution of altruistic behaviour between siblings. J Theor Biol 72: 297–319.

Charlesworth B (1979) Note on the evolution of altruism in structured demes. Am Nat 113: 601–605.

Charlesworth B (1994) Evolution in age-structured populations. Cambridge, UK: Cambridge University Press.

Charlton BG (1996) What is the ultimate cause of socio-economic inequalities in health? An explanation in terms of evolutionary psychology. J R Soc Med 89: 3–8.

Charnov EL (1978) Evolution of eusocial behavior: offspring choice or parental parasitism? J Theor Biol 75: 451–465.

Charnov EL, Krebs JR (1975) The evolution of alarm calls: altruism or manipulation? Am Nat 109: 107–112.

Charnov EL, Berrigan D (1993) Why do female primates have such long lifespans and so few babies? Or life in the slow lane. Evol Anthropol 1: 191–194.

Chastanet A, Vitkup D, Yuan GC, Norman TM, Liu JS, Losick RM (2010) Broadly heterogeneous activation of the master regulator for sporulation in Bacillus subtilis. Proc Natl Acad Sci USA 107: 8486–8491.

Chastanet A, Losick R (2011) Just-in-time control of Spo0A synthesis in Bacillus subtilis by multiple regulatory mechanisms. J Bacteriol 193: 6366–6374.

Chaston J, Goodrich-Blair H (2010) Common trends in mutualism revealed by model associations between invertebrates and bacteria. FEMS Microbiol Rev 34: 41–58.

Chattwood A, Thompson CR (2011) Non-genetic heterogeneity and cell fate choice in Dictyostelium discoideum. Dev Growth Differ 53: 558–566.

Chattwood A, Nagayama K, Bolourani P, Harkin L, Kamjoo M, Weeks G, Thompson CRL (2013) Developmental lineage priming in Dictyostelium by heterogeneous Ras activation. eLife 2: e01067.

Chen ICK, Griesenauer B, Yu YTN, Velicer GJ (2014) A recent evolutionary origin of a bacterial small RNA that controls multicellular fruiting body development. Mol Phylogenet Evol 73: 1–9.

Chen L, Wang R, Zhou T, Aihara K (2005) Noise-induced cooperative behavior in a multicell system. Bioinformatics 21: 2722–2729.

Chen X, Fu F, Wang L (2008a) Interaction stochasticity supports cooperation in spatial prisoner’s dilemma. Phys Rev E 78: 051120.

Chen X, Fu F, Wang L (2008b) Promoting cooperation by local contribution under stochastic win-stay-lose-shift mechanism. Physica A 387: 5609–5615.

Chen X, Perc M (2014) Excessive abundance of common resources deters social responsibility. Sci Rep 4: 4161.

Chen Y, Zhu J, Lum PY, Yang X, Pinto S, et al. (2008) Variations in DNA elucidate molecular networks that cause disease. Nature 452: 429–435.

Cheney DL, Seyfarth RM, Fischer J, Beehner J, Bergman T, Johnson SE, et al. (2004) Factors affecting reproduction and mortality among baboons in the Okavango Delta, Botswana. Int J Primatol 25: 401–428.

Cheney DL, Moscovice LR, Heesen M, Mundry R, Seyfarth RM (2010) Contingent cooperation between wild female baboons. Proc Natl Acad Sci USA 107: 9562–9566.

Cheplick GP (1992) Sibling competition in plants. J Ecol 80: 567–575.

Cheplick GP (1993a) Reproductive systems and sibling competition in plants. Plant Species Biol 8: 131–139.

Cheplick GP (1993b) Sibling competition is a consequence of restricted dispersal in an annual cleistogamous grass. Ecology 74: 2161–2164.

Cheplick GP (1996) Do seed germination patterns in cleistogamous annual grasses reduce the risk of sibling competition? J Ecol 84: 247–255.

Cheplick GP, Kane KH (2004) Genetic relatedness and competition in Triplasis purpurea (Poaceae): resource partitioning or kin selection? Int J Plant Sci 165: 623–630.

Chesson P, Huntly N (1997) The roles of harsh and fluctuating conditions in the dynamics of ecological communities. Am Nat 150: 519–553.

Cheverud JM (1996) Developmental integration and the evolution of pleiotropy. Am Zool 36: 44–50.

Chiavegatto S, Nelson RJ (2003) Interaction of nitric oxide and serotonin in aggressive behavior. Horm Behav 44: 233–241.

Chigira M, Noda K, Watanabe H (1990) Autonomy in tumor cell proliferation. Med Hypotheses 32: 249–254.

Chin-Baarstad A, Klug H, Lindström K (2009) Should you eat your offspring before someone else does? Effect of an egg predator on filial cannibalism in the sand goby. Anim Behav 78: 203–208.

Chinnusamy V, Zhu JK (2009) Epigenetic regulation of stress responses in plants. Curr Opin Plant Biol 12: 133–139.

Chittka A, Chittka L (2010) Epigenetics of royalty. PLoS Biol. 8: e1000532.

Chittka A, Wurm Y, Chittka L (2012) Epigenetics: The making of ant castes. Curr Biol 22: R835–R838.

Chittka L, Muller H (2009) Learning, specialization, efficiency and task allocation in social insects. Commun Integr Biol 2: 151–154.

Cho MM, DeVries AC, Williams JR, Carter CS (1999) The effects of oxytocin and vasopressin on partner preferences in male and female prairie voles (Microtus ochrogaster). Behav Neurosci 113: 1071–1079.

Choe JC, Crespi BJ, eds. (1997) The evolution of social behavior in insects and arachnids. Cambridge, UK: Cambridge University Press.

Choe JC, Perlman DL (1997) Social conflict and cooperation among founding queens in ants (Hymenoptera: Formicidae). In: Choe JC, Crespi BJ, eds. The evolution of social behavior in insects and arachnids. Cambridge, UK: Cambridge University Press. pp 392–406.

Choi JK, Bowles S (2007) The coevolution of parochial altruism and war. Science 318: 636–640.

Chou HH, Chiu HC, Delaney NF, Segre D, Marx CJ (2011) Diminishing returns epistasis among beneficial mutations decelerates adaptation. Science 332: 1190–1192.

Choudhury S, Panda P, Sahoo L, Panda SK (2013) Reactive oxygen species signaling in plants under abiotic stress. Plant Signal Behav 8: e23681.

Chown SL (2012) Biotic interactions modify the effects of oxygen on insect gigantism. Proc Natl Acad Sci USA 109: 10745–10746.

Chown SL, Gaston KJ (2010) Body size variation in insects: a macroecological perspective. Biol Rev 85: 139–169.

Christensen KM, Whitham TG, Balda RP (1991) Discrimination among pinyon pine trees by Clark’s nutcrackers: effects of cone crop size and cone characters. Oecologia 86: 402–407.

Chuang JS, Rivoire O, Leibler S (2010) Cooperation and Hamilton's rule in a simple synthetic microbial system. Mol Syst Biol 6: 398.

Cini A, Gioli L, Cervo R (2009) A quantitative threshold for nest-mate recognition in a paper social wasp. Biol Lett doi:10.1098/rsbl.2009.0140.

Civelek M, Lusis AJ (2014) Systems genetics approaches to understand complex traits. Nat Rev Genet 15: 34–48.

Claessen D, Rozen DE, Kuipers OP, Søgaard-Andersen L, van Wezel GP (2014) Bacterial solutions to multicellularity: a tale of biofilms, filaments and fruiting bodies. Nat Rev Microbiol 12: 115–124.

Clapham ME, Karr JA (2012) Environmental and biotic controls on the evolutionary history of insect body size. Proc Natl Acad Sci USA 109: 10927–10930.

Clarin TMA, Ruczy?ski I, Page RA, Siemers BM (2013) Foraging ecology predicts learning performance in insectivorous bats. PLoS ONE 8: e64823.

Clark CW, Mangel M (1986) The evolutionary advantages of group foraging. Theor Popul Biol 30: 45–75.

Clark CW, Yoshimura J (1993) Optimization and ESS analysis for populations in stochastic environments. In: Yoshimura J, Clark CW, eds. Adaptation in stochastic environments. Springer Berlin Heidelberg. pp. 122–131.

Clark CW, Dukas R (1994) Balancing foraging and antipredator demands—an advantage of sociality. Am Nat 144: 542–548.

Clark RM, Fewell JH (2014) Social dynamics drive selection in cooperative associations of ant queens. Behav Ecol 25: 117–123.

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: 4821–4824.

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 227–251.

Clarke FM, Faulkes CG (1997) Dominance and queen succession in captive colonies of the eusocial naked mole-rat, Heterocephalus glaber. Proc R Soc Lond B 264: 993–1000.

Clarke FM, Faulkes CG (2001) Intracolony aggression in the eusocial naked mole-rat, Heterocephalus glaber. Anim Behav 61: 311–324.

Claverys JP, Håvarstein LS (2007) Cannibalism and fratricide: Mechanisms and raisons d’être. Nature 5: 219–229.

Clément JL, Bagnères AG (1998) Nestmate recognition in termites. In: Vander Meer RK, Breed MD, Espelie KE, Winston ML, eds. Pheromone communication in social insects. Boulder, CO: Westview Press. pp 126–155.

Clotfelter CT, Cook PJ (1987) Implicit taxation in lottery finance. Natl Tax J 40: 533–546.

Clotfelter CT, Cook PJ (1989) Selling hope: State lotteries in America. Cambridge, MA: Harvard University Press.

Clotfelter CT, Cook PJ, Edell JA, Moore M (1999) State lotteries at the turn of the century: Report to the national gambling impact study commission. http://govinfo.library.unt.edu/ngis c/reports/lotfinal.pdf.

Clouse R (2001) Some effects of group size on the output of beginning nests of Mischocyttarus mexicanus (Hymenoptera: Vespidae). Fla Entomol 84: 418–425.

Clutton-Brock TH (1975) Feeding behaviour of red colobus and black and white colobus in East Africa. Folia Primatol 23: 165–207.

Clutton-Brock TH, ed. (1988) Reproductive success. Studies of individual variation in contrasting breeding seasons. Chicago, IL: University of Chicago Press.

Clutton-Brock TH (1991) The evolution of parental care. Princeton, NJ: Princeton University Press.

Clutton-Brock TH (1998) Reproductive skew, concessions and limited control. Trends Ecol Evol 13: 288–292.

Clutton-Brock TH (2002) Breeding together: kin selection and mutualism in cooperative vertebrates. Science 296: 69–72.

Clutton-Brock TH (2009a) Cooperation between non-kin in animal societies. Nature 462: 51–57.

Clutton-Brock T (2009b) Structure and function in mammalian societies. Phil Trans R Soc B Biol Sci 364: 3229–3242.

Clutton-Brock TH, Harvey PH (1977) Primate ecology and social organization. J Zool 183: 1–39.

Clutton-Brock TH, Albon SD, Guinness FE (1982) Competition between female relatives in a matrilocal mammal. Nature 300: 178–180.

Clutton-Brock TH, Parker GA (1995) Punishment in animal societies. Nature 373: 209–216.

Clutton-Brock TH, Brotherton PN, Smith R, McIlrath GM, Kansky R, Gaynor D, O’Riain MJ, Skinner JD (1998) Infanticide and expulsion of females in a cooperative mammal. Proc R Soc Lond Ser B 265: 2291–2295.

Clutton-Brock TH, O’Riain MJ, Brotherton PNM, Gaynor D, Kansky R, Griffin AS, et al. (1999) Selfish sentinels in cooperative mammals. Science 284: 1640–1644.

Clutton-Brock TH, Brotherton PNM, O’Riain MJ, Griffin AS, Gaynor D, Sharpe L, Kansky R, Manser M, McIlrath GM (2000) Individual contributions to babysitting in a cooperative mongoose, Suricata suricatta. Proc R Soc Lond B Biol Sci 267: 301–305.

Clutton-Brock TH, Brotherton PNM, O’Riain MJ, Griffin AS, Gaynor D, Kansky R, Sharpe L, McIlrath GM (2001a) Contributions to cooperative rearing in meerkats, Suricata suricatta. Anim Behav 61: 705–710.

Clutton-Brock TH, Brotherton PNM, Russell AF, O’Riain MJ, Gaynor D, Kansky R, Griffin A, et al. (2001b) Cooperation, control, and concession in meerkat groups. Science 291: 478–481.

Clutton-Brock TH, Russell AF, Sharpe LL (2003) Meerkat helpers do not specialize in particular activities. Anim Behav 66: 531–540.

Clutton-Brock TH, Hodge SJ, Spong G, Russell AF, Jordan NR, Bennett NC, Manser MB (2006) Intrasexual competition and sexual selection in cooperative meerkats. Nature 444: 1065–1068.

Clutton-Brock TH, Hodge SJ, Flower T (2008) Group size and the suppression of subordinate reproduction in Kalahari meerkats. Anim Behav 76: 689–700.

Clutton-Brock TH, Hodge SJ, Flower T, Spong GF, Young AJ (2010) Adaptive suppression of subordinate reproduction in cooperative mammals. Am Nat 176: 664–673.

Clutton-Brock TH, Lukas D (2012) The evolution of social philopatry and dispersal in female mammals. Mol Ecol 21: 472–492.

Clutton-Brock TH, Huchard E (2013) Social competition and its consequences in female mammals. J Zool 289: 151–171.

Cnaani J, Robinson GE, Bloch G, Borst D, Hefetz A (2000) The effect of queen–worker conflict on caste determination in the bumblebee, Bombus terrestris. Behav Ecol Sociobiol 47: 346–352.

Cockburn A (1996) Why do so many Australian birds cooperate: social evolution in the Corvida? In: Floyd RB, Sheppard AW, De Barro PJ, eds. Frontiers of population ecology. Melbourne, Australia: CSIROPublishing. pp 451–472.

Cockburn A (1998) Evolution of helping behavior in cooperatively breeding birds. Annu Rev Ecol Syst 29: 141–177.

Cockburn A (2003) Cooperative breeding in Oscine passerines: does sociality inhibit speciation? Proc R Soc B 270: 2207–2214.

Cockburn A, Russell AF (2011) Cooperative breeding: a question of climate? Curr Biol 21: R195–R197.

Cohen J, Stewart I (1994) The collapse of chaos: discovering simplicity in a complex world. NewYork, NY: Viking Penguin.

Cohen S, Janicki-Deverts D (2009) Can we improve our physical health by altering our social networks? Perspect Psychol Sci 4: 375–378.

Cole BJ (1983) Multiple mating and the evolution of social behavior in the Hymenoptera. Behav Ecol Sociobiol 12: 191–201.

Cole B (2009) The ecological setting of social evolution: the demography of ant populations. In: Gadau J, Fewell JH, eds. Organization of insect societies: from genome to sociocomplexity. Cambridge, MA: Harvard University Press. pp 74–104.

Cole BJ, Wiernasz DC (1999) The selective advantage of low relatedness. Science 285: 891–893.

Coleman RA, Browne M, Theobalds T (2004) Aggregation as a defense: limpet tenacity changes in response to simulated predator attack. Ecology 85: 1153–1159.

Colgan P (1983) Comparative social recognition. New York, NY: Wiley.

Collias N, Collias E, Jennrich RI (1994) Dominant red junglefowl (Gallus gallus) hens in an unconfined flock rear the most young over their lifetime. Auk 111: 869–872.

Collie K, Kim SJ, Baker MB (2013) Fitness consequences of sibling egg cannibalism by neonates of the Colorado potato beetle, Leptinotarsa decemlineata. Anim Behav 85: 329–338.

Colman-Lerner A, Gordon A, Serra E, Chin T, Resnekov O, Endy D, et al. (2005) Regulated cell-to-cell variation in a cell-fate decision system. Nature 437: 699–706.

Cologer-Clifford A, Simon NG, Lu SF, Smoluk SA (1997) Serotonin agonist-induced decreases in intermale aggression are dependent on brain region and receptor subtype. Pharmacol Biochem Behav 58: 425–430.

Colwell RK (1981) Group selection is implicated in the evolution of female-biased sex ratios. Nature 290: 401–404.

Comai S, Tau M, Gobbi G (2012) The psychopharmacology of aggressive behavior: a translational approach. Part 1: neurobiology. J Clin Psychopharmacol 32: 83–94.

Comins HN, Hamilton WD, May RM (1980) Evolutionarily stable dispersal strategies. J Theor Biol 82: 205–230.

Connell JH (1983) On the prevalence and relative importance of interspeci®c competition: Evidence from field experiments. Am Nat 122: 661–696.

Connor RC (1986) Pseudo-reciprocity: investing in mutualism. Anim Behav 34: 1562–1566.

Connor RC (1995) Altruism among non-relatives: alternatives to the ‘Prisoner’s Dilemma’. Trends Ecol Evol 10: 84–86.

Connor RC,Curry RL (1995) Helping non-relatives: a role for deceit? Anim Behav 49: 389–393.

Connor RC (2000) Group living in whales and dolphins. In: Mann J, Connor RC, Tyack PL, Whitehead H, eds. Cetacean societies: field studies of dolphins and whales. Chicago, IL: University of Chicago Press. pp 199–218.

Connor RC (2010) Cooperation beyond the dyad: on simple models and a complex society. Phil Trans R Soc B 365: 2687–2697.

Conradt L, Roper TJ (2005) Consensus decision making in animals. Trends Ecol Evol 20: 449–456.

Cook MI, Monaghan P, Burns MD (2000) Effects of short-term hunger and competitive asymmetry on facultative aggression in nestling black guillemots Cepphus grylle. Behav Ecol 11: 282–287.

Cook RE (1983) Clonal plant populations. Am Sci 71: 244–253.

Coolen I (2002) Increasing foraging group size increases scrounger use and reduces searching efficiency in nutmeg mannikins (Lonchura punctulata). Behav Ecol Sociobiol 52: 232–238.

Coolen I, Giraldeau LA, Lavoie M (2001) Head position as an indicator of producer and scrounger tactics in a ground-feeding bird. Anim Behav 61: 895–903.

Coolen I, van Bergen Y, Day RL, Laland KN (2003) Species difference in adaptive use of public information in sticklebacks. Proc R Soc Lond B 270: 2413–2419.

Coolen I, Dangles O, Casas J (2005) Social learning in noncolonial insects? Curr Biol 15: 1931–1935.

Coolen I, Giraldeau LA, Vickery W (2007) Scrounging behavior regulates population dynamics. Oikos 116: 533–539.

Cooper WS (1984) Expected time to extinction and the concept of fundamental fitness. J Theor Biol 107: 603–629.

Cooper WS, Kaplan RH (1982) Adaptive ’coin flipping’: a decision theoretic examination of natural selection for random individual variation. J Theoret Biol 94: 135–151.

Copren KA, Nelson LJ, Vargo EL, Haverty MI (2005) Phylogenetic analyses of mtDNA sequences corroborate taxonomic designations based on cuticular hydrocarbons in subterranean termites. Mol Phylogenet Evol 35: 689–700.

Cordero OX, Wildschutte H, Kirkup B, Proehl S, Ngo L, Hussain F, et al. (2012) Ecological populations of bacteria act as socially cohesive units of antibiotic production and resistance. Science 337: 1228–1231.

Cornforth DM, Foster KR (2013) Competition sensing: the social side of bacterial stress responses. Nat Rev Microbiol 11: 285–293.

Corning PA (1996) The co-operative gene: on the role of synergy in evolution. Evol Theor 11: 183–207.

Corning PA (1997) Holistic Darwinism: “Synergistic selection” and the evolutionary process. J Soc Evol Syst 20: 363–400.

Cornwallis CK, West SA, Davis KE, Griffin AS (2010) Promiscuity and the evolutionary transition to complex societies. Nature 466: 969–972.

Costa JT (2006) The other insect societies. Cambridge, MA: Belknap Press of Harvard University Press.

Costa JT, Fitzgerald TD (1996) Developments in social terminology: semantic battles in a conceptual war. Trends Ecol Evol 11: 285–289.

Costerton JW, Cheng KJ, Geesey GG, Ladd TI, Nickel JC, Dasgupta M, Marrie TJ (1987) Bacterial biofilms in nature and disease. Annu Rev Microbiol 41: 435–464.

Costerton JW, Stewart PS, Greenberg EP (1999) Bacterial biofilms: a common cause of persistent infections. Science 284: 1318–1322.

Coufal NG, Garcia-Perez JL, Peng GE, Yeo GW, Mu Y, Lovci MT, et al. (2009) L1 retrotransposition in human neural progenitor cells. Nature 460: 1127–1131.

Courchamp F, Clutton-Brock T, Grenfell B (1999a) Inverse density dependence and the Allee effect. Trends Ecol Evol 14: 405–410.

Courchamp F, Grenfell B, Clutton-Brock TH (1999b) Population dynamics of obligate cooperators. Proc R Soc Lond Ser B 266: 557–564.

Courchamp F, Macdonald DW (2001) Crucial importance of pack size in the African wild dog Lycaon pictus. Anim Conserv 4: 169–174.

Courtenay SC, Quinn TP, Dupuis HMC, Groot C, Larkin PA (2001) Discrimination of family-specific odours by juvenile coho salmon: roles of learning and odour concentration. J Fish Biol 58: 107–125.

Couzin ID (2009) Collective cognition in animal groups. Trends Cogn Sci 13: 36–43.

Couzin ID, Krause J (2003) Self-organization and collective behavior in vertebrates. Adv Study Behav 32: 1–75.

Couzin ID, Krause J, Franks NR, Levin SA (2005) Effective leadership and decision-making in animal groups on the move. Nature 433: 513–516.

Covas R, Griesser M (2007) Life history and the evolution of family living in birds. Proc R Soc B Biol Sci 274: 1349–1357.

Covas R, du Plessis MA, Doutrelant C (2008) Helpers in colonial cooperatively breeding sociable weavers Philetairus socius contribute to buffer the effects of adverse breeding conditions. Behav Ecol Sociobiol 63: 103–112.

Coveney PV (2003) Self-organization and complexity: a new age for theory, computation and experiment. Phil Trans R Soc Lond A 361: 1057–1079.

Crabbe JC, Wahlsten D, Dudek BC (1999) Genetics of mouse behavior: interactions with laboratory environment. Science 284: 1670–1672.

Craig DM (1982) Group selection versus individual selection: an experimental analysis. Evolution 36: 271–282.

Craig DM, Muir WM (1996) Group selection for adaptation to multiple-hen cages: beak related mortality, feathering, and body weight responses. Poultry Sci 75: 294–302.

Craig R (1979) Parental manipulation, kin selection, and the evolution of altruism. Evolution 33: 319–334.

Craig R (1982) Evolution of male workers in the Hymenoptera. J Theor Biol 94: 95–105.

Creel SR (1990a) How to measure inclusive fitness. Proc R Soc Lond Ser B 241: 229–231.

Creel SR (1990b) The future components of inclusive fitness: acounting for interactions between members of overlapping generations. Anim Behav 40: 127–134.

Creel S (1997) Cooperative hunting and group size: assumptions and currencies. Anim Behav 54: 1319–1324.

Creel SR, Creel NM (1991) Energetics, reproductive suppression and obligate communal breeding in carnivores. Behav Ecol Sociobiol 28: 263–270.

Creel SR, Monfort SL, Wildt DE, Waser PM (1991) Spontaneous lactation is an adaptive result of pseudopregnancy. Nature 351: 660–662.

Creel S, Creel N, Wildt DE, Monfort SL (1992) Behavioural and endocrine mechanisms of reproductive suppression in Serengeti dwarf mongooses. Anim Behav 43: 231–245.

Creel SR, Waser PM (1994) Inclusive fitness and reproductive strategies in dwarf mongooses. Behav Ecol 5: 339–348.

Creel S, Creel NM (1995) Communal hunting and pack size in african wild dogs, Lycaon pictus. Anim Behav 50: 1325–1339.

Creel S, Macdonald DW (1995) Sociality, group size, and reproductive supression among carnivores. Adv Study Behav 24: 203–257.

Creel S, Creel NM, Mills MGL, Monfort SL (1997) Rank and reproduction in cooperatively breeding African wild dogs: behavioral and endocrine correlates. Behav Ecol 8: 298–306.

Creel SR, Waser PM (1997) Variation in reproductive suppression among dwarf mongooses: interplay between mechanisms and evolution. In: Solomon NG, French JA, eds. Cooperative breeding in mammals. Cambridge, UK: Cambridge University Press. pp 150–170.

Creel S, Creel NM (2002) The African wild dog: behavior, ecology and conservation. Princeton, NJ: Princeton University Press.

Creighton JC (2005) Population density, body size, and phenotypic plasticity of brood size in a burying beetle. Behav Ecol 16: 1031–1036.

Cremer S, Armitage SAO, Schmid-Hempel P (2007) Social immunity. Curr Biol 17: R693–R702.

Cremer S, Sixt M (2009) Analogies in the evolution of individual and social immunity. Phil Trans R Soc B Biol Sci 364: 129–142.

Crespi BJ (1992) Cannibalism and trophic eggs in subsocial and eusocial insects. In: Elgar MA, Crespi BJ, eds. Cannibalism: ecology and evolution in diverse taxa. Oxford, UK: Oxford University Press. pp 176–213.

Crespi BJ (1994) Three conditions for the evolution of eusociality–are they sufficient? Insect Soc 41: 395–400.

Crespi BJ (2001) The evolution of social behavior in microorganisms. Trends Ecol Evol 16: 178–183.

Crespi BJ, Yanega D (1995) The definition of eusociality. Behav Ecol 6: 109–115.

Crespi BJ, Choe JC (1997) Explanation and evolution of social systems. In: Choe JC, Crespi BJ, eds. The evolution of social behavior in insects and arachnids. Cambridge, UK: Cambridge University Press. pp 499–524.

Crespi BJ, Ragsdale JE (2000) A skew model for the evolution of sociality via manipulation: why is it better to be feared than loved. Proc R Soc Lond Ser B 267: 821–828.

Crespi B, Summers K (2000) Evolutionary biology of cancer. Trends Ecol Evol 20: 545–552.

Crespi BJ, Morris DC, Mound LA (2004) Evolution of ecological and behavioural diversity: Australian Acacia thrips as model organisms. Australian Biological Resources Study and CSIRO, Canberra.

Cresswell W (1994) Flocking is an effective antipredator strategy in redshanks, Tringa tetanus. Anim Behav 47: 433–442.

Crist TO, MacMahon JA (1991) Foraging patterns of Pogonomyrmex occidentalis (Hymenoptera: Formicidae) in a shrubsteppe ecosystem: the roles of temperature, trunk trails, and seed resources. Environ Entomol 20: 265–275.

Crockett CM, Pope TR (1993) Consequences of sex differences in dispersal for juvenile red howler monkeys. In: Pereira ME, Fairbanks LA, eds. Juvenile primates. New York, NY: Oxford University Press. pp 104–118.

Crockford C, Wittig RM, Langergraber K, Ziegler TE, Zuberbühler K, Deschner T (2013) Urinary oxytocin and social bonding in related and unrelated wild chimpanzees. Proc R Soc B Biol Sci 280: 20122765.

Cronin AL, Hirata M (2003) Social polymorphism in the sweat bee Lasioglossum (Evylaeus) baleicum (Cockerell) (Hymenoptera, Halictidae) in Hokkaido, northern Japan. Insect Soc 50: 379–386.

Cronin AL, Molet M, Doums C, Monnin T, Peeters C (2013) Recurrent evolution of dependent colony foundation across eusocial insects. Annu Rev Entomol 58: 37–55.

Cronin H (1991) The ant and the peacock. Cambridge, UK: Cambridge University Press.

Crook JH (1964) The evolution of social organisation and visual communication in weaverbirds (Ploceinae). Behaviour 10: 1–178.

Crook JH (1965) The adaptive significance of avian social organizations. Symp Zool Soc Lond 14: 181–218.

Crook JH (1970) Social organization and the environment: Aspects of contemporary social ethology. Anim Behav 18: 197–209.

Crook JH (1970) The socio-ecology of primates. In: Cook JH, ed. Social behaviour in birds and mammals. London, UK: Academic Press. pp 103–166.

Crook JH (1972) Sexual selection, dimorphism, and social organization in the primates. In: Campbell BG, ed. Sexual selection and the descent of man. New York, NY: Aldine. pp 231–281.

Cropper M (2012) How should benefits and costs be discounted in an intergenerational context? Resources for the Future Discussion Paper No. 12-42. Available at SSRN: http://ssrn.com/abstract=2165532.

Crosland MWJ (1989) Kin recognition in the ant Rhytidoponera confusa. I. Environmental Odour Anim Behav 37: 912–919.

Crowell S (1957) Differential responses of growth zones to nutritive level, age, and temperature in the colonial hydroid Campanularia. J Exp Zool 134: 63–90.

Crowley PH, Cox JJ (2011) Intraguild mutualism. Trends Ecol Evol 26: 627–633.

Crozier RΗ (1977) Evolutionary genetics of the Hymenoptera. Annu Rev Entomol 22: 263–288.

Crozier RH (1979) Genetics of sociality. In: Herman HR, ed. Social Insects (Vol. 1). New York, NY: Academic Press. pp 223–286.

Crozier RH (1986) Genetic clonal recognition abilities in marine invertebrates must be maintained by selection for something else. Evolution 40: 1100–1101.

Crozier RH (1987) Genetic aspects of kin recognition: concepts, models, and synthesis. In: Fletcher DJC, Michener CD, eds. Kin recognition in animals. New York, NY: Wiley & Sons. pp 55–73.

Crozier RH (2008) Advanced eusociality, kin selection and male haploidy. Aust J Entomol 47: 2–8.

Crozier RH, Dix MW (1979) Analysis of two genetic models for the innate components of colony odor in social Hymenoptera. Behav Ecol Sociobiol 4: 217–224.

Crozier RH, Page RE (1985) On being the right size: male contributions and multiple mating in the social hymenoptera. Behav Ecol Sociobiol 18: 105–115.

Crozier RH, Fjerdingstad EJ (2001) Polyandry in social Hymenoptera - disunity in diversity. Ann Zool Fenn 38: 267–285.

Crozier R, Schlüns H (2008) Genetic caste determination in termites: out of the shade but not from Mars. Bioessays 30: 299–302.

Crump ML (1992) Cannibalism in amphibians. In: Elgar MA, Crespi BJ, eds. Cannibalism: ecology and evolution in diverse taxa. Oxford, UK: Oxford University Press. pp 256–276.

Crump ML (1996) Parental care among the Amphibia. Adv Study Behav 25: 109–144.

Csete ME, Doyle JC (2002) Reverse engineering of biological complexity. Science 295: 1664–1669.

Cullen PJ, Sprague GF (2000) Glucose depletion causes haploid invasive growth in yeast. Proc Natl Acad Sci USA 97: 13461–13463.

Curley JP, Jensen CL, Mashoodh R, Champagne FA (2011) Social influences on neurobiology and behavior: epigenetic effects during development. Psychoneuroendocrinology 36: 352–371.

Currie CR (2001) A community of ants, fungi, and bacteria: a multilateral approach to studying symbiosis. Annu Rev Microbiol 55: 357–380.

Currie CR, Scott JA, Summerbell RC, Malloch D (1999) Fungus-growing ants use antibiotic-producing bacteria to control garden parasites. Nature 398: 701–704.

Curry OS, Price ME, Price JG (2005) Is patience a virtue? Individual differences in discount rates and cooperativeness. Paper presented at the 17th meeting of the Human Behavior and Evolution Society, Austin, Texas. http://www.indiana.edu/~workshop/seminars/papers/y673_fall_2004_curry.pdf

Curry RL (1988a) Influence of kinship on helping behavior in Galapagos mockingbirds. Behav Ecol Sociobiol 22: 141–152.

Curry RL (1988b) Group structure, within-group conflict and reproductive tactics in cooperatively breeding Galapagos mockingbirds, Nesomimus parvulus. Anim Behav 36: 1708–1728.

Cushing JM (1986) Periodic Lotka–Volterra competition equations. J Math Biol 24: 381–403.

Cutts C, Speakman J (1994) Energy savings in formation flight of pink-footed geese. J Exp Biol 189: 251–261.

Cuvillier-Hot V, Cobb M, Malosse C, Peeters C (2001) Sex, age and ovarian activity affect cuticular hydrocarbons in Diacamma ceylonense, a queenless ant. J Insect Physiol 47: 485–493.

Cuvillier-Hot V, Lenoir A, Crewe R, Malosse C, Peeters C (2004a) Fertility signalling and reproductive skew in queenless ants. Anim Behav 68: 1209–1219.

Cuvillier-Hot V, Lenoir A, Peeters C (2004b) Reproductive monopoly enforced by sterile police workers in a queenless ant. Behav Ecol 15: 970–975.

Cuvillier-Hot V, Renault V, Peeters C (2005) Rapid modification in the olfactory signal of ants following a change in reproductive status. Naturwissenschaften 92: 73–77.

Cuvillier-Hot V, Lenoir A (2006) Biogenic amine levels, reproduction and social dominance in the queenless ant Streblognathus peetersi. Naturwissenschaften 93: 149–153.

Cyr AR, Domann FE (2011) The redox basis of epigenetic modifications: from mechanisms to functional consequences. Antioxid Redox Signal 15: 551–589.

Czárán TL, Hoekstra RF, Pagie L (2002) Chemical warfare between microbes promotes biodiversity. Proc Natl Acad Sci USA 99: 786–790.

Czárán T, Hoekstra RF (2009) Microbial communication, cooperation and cheating: quorum sensing drives the evolution of cooperation in bacteria. PLoS ONE 4: e6655.

Czárán T, Hoekstra RF (2010) Janus-headed communication promotes bacterial cooperation and cheating: is quorum sensing useful against infections? Virulence 1: 402–403.

Dabrowska J, Hazra R, Ahern TH, Guo JD, McDonald AJ, Mascagni F, et al. (2011) Neuroanatomical evidence for reciprocal regulation of the corticotrophin-releasing factor and oxytocin systems in the hypothalamus and the bed nucleus of the stria terminalis of the rat: implications for balancing stress and affect. Psychoneuroendocrinology 36: 1312–1326.

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: 473–479.

Dahbi A, Lenoir A (1998) Nest separation and the dynamics of the Gestalt odor in the polydomous ant Cataglyphis iberica (Hymenoptera, Formicidae). Behav Ecol Sociobiol 42: 349–355.

Dall SRX, Johnstone RA (2002) Managing uncertainty: information and insurance under the risk of starvation. Phil Trans R Soc Lond Ser B 357: 1519–1526.

Dallman MF, Akana SF, Strack AM, Hanson ES, Sebastian RJ (1995) The neural network that regulates energy balance is responsive to glucocorticoids and insulin and also regulates HPA axis responsivity at a site proximal to CRF neurons. Ann NY Acad Sci 771: 730–742.

Daly M, Wilson M (2002) Two special issues on risk. Evol Hum Behav 23: 1–2.

Daly M, Wilson M (2005) Carpe diem: Adaptation and devaluing the future. Q Rev Biol 80: 55–60.

Damer TE (1995) Attacking faulty reasoning: A practical guide to fallacy-free arguments. 3rd edn. Belmont, CA: Wadsworth Publishing.

Damore JA, Gore J (2012) Understanding microbial cooperation. J Theor Biol 299: 31–41.

Dandekar A, Chugani S, Greenberg E (2012) Bacterial quorum sensing and metabolic incentives to cooperate. Science 338: 264–266.

Danforth BN (1990) Provisioning behavior and the estimation of investment ratios in a solitary bee, Calliopsis (Hypomacrotera) persimilis (Cockerell) (Hymenoptera: Andrenidae). Behav Ecol Sociobiol 27: 159–168.

Danforth BN (2002) Evolution of sociality in a primitively eusocial lineage of bees. Proc Natl Acad Sci USA 99: 286–290.

Danforth BN (2007) Bees - a primer. Curr Biol 17: R156–R161.

Danforth BN, Neff JL, Barretto-Ko P (1996) Nestmate relatedness in a communal bee, Perdita texana (Hymenoptera: Andrenidae), based on DNA fingerprinting. Evolution 50: 276–284.

Danforth BN, Conway L, Ji S (2003) Phylogeny of eusocial Lasioglossum reveals multiple losses of eusociality within a primitively eusocial clade of bees (Hymenoptera: Halictidae). Syst Biol 52: 23–36.

Dani FR, Jones GR, Destri S, Spencer SH, Turillazzi S (2001) Deciphering the recognition signature within the cuticular chemical profile of paper wasps. Anim Behav 62: 165–171.

Danielson P (2002) Competition among cooperators: Altruism and reciprocity. Proc Natl Acad Sci USA 99 (Suppl. 3): 7237–7242.

Danks HV (1971) Nest mortality factors in stem-nesting aculeate Hymenoptera. J Anim Ecol 40: 79–82.

Dantzer B, Swanson EM (2012) Mediation of vertebrate life histories via insulin-like growth factor-1. Biol Rev 87: 414–429.

Dantzer R (1998) Vasopressin, gonadal steroids and social recognition. Prog Brain Res 119: 409–414.

Dao DN, Kessin RH, Ennis HL (2000) Developmental cheating and the evolutionary biology of Dictyostelium and Myxococcus. Microbiology 146: 1505–1512.

Darwin CR (1859, 6th edn. 1872) On the origin of species. London, UK: John Murray.

Da Silva J, Woodroffe R, Macdonald DW (1993) Habitat, food availability and group territoriality in the European badger. Oecologia 95: 558–564.

Daugherty THF, Toth AL, Robinson GE (2011) Nutrition and division of labor: effects on foraging and brain gene expression in the paper wasp Polistes metricus. Mol Ecol 20: 5337–5347.

Davey ME, O’Toole GA (2000) Microbial biofilms: from ecology to molecular genetics. Microbiol Mol Biol Rev 64: 847–867.

Davies DG, Parsek MR, Pearson JP, Iglewski BH, Costerton JW, Greenberg EP (1998) The involvement of cell-to-cell signals in the development of a bacterial biofilm. Science 280: 295–298.

Davies NB (1982) Behaviour and competition for scarce resources. In: King's College Sociobiology Group, ed. Current problems in sociobiology. Cambridge, UK: Cambridge University Press. pp 363–380.

Davies NB (1991) Mating systems. In: Krebs JR, Davies NB, eds. Behavioural ecology: an evolutionary approach. Oxford, UK: Blackwell Scientific Publications. pp 263–294.

Davies NB (1992) Dunnock behaviour and social evolution. Oxford, UK: Oxford University Press.

Davies NB (2000) Cuckoos, cowbirds and other cheats. London, UK: T & A.D. Poyser.

Davies NB, Lundberg A (1984) Food distribution and a variable mating system in the Dunnock, Prunella modularis. J Anim Ecol 53: 895–912.

Davies NB, Hatchwell BJ, Robson T, Burke T (1992) Paternity and parental effort in dunnocks Prunella modularis: how good are male chick-feeding rules? Anim Behav 43: 729–745.

Davies NB, Krebs JR, West SA (2012) An introduction to behavioural ecology. 4th edn. Chichester, UK: John Wiley & Sons.

Davies NG, Gardner A (2014) Evolution of paternal care in diploid and haplodiploid populations. J Evol Biol 27: 1012–1019.

Davies P, Demetrius LA, Tuszynski JA (2012) Implications of quantum metabolism and natural selection for the origin of cancer cells and tumor progression. AIP Adv 2: 011101.

Davis AR (2011) Kin presence drives philopatry and social aggregation in juvenile Desert Night Lizards (Xantusia vigilis). Behav Ecol 23: 18–24.

Daw ND, Kakade S, Dayan P (2002) Opponent interactions between serotonin and dopamine. Neural Network 15: 603−616.

Dawes RM (1980) Social dilemmas. Annu Rev Psychol 31: 169−193.

Dawkins R (1976) The selfish gene. Oxford, UK: Oxford University Press.

Dawkins R (2012) The descent of Edward Wilson. Prospect Magazine 195: 24th May 2012.

de Almeida RM, Rowlett JK, Cook JM, et al. (2004) GABAA/alpha1 receptor agonists and antagonists: effects on species-typical and heightened aggressive behavior after alcohol self-administration in mice. Psychopharmacology 172: 255−263.

de Almeida RM, Ferrari PF, Parmigiani S, Miczek KA (2005) Escalated aggressive behavior: dopamine, serotonin and GABA. Eur J Pharmacol 526: 51−64.

Debat V, Debelle A, Dworkin I (2009) Plasticity, canalization, and developmental stability of the Drosophila wing: joint effects of mutations and developmental temperature. Evolution 63: 2864–2876.

de Beco S, Ziosi M, Johnston LA (2012) New frontiers in cell competition. Dev Dyn 241: 831–841.

De Beer D, Srinivasan R, Stewart PS (1994) Direct measurement of chlorine penetraion into biofilms during disinfection. Appl Environ Microbiol 60: 4339–4344.

De Boer RJ, Freitas AA, Perelson AS (2001) Resource competition determines selection of B cell repertoires. J Theor Biol 212: 333–343.

de Boer SF, Koolhaas JM (2005) 5-HT1A and 5-HT1B receptor agonists and aggression: A pharmacological challenge of the serotonin deficiency hypothesis. Eur J Pharmacol 526: 125–139.

De Bono M, Tobin DM, Davis MW, Avery L, Bargmann CI (2002) Social feeding in Caenorhabditis elegans is induced by neurons that detect aversive stimuli. Nature 419: 899–903.

Decety J, Jackson PL (2004) The functional architecture of human empathy. Behav Cogn Neurosci Rev 3: 71–100.

Decety J, Jackson PL (2006) A social-neuroscience perspective on empathy. Curr Dir Psychol Sci 15: 54–58.

Decho AW (1994) Exopolymers in microbial mats: assessing their adaptive roles. In: Stal LJ, Caumette P, eds. Microbial mats: structure, development, and environmental significance. Berlin, Germany: Springer-Verlag. pp 215–219.

Declerck CH, Boone C, Kiyonari T (2010) Oxytocin and cooperation under conditions of uncertainty: the modulating role of incentives and social information. Horm Behav 57: 368–374.

De Dreu CKW (2012) Oxytocin modulates cooperation within and competition between groups: An integrative review and research agenda. Horm Behav 61: 419–428.

DeHeer CJ, Ross KG (1997) Lack of detectable nepotism in multiple-queen colonies of the fire ant Solenopsis invicta (Hymenoptera: Formicidae). Behav Ecol Sociobiol 40: 27–33.

DeHeer CJ, Vargo EL (2006) An indirect test of inbreeding depression in the termites Reticulitermes flavipes and Reticulitermes virginicus. Behav Ecol Sociobiol 59: 753–761.

de Kroon H, Huber H, Stuefer JF, van Groenendael JM (2005) A modular concept of phenotypic plasticity in plants. New Phytol 166: 73–82.

de la Cova C, Abril M, Bellosta P, Gallart P, Johnston LA (2004) Drosophila myc regulates organ size by inducing cell competition. Cell 117: 107–116.

DeLay LS, Faaborg J, Naranjo J, Paz SM, de Vrie T, Parker PG (1996) Paternal care in the cooperatively polyandrous Galapagos hawk. Condor 98: 300–311.

Delton AW, Krasnow MM, Leda Cosmides L, Tooby J (2011) Evolution of direct reciprocity under uncertainty can explain human generosity in one-shot encounters. Proc Natl Acad Sci USA 108: 13335–13340.

de Mazancourt C, Dieckmann U (2004) Trade-off geometries and frequency-dependent selection. Am Nat 164: 765–778.

Dempster JP, McLean IFG, eds. (1998) Insect populations: in theory and in practice. London, UK: Kluwer Academic Press.

Deneubourg JL, Goss S (1989) Collective patterns and decision making. Ethol Ecol Evol 1: 295–311.

Deneubourg JL, Lioni A, Detrain C (2002) Dynamics of aggregation and emergence of cooperation. Biol Bull 202: 262–267.

Dengler-Crish CM, Catania KC (2007) Phenotypic plasticity in female naked mole-rats after removal from reproductive suppression. J Exp Biol 210: 4351–4358.

Denison RF (2000) Legume sanctions and the evolution of symbiotic cooperation by rhizobia. Am Nat 156: 567–576.

Denver RJ (2009) Stress hormones mediate environment-genotype interactions during amphibian development. Gen Comp Endocr 164: 20–31.

Deppmann CD, Mihalas S, Sharma N, Lonze BE, Niebur E, Ginty DD (2008) A model for neuronal competition during development. Science 320: 369–373.

De Schutter G, Theraulaz G, Deneubourg JL (2001) Animal–robots collective intelligence. Ann Math Artif Intel 31: 223–238.

Després L, Jaeger N (1999) Evolution of oviposition strategies and speciation in the globeflower flies Chiastocheta spp. (Anthomyiidae). J Evol Biol 12: 822–831.

Detrain C, Deneubourg JL (2006) Self-organized structures in a superorganism: do ants “behave” like molecules? Phys Life Rev 3: 162–187.

d’Ettorre P (2007) Evolution of sociality: you are what you learn. Curr Biol 17: R766–R768.

d’Ettorre P, Mondi N, Lenoir A, Errard C (2002) Blending in with the crowd: social parasites integrate into their host colonies using a flexible chemical signature. Proc R Soc B 269: 1911–1918.

d’Ettorre P, Heinze J, Ratnieks FLW (2004) Worker policing by egg eating in the ponerine ant Pachycondyla inversa. Proc R Soc Lond B 271: 1427–1434.

de Valpine P, Eadie JM (2008) Conspecific brood parasitism and population dynamics. Am Nat 172: 547–562.

De Vos D, De Chial M, Cochez C, Jansen S, Tümmler B, Meyer JM, Cornelis P (2001) Study of pyoverdin type and production by Pseudomonas aeruginosa isolated from cystic fibrosis patients: prevalence of type II pyoverdine isolates and accumulation of pyoverdine-negative mutations. Arch Microbiol 175: 384–388.

DeVries AC, Guptaa T, Cardillo S, Cho M, Carter CS (2002) Corticotropin-releasing factor induces social preferences in male prairie voles. Psychoneuroendocrinology 27: 705–714.

de Vries GJ (2008) Sex differences in vasopressin and oxytocin innervation of the brain. Prog Brain Res 170: 17–27.

Dew RM, Rehan SM, Tierney SM, Chenoweth LB, Schwarz MP (2012) A single origin of large colony size in allodapine bees suggests a threshold event among 50 million years of evolutionary tinkering. Insect Soc 59: 207–214.

Dew RM, Gardner MG, Schwarz MP (2014) The problems of a priori categorisation of agonism and cooperation: circle-tube interactions in two allodapine bees. Ethology 120: 551–562.

de Waal FBM (1982) Chimpanzee politics: power and sex among apes. Baltimore, MD: John Hopkins University Press (reprinted 2007).

de Waal FBM (1992) Coalitions as part of reciprocal relations in the Arnhem chimpanzee colony. In: Harcourt A, de Waal FBM, eds. Coalitions and alliances in humans and other animals. Oxford, UK: Oxford University Press. pp 233–257.

de Waal FBM (1996) Good natured: the origins of right and wrong in humans and other animals. Cambridge, MA: Harvard University Press.

de Waal FBM (1997a) The chimpanzee’s service economy: food for grooming. Evol Hum Behav 18: 375–386.

de Waal FBM (1997b) Bonobo: the forgotten ape. Berkeley, CA: University of California Press.

de Waal FBM (2008) Putting the altruism back into altruism: the evolution of empathy. Annu Rev Psychol 59: 279–300.

de Waal FBM, Davis JM (2003) Capuchin cognitive ecology: cooperation based on projected returns. Neuropsychologia 41: 221–228.

de Waal FBM, Suchak M (2010) Prosocial primates: selfish and unselfish motivations. Phil Trans R Soc B 365: 2711–2722.

de Weerd H, Verbrugge R (2011) Evolution of altruistic punishment in heterogeneous populations. J Theor Biol 290: 88–103.

Dewhirst FE,Chen T, Izard J, Paster BJ, Tanner ACR,Yu WH, et al. (2010) The human oral microbiome. J Bacteriol 192: 5002–5017.

DeWoody JA, Fletcher DE, Wilkins SD, Avise JC (2001) Genetic documentation of filial cannibalism in nature. Proc Natl Acad Sci USA 98: 5090–5092.

Díaz B, Moreno E (2005) The competitive nature of cells. Exp Cell Res 306: 317–322.

Dickins DW, Clark RA (1987) Games theory and siblicide in the kittiwake gull, Rissa tridactyla. J Theor Biol 125: 301–305.

Dickinson JL, Hatchwell BJ (2004) Fitness consequences of helping. In: Koenig WD, Dickinson JL, eds. Ecology and evolution of cooperative breeding in birds. Cambridge, UK: Cambridge University Press. pp 48–66.

Dickinson JL, McGowan A (2005) Winter resource wealth drives delayed dispersal and family-group living in western bluebirds. Proc R Soc B Biol Sci 272: 2423–2428.

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: 354–357.

Diep DB, Håvarstein LS, Nes IF (1995) A bacteriocin-like peptide induces bacteriocin synthesis in Lactobacillus plantarum C11. Mol Microbiol 18: 631–639.

Dierkes P, Heg D, Taborsky M, Skubic E, Achmann R (2005) Genetic relatedness in groups is sex-specific and declines with age of helpers in a cooperatively breeding cichlid. Ecol Lett 8: 968–975.

Dietemann V, Peeters C, Liebig J, Thivet V, Hölldobler B (2003) Cuticular hydrocarbons mediate discrimination of repoductives and nonreproductives in the ant Myrmecia gulosa. Proc Natl Acad Sci USA 100: 10341–10346.

Dietemann V, Liebig J, Hölldobler B, Peeters C (2005) Changes in the cuticular hydrocarbons of incipient reproductives correlate with triggering of worker policing in the bulldog ant Myrmecia gulosa. Behav Ecol Sociobiol 58: 486–496.

Di Fiore A, Rendall D (1994) Evolution of social organization: a reappraisal for primates by using phylogenetic methods. Proc Natl Acad Sci USA 91: 9941–9945.

Digby LJ (1995) Infant care, infanticide, and female reproductive strategies in polygynous groups of common marmosets (Callithrix jacchus). Behav Ecol Sociobiol 37: 51–61.

Digby L (2000) Infanticide by female mammals: implication for the evolution of social systems. In: van Schaik CP, Jansen CH, eds. Infanticide by males and its implications. Cambridge, UK: Cambridge University Press. pp 423–446.

Diggle SP, Gardner A, West SA, Griffin AS (2007a) Evolutionary theory of bacterial quorum sensing: when is a signal not a signal? Phil Trans R Soc Lond B Biol Sci 362: 1241–1249.

Diggle SP, Griffin AS, Campell GS, West SA (2007b) Cooperation and conflict in quorum-sensing bacterial populations. Nature 450: 411–414.

Dilda CL, Mackay TFC (2002) The genetic architecture of Drosophila sensory bristle number. Genetics 162: 1655–1674.

Dillon RJ, Dillon VM (2004) The gut bacteria of insects: Nonpathogenic interactions. Annu Rev Entomol 49: 71–92.

Dixon PA, Milicich MJ, Sugihara G (1999) Episodic fluctuations in larval supply. Science 283: 1528–1530.

Dobata S, Tsuji K (2009) A cheater lineage in a social insect: Implications for the evolution of cooperation in the wild. Commun Integr Biol 2: 67–70.

Dobler R, Kölliker M (2010) Kin-selected siblicide and cannibalism in the European earwig. Behav Ecol 21: 257–263.

Dobler R, Kölliker M (2011) Influence of weight asymmetry and kinship on siblicidal and cannibalistic behaviour in earwigs. Anim Behav 82: 667–672.

Doe CQ, Goodman CS (1985) Early events in insect neurogenesis. II. The role of cell interactions and cell lineage in the determination of neuronal precursor cells. Dev Biol 111: 206–219.

Doebeli M (1996) An explicit genetic model for ecological character displacement. Ecology 77: 510–520.

Doebeli M (2010) Inclusive fitness is just bookkeeping. Nature 467: 661.

Doebeli M, Blarer A, Ackermann M (1997) Population dynamics, demographic stochasticity, and the evolution of cooperation. Proc Natl Acad Sci USA 94: 5167–5171.

Doebeli M, Knowlton N (1998) The evolution of interspecific mutualisms. Proc Natl Acad Sci USA 95: 8676–8680.

Doebeli M, Hauert C, Killingback T (2004) The evolutionary origins of cooperators and defectors. Science 306: 859–862.

Doebeli M, Hauert C (2005) Models of cooperation based on the Prisoner’s Dilemma and the Snowdrift game. Ecol Lett 8: 748–766.

Doebeli M, Hauert C (2006) Limits of Hamilton’s rule. J Evol Biol 19: 1386–1388.

Doligez B, Part T (2008) Estimating fitness consequences of dispersal: a road to ‘know-where’? Non-random dispersal and the underestimation of dispersers’ fitness. J Anim Ecol 77: 1199–1211.

Domes G, Heinrichs M, Glascher J, Buchel C, Braus DF, Herpertz SC (2007a) Oxytocin attenuates amygdala responses to emotional faces regardless of valence. Biol Psychiatry 62: 1187–1190.

Domes G, Heinrichs M, Michel A, Berger C, Herpertz SC (2007b) Oxytocin improves ‘‘mind-reading’’ in humans. Biol Psychiatry 61: 731–733.

Dominey WJ, Blumer LS (1984) Cannibalism of early life stages in fishes. In: Hausfater G, Blaffer Hrdy S, eds. Infanticide. comparative and evolutionary perspectives. New York, NY: Aldine. pp 43–64.

Dominy RG (1992) Aerodynamics of grand prix cars. J Autom Eng 206: 267–274.

Donaldson ZR, Young LJ (2008) Oxytocin, vasopressin, and the neurogenetics of sociality. Science 322: 900–904.

Donaldson ZR, Spiegel L, Young LJ (2010) Central vasopressin V1a receptor activation is independently necessary for both partner preference formation and expression in socially monogamous male prairie voles. Behav Neurosci 124: 159–163.

Doncaster CP, Jackson A, Watson RA (2013) Manipulated into giving: when parasitism drives apparent or incidental altruism. Proc R Soc B 280: 20130108.

Donohue K (2003) The influence of neighbor relatedness on multilevel selection in the Great Lakes sea rocket. Am Nat 162: 77–92.

Douglas AE (2008) Conflict, cheats and the persistence of symbioses. New Phytol 177: 849–858.

Douglas AE (2010) The symbiotic habit. Princeton, NJ: Princeton University Press.

Douglass JL, Wilkens L, Pantazelou E, Moss F (1993) Noise enhancement of information transfer in crayfish mechanoreceptors by stochastic resonance. Nature 365: 337–340.

Dovidio JF, Piliavin JA, Schroeder DA, Penner L (2006) The social psychology of prosocial behavior. Hillsdale, NY: Lawrence Erlbaum.

Dowding P, Chapin FS III, Wielgolaski FE, Kilfeather P (1981) Nutrients in Tundra ecosystems. In: Bliss LC, Heal OW, More JJ, eds. Tundra ecosystems: A comparative analysis. Cambridge, UK: Cambridge University Press. pp 647–683.

Downs SG, Ratnieks FL (1999) Recognition of conspecifics by honeybee guards uses nonheritable cues acquired in the adult stage. Anim Behav 58: 643–648.

Drake A, Fraser D, Weary DM (2008) Parent–offspring resource allocation in domestic pigs. Behav Ecol Sociobiol 62: 309–319.

Dreber A, Rand DG, Fudenberg D, Nowak MA (2008) Winners don’t punish. Nature 452: 348–351.

Dreller C, Fondrk MK, Page RE Jr (1995) Genetic variability affects the behavior of foragers in a feral honeybee colony. Naturwissenschaften 82: 243–245.

Dreyer AP, Shingleton AW (2011) The effect of genetic and environmental variation on genital size in male Drosophila: canalized but developmentally unstable. PLoS ONE 6: e28278.

Driscoll WW, Pepper JW (2010) Theory for the evolution of diffusible external goods. Evolution 64: 2682–2687.

Drummond H (2001) A revaluation of the role of food in broodmate aggression. Anim Behav 61: 517–526.

Drummond H, González E, Osorno JL (1986) Parent-offspring cooperation in the blue-footed boody (Sula nebouxii): social roles in infanticial brood reduction. Behav Ecol Sociobiol 19: 365–372.

Drummond-Barbosa D, Spradling AC (2001) Stem cells and their progeny respond to nutritional changes during Drosophila oogenesis. Dev Biol 231: 265–278.

Du F, Fu F (2011) Partner selection shapes the strategic and topological evolution of cooperation. Dyn Games Appl 1: 354–369.

Du J, Yu FH, Alpert P, Dong M (2009) Arbuscular mycorrhizal fungi reduce effects of physiological integration in Trifolium repens. Ann Bot 104: 335–344.

Du WB, Cao XB, Hu MB, Wang WX (2009) Asymmetric cost in snowdrift game on scale-free networks. EPL 87: 60004.

Duarte A, Weissing FJ, Pen I, Keller L (2011) An evolutionary perspective on self-organized division of labor in social insects. Annu Rev Ecol Evol Syst 42: 91–110.

Dubois F, Morand-Ferron J, Giraldeau LA (2010) Learning in a game context: strategy choice by some keeps learning from evolving in others. Proc R Soc B Biol Sci 277: 3609–3616.

Dubravcic D, van Baalen M and Nizak C (2014) An evolutionarily significant unicellular strategy in response to starvation stress in Dictyostelium social amoebae [v1; ref status: indexed, http://f1000r.es/3hg] F1000Research 3: 133.

Duckworth RA, Kruuk LEB (2009) Evolution of genetic integration between dispersal and colonization ability in a bird. Evolution 63: 968–977.

D'Udine B, Partridge L (1981) Olfactory preferences of inbred mice (Mus musculus) for their own strain and for siblings: effects of strain, sex and cross-fostering. Behaviour 78: 314–324.

Dudley R (1998) Atmospheric oxygen, giant Paleozoic insects and the evolution of aerial locomotor performance. J Exp Biol 201: 1043–1050.

Duffy JE (1996) Eusociality in a coral-reef shrimp. Nature 381: 512–514.

Duffy JE, Morrison CL, Macdonald KS (2002) Colony defence and behavioural differentiations in the eusocial shrimp Synalpheus regalis. Behav Ecol Sociobiol 51: 488–495.

Duffy JE, Macdonald KS (2009) Kin structure, ecology and the evolution of social organization in shrimp: a comparative analysis. Proc R Soc B Biol Sci 277: 575–584.

Duffy WG (1994) Demographics of Lestes disjunctus (Odonata: Zygoptera) in a riverine wetland. Can J Zool 72: 910–917.

Dugatkin LA (1990) N-person games and the evolution of cooperation: a model based on predator inspection behavior in fish. J Theor Biol 142: 123–135.

Dugatkin LA (1997) Cooperation among animals: a modern perspective. Oxford, UK: Oxford University Press.

Dugatkin LA (2002) Animal cooperation among unrelated individuals. Naturwissenschaften 89: 533–541.

Dugatkin LA (2007) Inclusive fitness theory from Darwin to Hamilton. Genetics 176: 1375–1380.

Dugatkin LA, Reeve HK (1994) Behavioral ecology and levels of selection: dissolving the group selection controversy. Adv Stud Behav 23: 101–133.

Dugatkin LA, Sih A (1995) Behavioral ecology and the study of partner choice. Ethology 99: 265–277.

Dugatkin LA, Perlin M, Atlas R (2003) The evolution of group-beneficial traits in the absence of between-group selection. J Theor Biol 220: 67–74.

Dugatkin LA, Perlin M, Lucas JS, Atlas R (2005) Group-beneficial traits, frequency dependent selection and genotypic diversity: an antibiotic resistance paradigm. Proc R Soc Lond B 272: 79–83.

Dukas R (1987) Foraging behavior of three bee species in a natural mimicry system: Female flowers which mimic male flowers in Ecballium elaterium (Cucurbitaceae). Oecologia 74: 256–263.

Dukas R, ed. (1998) Cognitive ecology: the evolutionary ecology of information processing and decision making. Chicago, IL: University of Chicago Press.

Dukas R (2001) Effects of predation risk on pollinators and plants. In: Chittka L, Thomson J, eds. Cognitive ecology of pollination. Cambridge, UK: Cambridge University Press. pp 214–236.

Dukas R, Real LA (1991) Learning foraging tasks by bees: A comparison between social and solitary species. Anim Behav 42: 269–276.

Dumont HJ, Ali AJ (2004) Stage-specific cannibalism and spontaneous cyst hatching in the freshwater fairy shrimp Streptocephalus proboscideus Frauenfeld. Hydrobiologia 524: 103–113.

Dunbar RIM (1988) Primate social systems. Ithaca, NY: Cornell University Press.

Dunbar RIM (1989) Social systems as optimal strategy sets: the costs and benefits of sociality. In: Standen V, Foley R, eds. Comparative socioecology: the behavioural eology of humans and animals. Oxford, UK: Blackwell Scientific. pp 131–149.

Dunbar RIM (1992) Time: a hidden constraint on the behavioural ecology of baboons. Behav Ecol Sociobiol 31: 35–49.

Dunbar RIM, Dunbar EP (1977) Dominance and reproductive success among female gelada baboons. Nature 266: 351–352.

Duncan GE, Moy SS, Perez A, Eddy DM, Zinzow WM, Lieberman JA, et al. (2004) Deficits in sensorimotor gating and tests of social behavior in a genetic model of reduced NMDA receptor function. Behav Brain Res 153: 507–519.

Duncan GE, Inada K, Farrington JS, Koller BH, Moy SS (2009) Neural activation deficits in a mouse genetic model of NMDA receptor hypofunction in tests of social aggression and swim stress. Brain Res 1265: 186–195.

Dunn J, Dunn DW, Strand MR, Hardy ICW (2014) Higher aggression towards closer relatives by soldier larvae in a polyembryonic wasp. Biol Lett 10: 20140229.

Dunn PO, Cockburn A, Mulder RA (1995) Fairy-wren helpers often care for young to which they are unrelated. Proc R Soc London B Biol Sci 259: 339–343.

du Plessis MA (1993) Helping behaviour in cooperatively-breeding green woodhoopoes: selected or unselected trait? Behaviour 127: 49–65.

du Plessis MA, Weathers WW, Koenig WD (1994) Energetic benefits of communal roosting by acorn woodpeckers during the nonbreeding season. Condor 96: 631–637.

du Plessis MA, Siegfried WR, Armstrong AJ (1995) Ecological and life history correlates of cooperative breeding in South African birds. Oecologia 102: 180–188.

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.

Dussutour A, Fourcassié V, Helbing D, Deneubourg JL (2004) Optimal traffic organization in ants under crowded conditions. Nature 428: 70–73.

Dutil JD (1986) Energetic constraints and spawning interval in the anadromous Arctic charr (Salvelinus alpinus). Copeia 1986: 945–955.

Dworkin M (1996) Recent advances in the social and developmental biology of the myxobacteria. Microbiol Mol Biol Rev 60: 70–102.

Earley RL (2010) Social eavesdropping and the evolution of conditional cooperation and cheating strategies. Phil Trans R Soc B 365: 2675–2686.

Early VE, Williams JG (1988) A Dictyostelium prespore-specific gene is transcriptionally repressed by DIF in vitro. Development 103: 519–524.

East ML, Hofer H (2000) Male spotted hyenas (Crocuta crocuta) queue for status in social groups dominated by females. Behav Ecol 12: 558–568.

Ebensperger LA (1998) Strategies and counterstrategies to infanticide in mammals. Biol Rev 73: 321–346.

Ebensperger LA, Wallem PK (2002) Grouping increases the ability of the social rodent, Octodon degus, to detect predators when using exposed microhabitats. Oikos 98: 491–497.

Ebensperger LA, Hayes LD (2008) On the dynamics of rodent social groups. Behav Process 79: 85–92.

Ebner K, Wotjak CT, Landgraf R, Engelmann M (2000) A single social defeat experience selectively stimulates the release of oxytocin, but not vasopressin, within the septal brain area of male rats. Brain Res 872: 87–92.

Ebstein RP, Israel S, Chew SH, Zhong S, Knafo A (2010) Genetics of human social behavior. Neuron 65: 831–844.

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: 13763–13768.

Edgar BA (2006) How flies get their size: genetics meets physiology. Nat Rev Genet 7: 907–916.

Edwards AC, Mackay TFC (2009) Quantitative trait loci for aggressive behavior in Drosophila melanogaster. Genetics 182: 889–897.

Edwards AC, Ayroles JF, Stone EA, Carbone MA, Lyman RF, Mackay TFC (2009a) A transcriptional network associated with natural variation in Drosophila aggressive behavior. Genome Biol 10: R76.

Edwards AC, Zwarts L, Yamamoto A, Callaerts P, Mackay TFC (2009b) Mutations in many genes affect aggressive behaviour in Drosophila melanogaster. BMC Biol 7: 29.

Edwards JP (1987) Caste regulation in the pharaoh's ant Monomorium pharaonis: the influence of queens on the production of new sexual forms. Physiol Entomol 12: 31–39.

Edwards SC, Pratt SC (2009) Rationality in collective decision-making by ant colonies. Proc R Soc Lond B Biol Sci 276: 3655–3661.

Efferson C, Lalive R, Fehr E (2008) The coevolution of cultural groups and ingroup favoritism. Science 321: 1844–1849.

Egas M, Riedl A (2008) The economics of altruistic punishment and the maintenance of cooperation. Proc R Soc B 275: 871–878.

Egashira N, Tanoue A, Matsuda T, Koushi E, Harada S, Takano Y, et al. (2007) Impaired social interaction and reduced anxiety-related behavior in vasopressin V1a receptor knockout mice. Behav Brain Res 178: 123–127.

Eggert F, Müller-Ruchholtz W, Ferstl R (1998) Olfactory cues associated with the major histocompatibility complex. Genetica 104: 191–197.

Eichenberger P, Fujita M, Jensen ST, Conlon EM, Rudner DZ, Wang ST, et al. (2004) The program of gene transcription for a single differentiating cell type during sporulation in Bacillus subtilis. PLoS Biol. 2: e328.

Eickwort GC (1981) Presocial insects. In: Hermann HR, ed. Social Insects, vol. II. London, UK: Academic Press. pp 199–280.

Eickwort GC, Eickwort KR, Eickwort JM, Gordon JM, Eickwort A (1996) Solitary behavior in a high-altitude population of the social sweat bee Halictus rubicundus. Behav Ecol Sociobiol 38: 227–233.

Eickwort KR (1973) Cannibalism and kin selection in Labidomera clivicollis (Coleoptera: Chrysomelidae). Am Nat 107: 452–453.

Eigen M (1996) On the nature of viral quasispecies. Trends Microbiol 4: 216–218.

Eijsink VGH, Axelsson L, Diep DB, Håvarstein LS, Holo H, Nes IF (2002) Production of class II bacteriocins by lactic acid bacteria; an example of biological warfare and communication. Antonie Van Leeuwenhoek 81: 639–654.

Eikenaar C, Richardson DS, Brouwer L, Komdeur J (2007) Parent presence, delayed dispersal, and territory acquisition in the Seychelles warbler. Behav Ecol 18: 874–879.

Eisenberg N, Fabes RA (1998) Prosocial development. In: Damon W, Eisenberg N, eds. Handbook of child psychology, 5th edn., vol. 3. New York, NY: Wiley. pp 701–778.

Eisenegger C, Haushofer J, Fehr E (2011) The role of testosterone in social interaction. Trends Cogn Sci 15: 263–271.

Ekman J (2006) Family living among birds. J Avian Biol 37: 289–298.

Ekman J, Rosander B (1992) Survival enhancement through food sharing: a means for parental control of natal dispersal. Theor Popul Biol 42: 117–129.

Ekman J, Sklepkovych B, Tegelström H (1994) Offspring retention in the Siberian jay (Perisoreus infaustus): the prolonged brood care hypothesis. Behav Ecol 5: 245–253.

Ekman J, Bylin A, Tegelström H (1999) Increased lifetime reproductive success for Siberian jay (Perisoreus infaustus) males with delayed dispersal. Proc R Soc Lond B 266: 911–915.

Ekman J, Bylin A, Tegelström H (2000) Parental nepotism enhances survival of retained offspring in the Siberian jay. Behav Ecol 11: 416–420.

Ekman J, Baglione V, Eggers S, Griesser M (2001a) Delayed dispersal: living under the reign of nepotistic parents. Auk 118: 1–10.

Ekman J, Eggers S, Griesser M, Tegelstrom H (2001b) Queuing for preferred territories: delayed dispersal of Siberian jays. J Anim Ecol 70: 317–324.

Ekman J, Eggers S, Griesser M (2002) Fighting to stay: the role of sibling rivalry for delayed dispersal. Anim Behav 64: 453–459.

Elango N, Hunt BG, Goodisman MA, Soojin VY (2009) DNA methylation is widespread and associated with differential gene expression in castes of the honeybee, Apis mellifera. Proc Natl Acad Sci USA 106: 11206–11211.

Eldar A, Elowitz M (2010) Functional roles for noise in genetic circuits. Nature 467: 167–173.

Eldegard K, Sonerud GA (2009) Female offspring desertion and male-only care increase with natural and experimental increase in food abundance. Proc R Soc Lond Ser B 276: 1713–1721.

Eldholm V, Johnsborg O, Haugen K, Ohnstad HS, Håvarstein LS (2009) Fratricide in Streptococcus pneumoniae: contributions and role of the cell wall hydrolases CbpD, LytA and LytC. Microbiology 155: 2223–2234.

Elgar MA, Crespi BJ, eds. (1992) Cannibalism: ecology and evolution in diverse taxa. Oxford, UK: Oxford University Press.

Elias S, Banin E (2012) Multi-species biofilms: living with friendly neighbors. FEMS Microbiol Rev 36: 990–1004.

Ellermeier CD, Hobbs EC, González-Pastor JE, Losick R (2006) A three-protein signaling pathway governing immunity to a bacterial cannibalism toxin. Cell 124: 549–559.

Ellgaard L, Helenius A (2003) Quality control in the endoplasmic reticulum. Nat Rev Mol Cell Biol 4: 181–191.

Ellis BJ, Figueredo AJ, Brumbach BH, Schlomer GL (2009) Fundamental dimensions of environmental risk: the impact of harsh versus unpredictable environments on the evolution and development of life history strategies. Hum Nat 20: 204–268.

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.

Ellstrand N, Antonovics J (1985) Experimental studies of the evolutionary significance of sexual reproduction. II. A test of the density-dependent selection hypothesis. Evolution 39: 657–666.

El Mouden C, Gardner A (2008) Nice natives and mean migrants: the evolution of dispersal-dependent social behaviour in viscous populations. J Evol Biol 21: 1480–1491.

Else G, Felton J, Stubbs A (1978) The conservation of bees and wasps. Peterborough, UK: Nature Conservancy Council.

Elwood RW (1991) Ethical implications of studies on infanticide and maternal aggression in rodents. Anim Behav 42: 841–849.

Emlen ST (1982a) The evolution of helping. I. An ecological constraints model. Am Nat 119: 29–39.

Emlen ST (1982b) The evolution of helping. II. The role of behavioral conflict. Am Nat 119: 40–53.

Emlen ST (1984) Cooperative breeding in birds and mammals. In: Krebs JR, Davies NB, eds. Behavioural ecology: an evolutionary approach. Boston, MA: Blackwell Scientific Publication. pp 305–339.

Emlen ST (1991) Evolution of cooperative breeding in birds and mammals. In: Krebs JR, Davies NB, eds. Behavioural ecology: an evolutionary approach. 2nd edn. Oxford, UK: Blackwell Scientific. pp 301–337.

Emlen ST (1994) Benefits, constraints and the evolution of the family. Trends Ecol Evol 9: 282–285.

Emlen ST (1995) An evolutionary theory of the family. Proc Natl Acad Sci USA 92: 8092–8099.

Emlen ST (1997) Predicting family dynamics in social vertebrates. In: Krebs JR, Davies NB, eds. Behavioural ecology: an evolutionary approach, 4th edn. Oxford: Blackwell Scientific. pp 228–253.

Emlen ST, Oring LW (1977) Ecology, sexual selection and the evolution of mating systems. Science 197: 215–223.

Emlen ST, Wrege PH (1991) Breeding biology of white-fronted bee-eaters at Nakuru: the influence of helpers on breeder fitness. J Anim Ecol 60: 309–326.

Emlen ST, Wrege PH (1992) Parent-offspring conflict and the recruitment of helpers among bee-eaters. Nature 356: 331–333.

Emlen ST, Reeve HK, Keller L (1998) Reproductive skew: Disentangling concessions from control. Trends Ecol Evol 13: 458–459.

Endler A, Liebig J, Schmitt T, Parker JE, Jones GR, Schreier P, Hölldobler B (2004) Surface hydrocarbons of queen eggs regulate worker reproduction in a social insect. Proc Natl Acad Sci USA 101: 2945–2950.

Engqvist L, Sauer KP (2002) A life-history perspective on strategic mating effort in male scorpionflies. Behav Ecol 13: 632–636.

Ennis HL, Dao DN, Wu MY, Kessin RH (2003) Mutation of the Dictyostelium fbxA gene affects cell-fate decisions and spatial patterning. Protist 154: 419–429.

Ens BJ, Weissing FJ, Drent RH (1995) The despotic distribution and deferred maturity: two sides of the same coin. Am Nat 146: 625–650.

Enver T, Pera M, Peterson C, Andrews PW (2009) Stem cell states, fates, and the rules of attraction. Cell Stem Cell 4: 387–397.

Epple G, Katz Y (1984) Social influences on oestrogen excretion and ovarian cyclicity in saddle back tamarins (Saguinus fuscicollis). Am J Primatol 6: 215–227.

Erckmann WJ (1983) The evolution of polyandry in shorebirds: An evaluation of hypotheses. In: Wasser SK, ed. Social behavior of female vertebrates. New York, NY: Academic Press. pp 113–168.

Erev I, Bornstein G, Galili R (1993) Constructive intergroup competition as a solution to the free rider problem: a field experiment. J Exp Soc Psychol 29: 463–478.

Eriksson A, Lindgren K (2002) Cooperation in an unpredictable environment. In: Standish RK, Abbass HA, Bedau MA, eds. Artificial Life VIII. Cambridge, MA: MIT Press. pp 394–399.

Erlandsson A (1992) Asymmetric interactions in semiaquatic insects. Oecologia 90: 153–157.

Errington PL (1963) Muskrat populations. Ames, IA: Iowa State University Press.

Eshel I (1972) On the neighbor effect and the evolution of altruistic traits. Theor Popul Biol 3: 258–277.

Eshel I, Cavalli-Sforza LL (1982) Assortment of encounters and evolution of cooperativeness. Proc Natl Acad Sci USA 79: 1331–1335.

Esser C, Alberti S, Höhfeld J (2004) Cooperation of molecular chaperones with the ubiquitin/proteasome system. Biochim Biophys Acta 1695: 171–188.

Eungdamrong NJ, Iyengar R (2004) Modeling cell signaling networks. Biol Cell 96: 355–362.

Evans DJ, Brown MRW, Gilbert P (1990) Susceptibility of bacterial biofilms to tobramycin: role of specific growth rate and phase in the division cycle. J Antimicrob Chemother 25: 585–591.

Evans HE (1977) Extrinsic vs. intrinsic factors in the evolution of insect sociality. BioScience 27: 613–617.

Everaerts C, Farine J-P, Cobb M, Ferveur J-F (2010) Drosophila cuticular hydrocarbons revisited: mating status alters cuticular profiles. PLoS ONE 5: e9607.

Ewald PW (1987) Transmission modes and evolution of the parasitism–mutation continuum. Ann NY Acad Sci 503: 295–306.

Ewen JG, Armstrong DP (2000) Male provisioning is negatively correlated with attempted extrapair copulation frequency in the stitchbird (or hihi). Anim Behav 60: 429–433.

Ezard THG, Aze T, Pearson PN, Purvis A (2011) Interplay between changing climate and species’ ecology drives macroevolutionary dynamics. Science 332: 349–351.

Faaborg J, Bednarz JC (1990) Galápagos and Harris’ hawks: divergent causes of sociality in two raptors. In: Stacey PB, Koenig WD, eds. Cooperative breeding in birds. Cambridge, UK: Cambridge University Press. pp 357–383.

Fairbanks LA, Mcguire MT (1987) Mother infant relationships in vervet monkeys—response to new adult males. Int J Primatol 8: 351–366.

Falconer DS (1989) Introduction to Quantitative Genetics. New York, NY: Longmans.

Falk J, Wong JW, Kölliker M, Meunier J (2014) Sibling cooperation in earwig families provides insights into the early evolution of social life. Am Nat 183: 547–557.

Fanelli D, Boomsma JJ, Turillazzi S (2005) Multiple reproductive strategies in a tropical hover wasp. Behav Ecol Sociobiol 58: 190–199.

Fang A, Pierson DL, Koenig DW, Mishra SK, Demain AL (1997) Effect of simulated microgravity and shear stress on microcin B17 production by Escherichia coli and on its excretion into the medium. Appl Environ Microbiol 63: 4090–4092.

Fanshawe JH, FitzGibbon CD (1993) Factors influencing the hunting success of an african wild dog pack. Anim Behav 45: 479–490.

Farr JA (1975) The role of predation in the evolution of social behavior of natural populations of the guppy, Poecilia reticulata (Pisces: Poeciliidae). Evolution 29: 151–158.

Faucon PC, Pardee K, Kumar RM, Li H, Loh Y-H, et al. (2014) Gene networks of fully connected triads with complete auto-activation enable multistability and stepwise stochastic transitions. PLoS ONE 9: e102873.

Faulkes CG, Abbott DH (1997) The physiology of a reproductive dictatorship: regulation of male and female reproduction by a single female in colonies of naked mole rats. In: Solomon NG, French JA, eds. Cooperative breeding in mammals. Cambridge, UK: Cambridge University Press. pp 302–333.

Faulkes CG, Bennett NC, Bruford MW, O’brien HP, Aguilar GH, Jarvis JUM (1997) Ecological constraints drive social evolution in the African mole-rats. Proc R Soc B Biol Sci 264: 1619–1627.

Fedigan LM, Carnegie SD, Jack KM (2008) Predictors of reproductive success in female white-faced capuchins (Cebus capucinus). Am J Phys Anthropol 137: 82–90.

Fedina TY, Kuo T-H, Dreisewerd K, Dierick HA, Yew JY, et al. (2012) Dietary effects on cuticular hydrocarbons and sexual attractiveness in Drosophila. PLoS ONE 7: e49799.

Fehl K, van der Post DJ, Semmann D (2011) Co-evolution of behaviour and social network structure promotes human cooperation. Ecol Lett 14: 546–551.

Fehr E, Schmidt KM (1999) A theory of fairness, competition, and cooperation. Q J Econ 114: 817–868.

Fehr E, Gächter S (2000) Cooperation and punishment in public good experiments. Am Econ Rev 90: 980–994.

Fehr E, Gächter S (2002) Altruistic punishment in humans. Nature 415: 137–140.

Fehr E, Fischbacher U (2003) The nature of human altruism. Nature 425: 785–791.

Fehr E, Fischbacher U (2004) Third-party punishment and social norms. Evol Hum Behav 25: 63–87.

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, MN: University of Minnesota Press. pp 3–16.

Feinberg AP (2007) Phenotypic plasticity and the epigenetics of human disease. Nature 447: 433–440.

Feinberg AP (2008) Epigenetics at the epicenter of modern medicine. J Am Med Assoc 299: 1345–1350.

Fellbaum CR, Gachomo EW, Beesetty Y, Choudhari S, Strahan GD, Pfeffer PE, et al. (2012) Carbon availability triggers fungal nitrogen uptake and transport in arbuscular mycorrhizal symbiosis. Proc Natl Acad Sci USA 109: 2666–2671.

Ferguson JN, Young LJ, Hearn EF, Matzuk MM, Insel TR, Winslow JT (2000) Social amnesia in mice lacking the oxytocin gene. Nat Genet 25: 284–288.

Ferguson JN, Aldag JM, Insel TR, Young LJ (2001) Oxytocin in the medial amygdala is essential for social recognition in the mouse. J Neurosci 21: 8278–8285.

Ferguson JN, Young LJ, Insel TR (2002) The neuroendocrine basis of social recognition. Front Neuroendocrinol 23: 200–224.

Fergusson LA, Winston ML (1988) The influence of wax deprivation on temporal polyethism in honey bee (Apis mellifera L.) colonies. Can J Zool 66: 1997–2001.

Fernandez-Duque E, Valeggia CR, Mendoza SP (2009) The biology of paternal care in human and nonhuman primates. Annu Rev Anthropol 38: 115–130.

Ferrell JE Jr (2002) Self-perpetuating states in signal transduction: Positive feedback, double-negative feedback and bistability. Curr Opin Cell Biol 14: 140–148.

Ferrell JE Jr, Pomerening JR, Kim SY, Trunnell NB, Xiong W, Huang CYF, Machleder EM (2009) Simple, realistic models of complex biological processes: positive feedback and bistability in a cell fate switch and a cell cycle oscillator. FEBS Lett 583: 3999–4005.

Ferriere R, Michod RE (1996) The evolution of cooperation in spatially heterogeneous populations. Am Nat 147: 692–717.

Ferriere R, Bronstein J, Rinaldi S, Law R, Gauduchon M (2002) Cheating and the evolutionary stability of mutualisms. Proc R Soc Lond B Biol Sci 269: 773–780.

Ferriere R, Michod RE (2011) Inclusive fitness in evolution. Nature 471: E6–E8.

Ferris CF, Melloni RH Jr, Koppel G, Perry KW, Fuller RW, Delville Y (1997) Vasopressin/serotonin interactions in the anterior hypothalamus control aggressive behavior in golden hamsters. J Neurosci 17: 4331–4340.

Ferveur JF (2005) Cuticular hydrocarbons: Their evolution and roles in Drosophila pheromonal communication. Behav Genet 35: 279–295.

Ferveur JF, Cobb M (2010) Behavioral and evolutionary roles of cuticular hydrocarbons in Diptera. In: Blomquist GJ, Bagnères AG, eds. Insect hydrocarbons: biology, biochemistry and chemical ecology. Cambridge, UK: Cambridge University Press. pp 325–343.

Fewell JH (2003) Social insect networks. Science 301: 1867–1870.

Fewell JH, Page RE (1999) The emergence of division of labour in forced associations of normally solitary ant queens. Evol Ecol Res 1: 537–548.

Fiegna F, Velicer GJ (2003) Competitive fates of bacterial social parasites: persistence and self-induced extinction of Myxococcus xanthus cheaters. Proc R Soc Lond B 270: 1527–1534.

Fiegna F, Velicer GJ (2005) Exploitative and hierarchical antagonism in a cooperative bacterium. PLoS Biol 3: e370.

Fiegna F, Yuen-Tsu NY, Kadam SV, Velicer GJ (2006) Evolution of an obligate social cheater to a superior cooperator. Nature 441: 310–314.

Field J (2005) The evolution of progressive provisioning. Behav Ecol 16: 770–778.

Field JP, Shreeves G, Sumner S (1999) Group size, queuing and helping decisions in facultatively eusocial hover wasps. Behav Ecol Sociobiol 45: 378–385.

Field J, Shreeves G, Sumner S, Casiraghi M (2000) Insurance-based advantage to helpers in a tropical hover wasp. Nature 404: 869–871.

Field J, Brace S (2004) Pre-social benefits of extended parental care. Nature 428: 650–652.

Field J, Cronin A, Bridge C (2006) Future fitness and helping in social queues. Nature 441: 214–217.

Field J, Cant MA (2009) Social stability and helping in small animal societies. Phil Trans R Soc B 364: 3181–3189.

Filby AL, Paull GC, Hickmore TF, Tyler CR (2010) Unravelling the neurophysiological basis of aggression in a fish model. BMC Genomics 11: 498.

Filotas E, Grant M, Parrott L, Rikvold PA (2010a) The effect of positive interactions on community structure in a multi-species metacommunity model along an environmental gradient. Ecol Model 221: 885–894.

Filotas E, Grant M, Parrott L, Rikvold PA (2010b) Positive interactions and the emergence of community structure in metacommunities. J Theor Biol 266: 419–429.

Finch CE (1990) Longevity, senescence and the genome. Chicago, IL: University of Chicago Press.

Fincke OM, Hadrys H (2001) Unpredictable offspring survivorship in the damselfly, Megaloprepus coerulatus, shapes parental behavior, constrains sexual selection, and challenges traditional fitness estimates. Evolution 55: 762–772.

Finkel T, Holbrook NJ (2000) Oxidants, oxidative stress and the biology of ageing. Nature 408: 239–247.

Firtel RA (1995) Integration of signaling information in controlling cell-fate decisions in Dictyostelium. Genes Dev 9: 1427–1444.

Fischer B, Taborsky B, Dieckmann U (2009) Unexpected patterns of plastic energy allocation in stochastic environments. Am Nat 173: E108–E120.

Fischer MK, Hoffmann KH, Volkl W (2001) Competition for mutualists in an ant–homopteran interaction mediated by hierarchies of ant attendance. Oikos 92: 531–541.

Fischman BJ, Woodard SH, Robinson GE (2011) Molecular evolutionary analyses of insect societies. Proc Natl Acad Sci USA 108(Suppl 2): 10847–10854.

Fish FE (1999) Energetics of swimming and flying in formation. Comments Theor Biol 5: 283–304.

Fisher DO, Lambin X, Yletyinen SM (2009) Experimental translocation of juvenile water voles in a Scottish lowland metapopulation. Popul Ecol 51: 289–295.

Fisher HE (1992) Anatomy of love: the natural history of monogamy, adultery, and divorce. New York, NY: Simon & Schuster.

Fisher HS,Hoekstra HE (2010)Competition drives cooperation among closely related sperm of deer mice.Nature 463: 801–803.

Fisher ML, Gold RE, Vargo EL, Cognato AI (2004) Behavioral and genetic analysis of the colony fusion in Reticulitermes flavipes (Isoptera: Rhinotermitidae). Sociobiology 44: 565–576.

Fisher RA (1918) The correlation between relatives on the supposition of Mendelian inheritance. Trans R Soc Edinburgh 52: 399–433.

Fisher RA (1930) The genetical theory of natural selection. Oxford, UK: Clarendon Press.

Fisher RM, Cornwallis CK, West SA (2013) Group formation, relatedness, and the evolution of multicellularity. Curr Biol 23: 1120–1125.

Fishman MA (2003) Indirect reciprocity among imperfect individuals. J Theor Biol 225: 285–292.

Fishman MA (2006) Involuntary defection and the evolutionary origins of empathy. J Theor Biol 242: 873–879.

Fishman MA, Lotem A, Stone L (2001) Heterogeneity stabilizes reciprocal altruism interactions. J Theor Biol 209: 87–95.

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.

Fittkau EJ, Klinge H (1973) On biomass and trophic structure of the central Amazonian rain forest ecosystem. Biotropica 5: 2–14.

FitzGerald GJ (1992) Filial cannibalism in fishes – why do parents eat their offspring? Trends Ecol Evol 7: 7–10.

Fitzgerald GJ, Whoriskey FG (1992) Empirical studies of cannibalism in fish. In: Elgar MA, Crespi BJ, eds. Cannibalism: ecology and evolution in diverse taxa. Oxford, UK: Oxford University Press. pp 238–255.

Fitzpatrick JW, Woolfenden GE, Kopeny MT (1991) Ecology and development-related habitat requirements of the Florida scrub jay (Aphelcoma coerulescens coerulescens). Tallahassee, FL: Office of Environmental Services.

Fitzpatrick MJ, Ben-Shahar Y, Smid HM, Vet LEM, Robinson GE, Sokolowski MB (2005) Candidate genes for behavioural ecology. Trends Ecol Evol 20: 96–104.

Flanagan TP, Letendre K, Burnside WR, Fricke GM, Moses ME (2012) Quantifying the effect of colony size and food distribution on harvester ant foraging. PLoS ONE 7: e39427.

Flemming HC, Neu TR, Wozniak D (2007) The EPS matrix: the ‘house of biofilm cells’. J Bacteriol 189: 7945–7947.

Fletcher DJC (1987) The behavioral analysis of kin recognition: Perspectives on methodology and interpretation. In: Fletcher DJC, Michener CD, eds. Kin recognition in animals. New York, NY: Wiley. pp 19–54.

Fletcher DJC, Ross KG (1985) Regulation of reproduction in eusocial Hymenoptera. Annu Rev Entomol 30: 319–343.

Fletcher DJC, Michener CD, eds. (1987) Kin recognition in animals. New York, NY: John Wiley.

Fletcher JA, Zwick M (2004) Strong altruism can evolve in randomly formed groups. J Theor Biol 228: 303–313.

Fletcher JA, Zwick M (2006) Unifying the theories of inclusive fitness and reciprocal altruism. Am Nat 168: 252–262.

Fletcher JA, Doebeli M (2009) A simple and general explanation for the evolution of altruism. Proc R Soc B: Biol Sci 276: 13–19.

Flinn MV (1988) Step- and genetic parent/offspring relationships in a Caribbean village. Ethnol Sociobiol 9: 335−369.

Flinn MV, Nepomnaschy PA, Muehlenbein MP, Ponzi D (2011) Evolutionary functions of early social modulation of hypothalamic-pituitary-adrenal axis development in humans. Neurosci Biobehav Rev 35: 1611–1629.

Florane CB, Bland JM, Husseneder C, Raina AK (2004) Diet-mediated inter-colonial aggression in the Formosan subterranean termite Coptotermes formosanus. J Chem Ecol 30: 2559–2574.

Flores M, Morales L, Gonzaga-Jauregui C, Domínguez-Vidaña R, Zepeda C, Yañez O, et al. (2007) Recurrent DNA inversion rearrangements in the human genome. Proc Natl Acad Sci USA 104: 6099–6106.

Flynn KM, Cooper TF, Moore FB-G, Cooper VS (2013) The environment affects epistatic interactions to alter the topology of an empirical fitness landscape. PLoS Genet 9: e1003426.

Foitzik S, Heinze J (1998) Nest site limitation and colony takeover in the ant Leptothorax nylanderi. Behav Ecol 9: 367–375.

Foitzik S, Heinze J (2000) Intraspecific parasitism and split sex ratios in a monogynous and monandrous ant (Leptothorax nylanderi). Behav Ecol Sociobiol 47: 424–431.

Foitzik S, Fischer B, Heinze J (2003) Arms races between social parasites and their hosts: geographic patterns of manipulation and resistance. Behav Ecol 14: 80–88.

Folkvord A (1997) Ontogeny of cannibalism in larval and juvenile fishes with special emphasis on Atlantic cod. In: Chambers RC, Trippel EA, eds. Early life history and recruitment in fish populations. London, UK: Chapman & Hall. pp 251–278.

Folse HJ III (2011) Evolution and individuality; beyond the genetically homogeneous organism. PhD thesis. Stanford, CA: Stanford University.

Forbes LS (1993) Avian brood reduction and parent-offspring “conflict”. Am Nat 142: 82–117.

Forbes S (2009) Portfolio theory and how parent birds manage investment risk. Oikos 118: 1561–1569.

Forbes LS, Ydenberg RC (1992) Sibling rivalry in a variable environment. Theor Popul Biol 41: 135–160.

Forbes S, Glassey B, Thornton S, Earle L (2001) The secondary adjustment of clutch size in red-winged blackbirds (Agelaius phoeniceus). Behav Ecol Sociobiol 50: 37–44.

Forbes S, Grosshans R, Glassey B (2002a) Multiple incentives for parental optimism and brood reduction in blackbirds. Ecology 83: 2529–2541.

Forbes SH, Adam RMM, Bitney C, Kukuk PF (2002b) Extended parental care in communal social groups. J Insect Sci 2: 22–28.

Foret S, Kucharski R, Pellegrini M, Feng SH, Jacobsen SE, Robinson GE, Maleszka R (2012) DNA methylation dynamics, metabolic fluxes, gene splicing, and alternative phenotypes in honey bees. Proc Natl Acad Sci USA 109: 4968–4973.

Fortin A, Stevenson MM, Gros P (2002) Susceptibility to malaria as a complex trait: big pressure from a tiny creature. Hum Mol Genet 11: 2469–2478.

Fortunato A, Queller DC, Strassmann JE (2003) A linear dominance hierarchy among clones in chimeras of the social amoeba Dictyostelium discoideum. J Evol Biol 16: 438–445.

Foster KR (2004) Diminishing returns in social evolution: the not-so-tragic commons. J Evol Biol 17: 1058–1072.

Foster KR (2011) The sociobiology of molecular systems. Nat Rev Genet 12: 193–203.

Foster KR, Ratnieks FLW (2000) Social insects - Facultative worker policing in a wasp. Nature 407: 692–693.

Foster KR, Ratnieks FLW, Raybould AF (2000) Do hornets have zombie workers? Mol Ecol 9: 735–742.

Foster KR, Ratnieks FLW (2001) Convergent evolution of worker policing by egg eating in the honeybee and common wasp. Proc R Soc Lond B 268: 169–174.

Foster KR, Gulliver J, Ratnieks FLW (2002) Worker policing in the European hornet Vespa crabro. Insect Soc 49: 41–44.

Foster KR, Shaulsky G, Strassmann JE, Queller DC, Thompson CR (2004) Pleiotropy as a mechanism to stabilize cooperation. Nature 431: 693–696.

Foster KR, Ratnieks FLW (2005) A new eusocial vertebrate? Trends Ecol Evol 20: 363–364.

Foster KR, Kokko H (2006) Cheating can stabilize cooperation in mutualisms. Proc Biol Sci 273: 2233–2239.

Foster KR, Wenseleers T (2006) A general model for the evolution of mutualisms. J Evol Biol 19: 1283–1293.

Foster KR, Wenseleers T, Ratnieks FLW, Queller DC (2006) There is nothing wrong with inclusive fitness. Trends Ecol Evol 21: 599–600.

Foster KR, Xavier JB (2007) Cooperation: bridging ecology and sociobiology. Curr Biol 17: R319–R321.

Foster KR, Parkinson K, Thompson CR (2007) What can microbial genetics teach sociobiology? Trends Genet 23: 74–80.

Foster MS, Delay LS (1998) Dispersal of mimetic seeds of three species of Ormosia (Leguminosae). J Trop Ecol 14: 389–411.

Foster WA, Benton TG (1992) Sex-ratio, local mate competition and mating-behavior in the aphid Pemphigus spyrothecae. Behav Ecol Sociobiol 30: 297–307.

Foster WA, Northcott PA (1994) Galls and the evolution of social behaviour in aphids. In: Williams MAJ, ed. Plant galls: organisms, interactions, populations. Oxford, UK: Clarendon Press. pp 161–182.

Fournier D, Battaille G, Timmermans I, Aron S (2008) Genetic diversity, worker size polymorphism and division of labour in the polyandrous ant Cataglyphis cursor. Anim Behav 75: 151–158.

Fowler K, Partridge L (1986) Variation in male fertility explains an apparent effect of genotypic diversity on success in larval competition in Drosophila melanogaster. Heredity 57: 31–36.

Francis DD, Young LJ, Meaney MJ, Insel TR (2002) Naturally occurring differences in maternal care are associated with the expression of oxytocin and vasopressin (V1a) receptors: gender differences. J Neuroendocrinol 14: 349–353.

Frank LG, Glickman SE, Licht P (1991) Fatal sibling aggression, precocial development, and androgens in neonatal spotted hyaenas. Science 252: 702–704.

Frank RH (1988) Passions within reason: The strategic role of the emotions. New York, NY: W. W. Norton and Company.

Frank SA (1986) Dispersal polymorphisms in subdivided populations. J Theor Biol 122: 303–309.

Frank SA (1994) Genetics of mutualism – the evolution of altruism between species. J Theor Biol 170: 393–400.

Frank SA (1995) Mutual policing and repression of competition in the evolution of cooperative groups. Nature 377: 520–522.

Frank SA (1996) Models of parasite virulence. Q Rev Biol 71: 37–78.

Frank SA (1998) Foundations of social evolution. Princeton, NJ: Princeton University Press.

Frank SA (2003) 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) Somatic evolutionary genomics: mutations during development cause highly variable genetic mosaicism with risk of cancer and neurodegeneration. Proc Natl Acad Sci USA 107(suppl 1): 1725–1730.

Frank SA (2013) A new theory of cooperation. In: Summers K, Crespi B, eds. Human social evolution: the foundational works of Richard D. Alexander. Oxford, UK: Oxford University Press. pp 40–47.

Frank SA (2014) Microbial metabolism: optimal control of uptake versus synthesis. PeerJ 2: e267.

Frank SA, Nowak MA (2004) Problems of somatic mutation and cancer. BioEssays 26: 291–299.

Franks NR, Scovell E (1983) Dominance and reproductive success among slave-making worker ants. Nature 304: 724–725.

Fraser D, Kærn M (2009) A chance at survival: gene expression noise and phenotypic diversification strategies. Mol Microbiol 71: 1333–1340.

Fraser JA, Heitman J (2003) Fungal mating-type loci. Curr Biol 13: R792–795.

Fraser JA, Heitman J (2004) Evolution of fungal sex chromosomes. Mol Microbiol 51: 299–306.

Frazier MR, Woods HA, Harrison JF (2001) Interactive effects of rearing temperature and oxygen on the development of Drosophila melanogaster. Physiol Biochem Zool 74: 641–650.

Frederick S, Loewenstein G, O'Donoghue T (2002) Time discounting and time preference: A critical review. J Econ Lit 40: 351–401.

Frederickson ME (2013) Rethinking mutualism stability: cheaters and the evolution of sanctions. Q Rev Biol 88: 269–295.

Free JB (1958) The drifting of honey-bees. J Agric Sci 51: 294–306.

Free JB, Weinberg I, Whiten A (1969) The egg-eating behavior of Bombus lapidaries L. Behaviour 35: 313–317.

Freeland WJ (1976) Pathogens and the evolution of primate sociality. Biotropica 8: 12–24.

Fréon P, Misund OA (1999) Dynamics of pelagic fish distribution and behaviour: effects on fisheries and stock assessment. Oxford, UK: Blackwell Science.

Frère CH, Krützen M, Mann J, Connor RC, Bejder L, Sherwin WB (2010) Social and genetic interactions drive fitness variation in a free-living dolphin population. Proc Natl Acad Sci USA 107: 19949–19954.

Frick T, Schuster S (2003) An example of the prisoner’s dilemma in biochemistry. Naturwissenschaften 90: 327–331.

Friend LA, Bourke AF (2012) Absence of within-colony kin discrimination in a multiple-queen ant, Leptothorax acervorum. Ethology 118: 1182–1190.

Frignani N, Ponti G (2008) Social vs. risk preferences under the veil of ignorance. Universita Degli Studi Di Ferrara, Quaderno No. 8/2008.

Fritts SH, Mech LD (1981) Dynamics, movements, and feeding ecology of a newly protected wolf population in Northwestern Minnesota. Wildlife Monogr 80: 6–79.

Frommen JG, Brendler C, Bakker TCM (2007) The tale of the bad stepfather: male three-spined sticklebacks Gasterosteus aculeatus L. recognize foreign eggs in their manipulated nest by egg cues alone. J Fish Biol 70: 1295–1301.

Fronhofer EA, Pasurka H, Mitesser O, Poethke HJ (2011) Scarce resources, risk sensitivity, and egalitarian resource sharing. Evol Ecol Res 13: 253–267.

Fruehauf JP, Meyskens FL Jr (2007) Reactive oxygen species: a breath of life or death? Clin Cancer Res 13: 789–794.

Frumhoff PC, Baker J (1988) A genetic component to division of labour within honey bee colonies. Nature 333: 358–361.

Fryxell JM, Mosser A, Sinclair AR, Packer C (2007) Group formation stabilizes predator–prey dynamics. Nature 449: 1041–1043.

Fu F, Wang L (2008) Coevolutionary dynamics of opinions and networks: From diversity to uniformity. Phys Rev E 78: 016104.

Fu F, Nowak MA, Hauert C (2010) Invasion and expansion of cooperators in lattice populations: Prisoner’s dilemma vs. snowdrift games. J Theor Biol 266: 358–366.

Fuchs S, Moritz RFA (1999) Evolution of extreme polyandry in the honeybee Apis mellifera L. Behav Ecol Sociobiol 45: 269–275.

Fudenberg D, Maskin E (1990) Evolution and cooperation in noisy repeated games. Am Econ Rev 80: 274–279.

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: 397–399.

Fuller RW (1996) The influence of fluoxetine on aggressive behavior. Neuropsychopharmacology 14: 77–81.

Furuichi T (1997) Agonistic interactions and matrifocal dominance rank of wild bonobos at Wamba. Int J Primatol 18: 855–875.

Futuyma DJ, Slatkin M, eds. (1983) Coevolution. Sunderland, MA: Sinauer Associates.

Gabor CS, Phan A, Clipperton-Allen AE, Kavaliers M, Choleris E (2012). Interplay of oxytocin, vasopressin, and sex hormones in the regulation of social recognition. Behav Neurosci 126: 97–109.

Gächter S, Renner E, Sefton M (2008) The long-run benefits of punishment. Science 322: 1510.

Gadagkar R (1985) Kin recognition in social insects and other animals–A review of recent findings and a consideration of their relevance for the theory of kin selection. Proc Indian Acad Sci 94: 587–621.

Gadagkar R (1990a) Social biology of Ropalidia marginata: Investigations into the origins of eusociality. In: Veeresh GK, Mallik B, Viraktmath CA, eds. Social insects and the environment. Proc. of the 11th International Congress of IUSSI, Bangalore, India, August 1990. New Delhi, India: Oxford and IBH. pp 9–11.

Gadagkar R (1990b) Evolution of insect societies: Some insights from studying tropical wasps. In: Veeresh GK, Kumar ARV, Shivshankar T, eds. Social insects: An Indian perspective. Bangalore, India: IUSSI-Indian Chapter. pp 129–152.

Gadagkar R (1990c) Evolution of eusociality: the advantage of assured fitness returns. Phil Trans R Soc Lond B 329: 17–25.

Gadagkar R (1990d) Origin and evolution of eusociality: A perspective from studying primitively eusocial wasps. J Genet 69: 113–125.

Gadagkar R (1991a) Belonogaster, Mischocyttarus, Parapolybia and independent founding Ropalidia. In: Ross KG, Matthews RW, eds. Social biology of wasps. Ithaca, NY: Cornell University Press. pp 149–190.

Gadagkar R (1991b) Demographic predisposition to the evolution of eusociality: a hierarchy of models. Proc Natl Acad Sci USA 88: 10993–10997.

Gadagkar R (1994a) Why the definition of eusociality is not helpful to understand its evolution and what should we do about it. Oikos 70: 485–488.

Gadagkar R (1994b) The evolution of altruism in insects-a case study. In: Agrawal OP, ed. Perspectives in entomological research. Jodhpur, India: Scientific Publishers. pp 263–275.

Gadagkar R (1996) The evolution of eusociality, including a review of the social status of Ropalidia marginata. In: Turillazzi S, West-Eberhard MJ, eds. Natural history and evolution of paper-wasps. Oxford, UK: Oxford University Press. pp 248–271.

Gadagkar R (2001) The social biology of Ropalidia marginata: toward understanding the evolution of eusociality. Harvard University Press.

Gadagkar K (2004a) Why do honey bee workers destroy each other’s eggs? J Biosci 29: 213–217.

Gadagkar R (2004b) Genetically engineered monogamy in voles lends credence to the modus operandi of behavioural ecology. J Genet 83: 109–111.

Gadagkar R (2010) Sociobiology in turmoil again. Curr Sci 99: 1036–1041.

Gadagkar R, Joshi NV (1982) Behaviour of the Indian social wasp Ropalidia cyathiformis on a nest of separate combs (Hymenoptera: Vespidae). J Zool 198: 27–37.

Gadagkar R, Joshi NV (1983) Quantitative ethology of social wasps: Time-activity budgets and caste differentiation in Ropalidia marginata (Lep.) (Hymenoptera: Vespidae). Anim Behav 31: 26–31.

Gadagkar R, Joshi NV (1984) Social organisation in the Indian wasp Ropalidia cyathiformis (Fab.) (Hymenoptera: Vespidae). Z Tierpsychol 64: 15–32.

Gadagkar R, Joshi NV (1985) Colony fission in a social wasp. Curr Sci 54: 57–62.

Gadagkar R, Chandrashekara K, Chandran S, Bhagavan S (1991) Worker-brood genetic relatedness in a primitively eusocial wasp. Naturwissenschaften 78: 523–526.

Gadagkar R, Chandrashekara K, Chandran S, Bhagavan S (1993) Serial polygyny in the primitively eusocial wasp Ropalidia marginata: implications for the evolution of sociality. In: Keller L, ed. Queen number and sociality in insects. Oxford, UK: Oxford University Press. pp 189–214.

Gadagkar R, Bonner JT (1994) Social insects and social amoebae. J Biosci 19: 219–245.

Gadau J, Gerloff CU, Kruger N, Chan H, Schmid-Hempel P, Wille A, Page RE Jr (2001) A linkage analysis of sex determination in Bombus terrestris (L.) (Hymenoptera: Apidae). Heredity 87: 234–242.

Gadgil M (1971) Dispersal: population consequences and evolution. Ecology 52: 253–261.

Gagne I, Coderre D, Mauffette Y (2002) Egg cannibalism by Coleomegilla maculata lengi neonates: preference even in the presence of essential prey. Ecol Entomol 27: 285–291.

Galef BG, Laland KN (2005) Social learning in animals: empirical studies and theoretical models. BioScience 55: 489–499.

Galizia CG, Rössler W (2010) Parallel olfactory systems in insects: anatomy and function. Annu Rev Entomol 55: 399–420.

Gallucci F, Ólafsson E (2007) Cannibalistic behaviour of rock-pool copepods: an experimental approach for space, food and kinship. J Exp Mar Biol Ecol 342: 325–331.

Gamboa GJ (1978) Intraspecific defense: advantage of social cooperation among paper wasp foundresses. Science 199: 1463–1465.

Gamboa GJ, Reeve HK, Pfennig DW (1986) The evolution and ontogeny of nestmate recognition in social wasps. Annu Rev Entomol 31: 431–454.

Gammie SC, Lonstein JS (2005) Maternal aggression. In: Nelson RJ, ed. Biology of aggression. New York, NY: Oxford University Press. pp 250–274.

Gammie SC, Auger AP, Jessen HM, Vanzo RJ, Awad TA, Stevenson SA (2007) Altered gene expression in mice selected for high maternal aggression. Genes Brain Behav 6: 432–443.

Gandolfi AE, Sachko Gandolfi A, Barash DP (2002) Economics as an evolutionary science: from utility to fitness. New Brunswick, NJ: Transaction Publishers.

Gandon S, Michalakis Y (1999) Evolutionarily stable dispersal rate in a metapopulation with extinctions and kin competition. J Theor Biol 199: 275–290.

Gao X, Lin TC (2011) Do individual investors trade stocks as gambling? Evidence from repeated natural experiments. Unpublished working paper, University of Hong Kong.

Garay J (2009) Cooperation in defence against a predator. J Theor Biol 257: 45–51.

Gardner A, West SA (2006) Demography, altruism, and the benefits of budding. J Evol Biol 19: 1707–1716.

Gardner A, West SA (2007) Social evolution: the decline and fall of genetic kin recognition. Curr Biol 17: R810–R812.

Gardner A, Kümmerli R (2008) Social evolution: this microbe will self-destruct. Curr Biol 18: R1021–R1023.

Gardner A, Grafen A (2009) Capturing the superorganism: a formal theory of group adaptation. J Evol Biol 22: 659–671.

Gardner A, West SA (2010) Greenbeards. Evolution 64: 25–38.

Gardner A, West SA, Wild G (2011) The genetical theory of kin selection. J Evol Biol 24: 1020–1043.

Gardner A, Alpedrinha J, West SA (2012) Haplodiploidy and the evolution of eusociality: split sex ratios. Am Nat 179: 240–256 (correction, Am Nat 179: 554–555).

Garlaschelli D, Capocci A, Caldarelli G (2007) Self-organized network evolution coupled to extremal dynamics. Nature Phys 3: 813–817.

Garnier S, Gautrais J, Theraulaz G (2007) The biological principles of swarm intelligence. Swarm Intell 1: 3–31.

Garrido MC, Herrero M, Kolter R, Moreno F (1988) The export of the DNA replication inhibitor microcin B17 provides immunity for the host cell. EMBO J 7: 1853–1862.

Garrigan D, Hedrick PW (2003) Perspective: detecting adaptive molecular polymorphism: lessons from the MHC. Evolution 57: 1707–1722.

Gaston AJ (1978) The evolution of group territorial behavior and cooperative breeding. Am Nat 112: 1091–1100.

Gaudeul A (2013) Social preferences under uncertainty. Jena Economic Research Papers 2013-024. Friedrich-Schiller-University Jena, Max-Planck-Institute of Economics.

Gaudieri S, Dawkins RL, Habara K, Kulski JK, Gojobori T (2000) SNP profile within the human major histocompatibility complex reveals an extreme and interrupted level of nucleotide diversity. Genome Res 10: 1579–1586.

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: 102–115.

Gaume L, Shenoy M, Zacharias M, Borges RM (2006) Co-existence of ants and an arboreal earthworm in a myrmecophyte of the Indian Western Ghats: Anti-predation effect of the earthworm mucus. J Trop Ecol 22: 341–344.

Gautvik KM, De Lecea L, Gautvik VT, Danielson PE, Tranque P, Dopazo A, et al. (1996) Overview of the most prevalent hypothalamus-specific mRNAs, as identified by directional tag PCR subtraction. Proc Natl Acad Sci USA 93: 8733–8738.

Gavrilets S (1997) Coevolutionary chase in exploiter–victim systems with polygenic characters. J Theor Biol 186: 527–534.

Gavrilets S (2000) Rapid evolution of reproductive barriers driven by sexual conflict. Nature 403: 886–889.

Gavrilets S (2004) Fitness landscapes and the origin of species. Princeton, NJ: Princeton University Press.

Gayou DC (1986) The social system of the Texas green jay. Auk 103: 560–547.

Gazda SK, Connor RC, Edgar RK, Cox F (2005) A division of labour with role specialization in group-hunting bottlenose dolphins (Tursiops truncatus) off Cedar Key, Florida. Proc R Soc B 272: 135–140.

Geier H, Mostowy S, Cangelosi GA, Behr MA, Ford TE (2008) Autoinducer-2 triggers the oxidative stress response in Mycobacterium avium, leading to biofilm formation. Appl Environ Microbiol 74: 1798–1804.

Genini J, Morellato LPC, Guimaraes PR Jr, Olesen JM (2010) Cheaters in mutualism networks. Biol Lett 6: 494–497.

Genova ML, Bianchi C, Lenaz G (2003) Structural organization of the mitochondrial respiratory chain. Ital J Biochem 52: 58–61.

Gerisch G (1968) Cell aggregation and differentiation in Dictyostelium. Curr Top Dev Biol 3: 157–197.

Gerisch G, Fromm H, Huesgen A, Wick U (1975) Control of cell contact sites by cAMP pulses in differentiating Dictyostelium cells. Nature 255: 547–549.

Gerlach G, Hodgins-Davis A, Avolio C, Schunter C (2008) Kin recognition in zebrafish: a 24-hour window for olfactory imprinting. Proc R Soc B Biol Sci 275: 2165–2170.

Gershenson C (2002) Complex philosophy. In: Proceedings of the 1st biennial seminar on philosophical, methodological & epistemological implications of complexity theory. La Habana, Cuba. http://arXiv.org/abs/nlin.AO/0109001.

Gershenson C (2004) Introduction to random boolean networks. In: Bedau M, Husbands P, Hutton T, Kumar S, Suzuki H, eds. Workshop and Tutorial Proceedings, Ninth International Conference on the Simulation and Synthesis of Living Systems (ALife IX). Boston, MA. pp 160–173. http://uk.arxiv.org/abs/nlin.AO/0408006.

Gershenson C (2005) A general methodology for designing self-organizing systems. CoRR abs/nlin/0505009.

Getz LL, McGuire B (1997) Communal nesting in prairie voles (Microtus ochrogaster): formation, composition, and persistence of communal groups. Can J Zool 75: 525–534.

Ghabrial AS, Krasnow MA (2006) Social interactions among epithelial cells during tracheal branching morphogenesis. Nature 441: 746–749.

Gherardi F, Atema J (2005) Memory of social partners in hermit crab dominance. Ethology 111: 271–285.

Ghoul M, Griffin AS, West SA (2014) Toward an evolutionary definition of cheating. Evolution 68: 318–331.

Giancotti FG, Tarone G (2003) Positional control of cell fate through joint integrin/receptor protein kinase signaling. Annu Rev Cell Dev Biol 19: 173–206.

Gibbs AG (2002) Lipid melting and cuticular permeability: new insights into an old problem. J Insect Physiol 48: 391–400.

Gibbs KA, Urbanowski ML, Greenberg EP (2008) Genetic determinants of self identity and social recognition in bacteria. Science 321: 256–259.

Gibo DL (1978) The selective advantage of foundress associations in Polistes fuscatus (Hymenoptera: Vespidae): a field study of the effects of predation on productivity. Can Entomol 110: 519–540.

Gilbert LI, Granger NA, Roe RM (2000) The juvenile hormones: historical facts and speculations on future research directions. Insect Biochem Mol Biol 30: 617–644.

Gilbert OM, Foster KR, Mehdiabadi NJ, Strassmann JE, Queller DC (2007) High relatedness maintains multicellular cooperation in a social amoeba by controlling cheater mutants. Proc Natl Acad Sci USA 104: 8913–8917.

Gilbert WM, Nolan PM, Stoehr AM, Hill GE (2005) Filial cannibalism at a house finch nest. Wilson Bull 117: 413–415.

Gilchrist JS (2006) Female eviction, abortion, and infanticide in banded mongooses (Mungos mungo): implications for social control of reproduction and synchronized parturition. Behav Ecol 17: 664–669.

Gilder PM, Slater PJB (1978) Interest of mice in conspecific male odours is influenced by degree of kinship. Nature 274: 364–365.

Gill DE, Chao L, Perkins SL, Wolf JB (1995) Genetic mosaicism in plants and clonal animals. Annu Rev Ecol Syst 26: 423–444.

Gill SA, Stutchbury BJ (2010) Delayed dispersal and territory acquisition in Neotropical buff-breasted wrens (Thryothorus leucotis). Auk 127: 372–378.

Gillespie DT (1977) Exact stochastic simulation of coupled chemical reactions. J Phys Chem 81: 2340–2361.

Gillespie J (1974) Polymorphism in patchy environments. Am Nat 108: 145–151.

Gillor O (2007) Bacteriocins’ role in bacterial communication. In: Riley MA, Chavan M. eds. Bacteriocins: ecology and evolution. Berlin, Germany: Springer. pp 135–146.

Gimeno CJ, Ljungdahl PO, Styles CA, Fink GR (1992) Unipolar cell divisions in the yeast S. cerevisiae lead to flamentous growth: regulation by starvation and RAS. Cell 68: 1077–1090.

Gimpl G, Fahrenholz F (2001) The oxytocin receptor system: structure, function, and regulation. Physiol Rev 81: 629–683.

Gintis H, Bowles S, Boyd R, Fehr E (2003) Explaining altruistic behavior in humans. Evol Hum Behav 24: 153–172.

Giraldeau LA (1988) The stable group and the determinants of foraging group size. In: Slobodchikoff CN, ed. The ecology of social behavior. New York, NY: Academic Press. pp 33–53.

Giraldeau LA, Caraco T (2000) Social foraging theory. Princeton, NJ: Princeton University Press.

Giraldeau LA, Dubois F (2008) Social foraging and the study of exploitative behavior. Adv Study Behav 38: 59–104.

Gjuvsland AB, Vik JO, Beard DA, Hunter PJ, Omholt SW (2013) Bridging the genotype–phenotype gap: what does it take? J Physiol 591: 2055–2066.

Glass NL, Jacobson DJ, Shiu PK (2000) The genetics of hyphal fusion and vegetative incompatibility in filamentous ascomycete fungi. Annu Rev Genet 34: 165–186.

Glazier AM, Nadeau JH, Aitman TJ (2002) Finding genes that underlie complex traits. Science 298: 2345–2349.

Glimcher PW (2005) Indeterminacy in brain and behavior. Annu Rev Psychol 56: 25–56.

Gneezy A, Fessler DMT (2011) Conflict, sticks and carrots: war increases prosocial punishments and rewards. Proc R Soc B Biol Sci 279: 219–223.

Gobbetti M, De Angelis M, Di Cagno R, Minervini F, Limitone A (2007) Cell–cell communication in food related bacteria. Int J Food Microbiol 120: 34–45.

Godfray HCJ, Grafen A (1988) Unmatedness and the evolution of eusociality. Am Nat 131: 303–305.

Godfray HCJ, Harper AB (1990) The evolution of brood reduction by siblicide in birds. J Theor Biol 145: 163–175.

Godfray HCJ, Parker GA (1992) Sibling competition, parent-offspring conflict and clutch size. Anim Behav 43: 473–490.

Godfrey-Smith P (2002) Environmental complexity and the evolution of cognition. In: Sternberg R, Kaufmann J, eds. The evolution of intelligence. London, UK: Lawrence Erlbaum Associates. pp 233–249.

Godin JGJ (1986) Antipredator function of shoaling in teleost fishes: a selective review. Nat Can 113: 241–250.

Godin JGJ (1995) Predation risk and alternative mating tactics in male Trinidadian guppies (Poecilia reticulata). Oecologia 103: 224–229.

Goewie EA (1977) Induction of caste differentiation in honey bee (Apis mellifera L.) after topical application of JH-III. Insect Soc 24: 265.

Goldbeter A (2006) Oscillations and waves of cyclic AMP in Dictyostelium: a prototype for spatio-temporal organization and pulsatile intercellular communication. Bull Math Biol 68: 1095–1109.

Goldenfeld N, Kadanoff LP (1999) Simple lessons from complexity. Science 284: 87–89.

Gómez Y, Kölliker M (2013) Maternal care, mother-offspring aggregation and age-dependent co-adaptation in the European earwig. J Evol Biol 26: 1903–1911.

Gómez-Gardeñes J, Campillo M, Moreno Y, Floría LM (2007) Dynamical organization of cooperation in complex networks. Phys Rev Lett 98: 108103.

Gómez-Gardeñes J, Poncela J, Floría LM, Moreno Y (2008) Natural selection of cooperation and degree hierarchy in heterogeneous populations. J Theor Biol 253: 296–301.

González MC, Hidalgo CA, Barabási AL (2008) Understanding individual human mobility patterns. Nature 453: 779–782.

González-Pastor JE (2011) Cannibalism: a social behavior in sporulating Bacillus subtilis. FEMS Microbiol Rev 35: 415–424.

González-Pastor JE, Hobbs EC, Losick R (2003) Cannibalism by sporulating bacteria. Science 301: 510–513.

Goo E, Majerczyk CD, An JH, Chandler JR, Seo YS, Ham H, et al. (2012) Bacterial quorum sensing, cooperativity, and anticipation of stationary-phase stress. Proc Natl Acad Sci USA 109: 19775–19780.

Goodisman MAD, Matthews RW, Crozier RH (2002) Mating and reproduction in the wasp Vespula germanica. Behav Ecol Sociobiol 51: 497–502.

Goodisman MAD, Kovacs JL, Hoffman EA (2007) Lack of conflict during queen production in the social wasp Vespula maculifrons. Mol Ecol 16: 2589–2595.

Goodman D (1974) Natural selection and a cost ceiling on reproductive effort. Am Nat 108: 247–268.

Goodnight CJ (1985) The influence of environmental variation on group and individual selection in a cress. Evolution 39: 545–558.

Goodnight CJ (2005) Multilevel selection: the evolution of cooperation in non-kin groups. Popul Ecol 47: 3–12.

Goodnight CJ, Schwartz JM, Stevens L (1992) Contextual analysis of models of group selection, soft selection, hard selection, and the evolution of altruism. Am Nat 140: 743–761.

Goodnight CJ, Stevens L (1997) Experimental studies of group selection: what do they tell us about group selection in nature? Am Nat 150(S1): S59–S79.

Goodnight KF (1992) The effect of stochastic variation on kin selection in a budding-viscous population. Am Nat 140: 1028–1040.

Goodson JL (2013) Deconstructing sociality, social evolution and relevant nonapeptide functions. Psychoneuroendocrinology 38: 465–478.

Goodson JL, Schrock SE, Klatt JD, Kabelik D, Kingsbury MA (2009) Mesotocin and nonapeptide receptors promote estrildid flocking behavior. Science 325: 862–866.

Goody E (1991) The learning of prosocial behaviour in small-scale egalitarian societies: an anthropological view. In: Hinde RAGJ, ed. Cooperation and Prosocial Behaviour. Cambridge, UK: Cambridge University Press. pp 106–128.

Gordon DM (1991) Behavioral flexibility and the foraging ecology of seed-eating ants. Am Nat 138: 379–411.

Gordon DM (2014) The ecology of collective behavior. PLoS Biol 12: e1001805.

Gordon DM, Ryder MH, Heinrich K, Murphy PJ (1996) An experimental test of the rhizopine concept in Rhizobium meliloti. Appl Environ Microbiol 62: 3991–3996.

Gordon DM, Kulig AW (1996) Founding, foraging and fighting: Colony size and the spatial distribution of harvester ant nests. Ecology 77: 2393–2409.

Gordon I, Martin C, Feldman R, Leckman J (2011) Oxytocin and social motivation. Dev Cogn Neurosci 1: 471–493.

Gore AC, Attardi B, DeFranco DB (2006) Glucocorticoid repression of the reproductive axis: effects on GnRH and gonadotropin subunit mRNA levels. Mol Cell Endocrinol 256: 40–48.

Gore J, Youk H, van Oudenaarden A (2009) Snowdrift game dynamics and facultative cheating in yeast. Nature 459: 253–256.

Gorman ML, Mills MG, Raath JP, Speakman JR (1998) High hunting costs make African wild dogs vulnerable to kleptoparasitism by hyaenas. Nature 391: 479–480.

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.

Gourley SL, Debold JF, Yin W, Cook J, Miczek KA (2005) Benzodiazepines and heightened aggressive behavior in rats: reduction by GABA(A)/alpha(1) receptor antagonists. Psychopharmacology 178: 232–240.

Gove R, Hayworth M, Chhetri M, Rueppell O (2009) Division of labour and social insect colony performance in relation to task and mating number under two alternative response threshold models. Insect Soc 56: 319–331.

Grafen A (1982) How not to measure inclusive fitness. Nature 298: 425–426.

Grafen A (1984) Natural selection, kin selection, and group selection. In: Krebs JR; Davies NB, eds. Behavioural ecology: an evolutionary approach. 2nd ed. Oxford, UK: Blackwell Scientific Publications. pp 62–84.

Grafen A (1986) Split sex ratios and the evolutionary origins of eusociality. J Theor Biol 122: 95–121.

Grafen A (1990) Do animals really recognize kin? Anim Behav 39: 42–54.

Grafen A (1991) Kin vision? a reply to Stuart. Anim Behav 41: 1095–1096.

Grafen A (1998) Evolutionary biology: Green beard as death warrant. Nature 394: 521–522.

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 and inclusive fitness. J Theor Biol 238: 541–563.

Grafen A, Archetti M (2008) Natural selection of altruism in inelastic viscous homogeneous populations. J Theor Biol 252: 694–710.

Graham JB, Dudley R, Aguilar NM, Gans C (1995) Implications of the later Palaeozoic oxygen pulse for physiology and evolution. Nature 375: 117–120.

Gray SM, Dill LM, Mckinnon JS (2007) Cuckoldry incites cannibalism: male fish turn to cannibalism when perceived certainty of paternity decreases. Am Nat 169: 258–263.

Green DJ, Cockburn A (1999) Life history and demography of an uncooperative Australian passerine, the brown thornbill. Aust J Zool 47: 633–649.

Green DJ, Cockburn A (2001) Post-fledging care, philopatry and recruitment in brown thornbills. J Anim Ecol 70: 505–514.

Green DR (2010) Cell competition: pirates on the tangled bank. Cell Stem Cell 6: 287–288.

Green L, Myerson J, Lichtman D, Rosen S, Fry A (1996) Temporal discounting in choice between delayed rewards: the role of age and income. Psychol Aging 11: 79–84.

Green L, Myerson J, Ostaszewski P (1999) Amount of reward has opposite effects on the discounting of delayed and probabilistic outcomes. J Exp Psychol Learn Mem Cogn 25: 418–427.

Green RE (1997) The influence of numbers released on the outcome of attempts to introduce exotic bird species to New Zealand. J Anim Ecol 66: 25–35.

Greene A (1991) Dolichovespula and Vespula. In: Ross KG, Matthews RW, eds. The social biology of wasps. Ithaca, NY: Cornell University Press. pp 263–308.

Greenlee JT, Callaway RM (1996) Abiotic stress and the relative importance of interference and facilitation in montane bunchgrass communities in western Montana. Am Nat 148: 386–396.

Greenlee KJ, Henry JR, Kirkton SD, Westneat MW, Fezzaa K, Lee WK, Harrison JF (2009) Synchrotron imaging of the grasshopper tracheal system: morphological components of tracheal hypermetry and the effect of age and stage on abdominal air sac volumes and convection. Am J Physiol Regul Integr Comp Physiol 297: 1343–1350.

Gregor T, Fujimoto K, Masaki N, Sawai S (2010) The onset of collective behavior in social amoebae. Science 328: 1021–1025.

Greig D, Travisano M (2004) The Prisoner’s Dilemma and polymorphism in yeast SUC genes. Proc R Soc B Biol Sci 271(Suppl. 3): S25–S26.

Grether GF (2011) The neuroecology of competitor recognition. Integr Comp Biol 51: 807–818.

Griesser M, Nystrand M, Ekman J (2006) Reduced mortality selects for family cohesion in a social species. Proc R Soc B 273: 1881–1886.

Griesser M, Barnaby J (2010) Families, a place of loving care and violent conflicts. The role of nepotism, cooperation and competition in the evolution of avian families. In: Zhang W, Liu H, eds. Behavioral and chemical ecology. New York, NY: Nova Science Publishers. pp 47–91.

Griffin AS, West SA (2002) Kin selection: fact and fiction. Trends Ecol Evol 17: 15–21.

Griffin AS, West SA (2003) Kin discrimination and the benefits of helping in cooperatively breeding vertebrates. Science 302: 634–636.

Griffin AS, West SA, Buckling A (2004) Cooperation and competition in pathogenic bacteria. Nature 430: 1024–1027.

Griffiths SW, MagurranAE (1999) Schooling decisions in guppies (Poecilia reticulata) are based on familiarity rather than kin recognition by phenotype matching. Behav Ecol Sociobiol 45: 437–443.

Griffiths SW, Armstrong JD (2001) The benefits of genetic diversity outweigh those of kin association in a territorial animal. Proc R Soc Lond Ser B Biol Sci 268: 1293–1296.

Grim P, Wardach S, Beltrani V (2006) Location, location, location: The importance of spatialization in modeling cooperation and communication. Interact Stud 7: 43–78.

Grimaldi DA, Engel M (2005) Evolution of the insects. New York, NY: Cambridge University Press.

Grime JP (1973) Competitive exclusion in herbaceous vegetation. Nature 242: 344–347.

Grishkevich V, Yanai I (2013) The genomic determinants of genotype x environment interactions in gene expression. Trends Genet 29: 479–487.

Grobstein P (1994) Variability in behavior and the nervous system. In: Ramachandran VS, ed. Encyclopedia of Human Behavior. New York: Academic Press. pp 447–458.

Groenendyk J, Michalak M (2005) Endoplasmic reticulum quality control and apoptosis. Acta Biochim Pol 52: 381–395.

Grosberg RK, Quinn JF (1986) The genetic control and consequences of kin recognition by the larvae of a colonial marine invertebrate. Nature 322: 456–459.

Grosberg RK, Quinn JF (1989) The evolution of selective aggression conditioned on allorecognition specificity. Evolution 43: 504–515.

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: 112–116.

Grosberg RK, Strathmann RR (2007) The evolution of multicellularity: a minor major transition? Annu Rev Ecol Evol Syst 38: 621–654.

Gross T, D’Lima CJD, Blasius B (2006) Epidemic dynamics on an adaptive network. Phys Rev Lett 96: 208701.

Gross T, Blasius B (2008) Adaptive coevolutionary networks: a review. J R Soc Interface 5: 259–271.

Grover JP (1997) Resource competition. London, UK: Chapman & Hall.

Grozinger CM, Sharabash NM, Whitfield CW, Robinson GE (2003) Pheromone-mediated gene expression in the honey bee brain. Proc Natl Acad Sci USA 100: 14519–14525.

Gruder-Adams S, Getz LL (1985) Comparison of the mating system and paternal behavior in Microtus ochragaster and M. pennsylvanicus. J Mamm 66: 165–167.

Grünbaum D (1998) Schooling as a strategy for taxis in a noisy environment. Evol Ecol 12: 503–522.

Guerrieri FJ, Nehring V, Jørgensen CG, Nielsen J, Galizia CG, et al. (2009) Ants recognize foes and not friends. Proc R Soc B Biol Sci 276: 2461–2468.

Guilford T (1988) The evolution of conspicuous coloration. Am Nat 131: S7–S21.

Guimarães PR Jr, Rico-Gray V, Oliveira PS, Izzo TJ, dos Reis SF, Thompson JN (2007) Interaction intimacy affects structure and coevolutionary dynamics in mutualistic networks. Curr Biol 17: 1797–1803.

Guiral S, Mitchell TJ, Martin B, Claverys JP (2005) Competence-programmed predation of noncompetent cells in the human pathogen Streptococcus pneumoniae: Genetic requirements. Proc Natl Acad Sci USA 102: 8710–8715.

Gunnthorsdottir A, Rapoport A (2006) Embedding social dilemmas in intergroup competition reduces free-riding. Organ Behav Hum Decis Process 101: 184–199.

Guo L, He X, Shi W (2014) Intercellular communications in multispecies oral microbial communities. Front Microbiol 5: 328.

Gupta J, Russell RJ, Wayman CP, Hurley D, Jackson VM (2008) Oxytocin-induced contractions within rat and rabbit ejaculatory tissues are mediated by vasopressin V1A receptors and not oxytocin receptors. Br J Pharmacol 155: 118–126.

Gurven M (2004) To give and to give not: The behavioral ecology of human food transfers. Behav Brain Sci 27: 543–583.

Gurven M, Hill K, Kaplan H, Hurtado A, Lyles R (2000) Food transfers among Hiwi foragers of Venezuela: tests of reciprocity. Hum Ecol 28: 171−218.

Gust DA (1995) Moving up the dominance hierarchy in young sooty mangabeys. Anim Behav 50: 15–21.

Gutteling EW, Riksen JAG, Bakker J, Kammenga JE (2007a) Mapping phenotypic plasticity and genotype–environment interactions affecting life-history traits in Caenorhabditis elegans. Heredity 98: 28–37.

Gutteling EW, Doroszuk A, Riksen JAG, Prokop Z, Reszka J, Kammenga JE (2007b) Environmental influence on the genetic correlations between life-history traits in Caenorhabditis elegans. Heredity 98: 206–213.

Guyon P, Petit A, Tempe J, Dessaux Y (1993) Transformed plants producing opines specifically promote growth of opine-degrading agrobacteria. Mol Plant Microbe Interact 6: 92–98.

Gwynn DT (1988) Courtship feeding in katydids benefits the mating male’s offspring. Behav Ecol Sociobiol 23: 373–377.

Haberstroh L, Firtel RA (1990) A spatial gradient of expression of a cAMP-regulated prespore cell-type-specific gene in Dictyostelium. Genes Dev 4: 596–612.

Hacker M, Kaib M, Bagine RKN, Epplen JT, Brandl R (2005) Unrelated queens coexist in colonies of the termite Macrotermes michaelseni. Mol Ecol 14: 1527–1532.

Hacker S, Gaines S (1997) Some implications of direct positive interactions for community species diversity. Ecology 78: 1990–2003.

Hackländer K, Möstl E, Arnold W (2003) Reproductive suppression in female Alpine marmots, Marmota marmota. Anim Behav 65: 1133–1140.

Hagen RH, Smith DR, Rissing SW (1988) Genetic relatedness among co-foundresses of two desert ants, Veromessor pergandei and Acromyrmex versicolor (Hymenoptera, Formicidae). Psyche 95: 191–202.

Haig D (1996) Gestational drive and the green-bearded placenta. Proc Natl Acad Sci USA 93: 6547–6551.

Hailman JP, McGowan KJ, Woolfenden GE (1994) Role of helpers in the sentinel behaviour of the Florida scrub jay (Aphelocoma c. coerulescens). Ethology 97: 119–140.

Hairston NG Jr, 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, Fox JA (2005) Rapid evolution and the convergence of ecological and evolutionary time. Ecol Lett 8: 1114–1127.

Haisley E, Mostafa R, Loewenstein G (2008) Subjective relative income and lottery ticket purchases. J Behav Decis Making 21: 283–295.

Haldane JBS (1955) Population genetics. In: Johnson ML, Abercrombie M, Fogg GE, eds. New Biology 18. Harmondsworth, UK: Penguin. pp 34–51.

Hall AR, Colegrave N (2007) How does resource supply affect evolutionary diversification? Proc R Soc B 274: 73–78.

Hall CL, Fegigan LM (1997) Spatial benefits afforded by high rank in white-faced capuchins. Anim Behav 53: 1069–1082.

Hall DW, Goodisman MAD (2012) The effects of kin selection on rates of molecular evolution in social insects. Evolution 66: 2080–2093.

Hall KRL (1964) Aggression in monkey and ape societies. In: Carthy J, Ebling F, eds. The natural history of aggression. New York, NY: Academic Press. pp 51–64.

Haller J, Barna I, Baranyi M (1995) Hormonal and metabolic responses during psychosocial stimulation in aggressive and nonaggressive rats. Psychoneuroendocrinology 20: 65–74.

Haller J, Kiem DT, Makara GB (1996) The physiology of social conflict in rats: What is particularly stressful? Behav Neurosci 110: 353–359.

Halpin ZT (1986) Individual odors among mammals: origins and functions. Adv Stud Behav 16: 39–70.

Hamblin S, Giraldeau LA (2009) Finding the evolutionarily stable learning rule for frequency-dependent foraging. Anim Behav 78: 1343–1350.

Hames R (1987) Relatedness and garden labor exchange among the Ye'kwana. Ethol Sociobiol 8: 354−392.

Hames R (2004) The purpose of exchange helps shape the mode of exchange. Behav Brain Sci 27: 564–565.

Hamilton WD (1963) The evolution of altruistic behaviour. Am Nat 97: 354–356.

Hamilton WD (1964) The genetical evolution of social behaviour, I & II. J Theor Biol 7: 1–52.

Hamilton WD (1970) Selfish and spiteful behaviour in an evolutionary model. Nature 228: 1218–1220.

Hamilton WD (1971) Geometry for the selfish herd. J Theor Biol 31: 295–311.

Hamilton WD (1972) Altruism and related phenomena, mainly in social insects. Annu Rev Ecol Syst 3: 193–232.

Hamilton WD (1975a) Innate social aptitudes in man: an approach from evolutionary genetics. In Fox R, ed. Biosocial anthropology. New York, NY: Wiley. pp 133–155.

Hamilton WD (1975b) Gamblers since life began: barnacles, aphids, elms. Q Rev Biol 50: 175–180.

Hamilton WD (1987) Discriminating nepotism: expectable, common and overlooked. In: Fletcher DJC, Michener CD, eds. Kin recognition in animals. New York, NY: Wiley. pp 417–437.

Hamilton WD (1996) Narrow roads of gene land: I. Evolution of social behaviour. Oxford, UK: W. H. Freeman.

Hamilton WD, May RM (1977) Dispersal in stable habitats. Nature 269: 578–581.

Hammer EC, Pallon J, Wallander H, Olsson PA (2011) Tit for tat? A mycorrhizal fungus accumulates phosphorus under low plant carbon availability. FEMS Microbiol Ecol 76: 236–244.

Hammerstein P (2003) Genetic and cultural evolution of cooperation. Cambridge, MA: MIT Press.

Hammock EAD, Young LJ (2004) Functional microsatellite polymorphism associated with divergent social structure in vole species. Mol Biol Evol 21: 1057–1063.

Hammock EAD, Young LJ (2005) Microsatellite instability generates diversity in brain and sociobehavioral traits. Science 308: 1630–1634.

Hammond RL, Keller L (2004) Conflict over male parentage in social insects. PLoS Biol 2: 1–11.

Han M, Sternberg PW (1990) let-60, a gene that specifies cell fates during C. elegans vulval induction, encodes a ras protein. Cell 63: 921–931.

Handa RJ, Burgess L, Kerr J, O’Keefe J (1994) Gonadal steroid hormone receptors and sex differences in the hypothalamo-pituitary-adrenal axis. Horm Behav 28: 464–476.

Hansen A (1995) The tax incidence of the Colorado state lottery instant games. Public Financ Quart 23: 385–399.

Hansen SK, Rainey PB, Haagensen JA, Molin S (2007) Evolution of species interactions in a biofilm community. Nature 445: 533–536.

Hanahan D, Weinberg RA (2000) The hallmarks of cancer. Cell 100: 57–70.

Hanahan D, Weinberg RA (2011) The hallmarks of cancer: the next generation. Cell 144: 646–674.

Hansen A, Miyazaki AD, Sprott DE (2000) The tax incidence of lotteries: Evidence from five states. J Consum Aff 34: 182–203.

Hanson BW (2013) Dynastic theory: The evolution of altruism in animal societies. http://brittwhan son.com/wp-content/uploads/2013/04/dynastic-theory-final.pdf

Hanson ES, Leibold EA (1998) Regulation of iron regulatory protein 1 during hypoxia and hypoxia/reoxygenation. J Biol Chem 273: 7588–7593.

Harcombe W (2010) Novel cooperation experimentally evolved between species. Evolution 64: 2166–2172.

Harcourt AH (1987) Dominance and fertility among female primates. J Zool 213: 471–487.

Hardin G (1968) The tragedy of the commons. Science 162: 1243–1248.

Harley JL,Smith SE (1983) Mycorrhizal symbiosis. London, UK: Academic Press.

Harms W (2001) Cooperative boundary populations: the evolution of cooperation on mortality risk gradients. J Theor Biol 213: 299–313.

Harper JL (1977) Population biology of plants. New York, NY: Academic Press.

Harper JL (1985) Modules, branches, and the capture of resources. In: Jackson JBC, Buss LW, Cook RE, eds. Population biology and evolution of clonal organisms. New Haven, CT: Yale University Press. pp 1–33.

Harpur BA, Kent CF, Molodtsova D, Lebon JM, Alqarni AS, Owayss AA, Zayed A (2014) Population genomics of the honey bee reveals strong signatures of positive selection on worker traits. Proc Natl Acad Sci USA 111: 2614–2619.

Harris AE, Ermentrout GB, Small SL (1997) A model of ocular column development by competition for trophic factor. Proc Natl Acad Sci USA 94: 9944–9949.

Harris RN, Ludwig PM (2004) Resource level and reproductive frequency in female four-toed salamanders, Hemidactylium scutatum. Ecology 85: 1585–1590.

Harrison F, Browning LE, Vos M, Buckling A (2006) Cooperation and virulence in acute Pseudomonas aeruginosa infections. BMC Biol 4: 21.

Harrison F, Buckling A (2005) Hypermutability impedes cooperation in pathogenic bacteria. Curr Biol 15: 1968–1971.

Harrison F, Buckling A (2009) Siderophore production and biofilm formation as linked social traits. ISME J 3: 632–634.

Harrison F, Buckling A (2011) Wider access to genotypic space facilitates loss of cooperation in a bacterial mutator. PLoS ONE 6: e17254.

Harrison GW, Lau MI, Williams MB (2002) Estimating individual discount rates in Denmark: a field experiment. Am Econ Rev 92: 1606–1617.

Harrison JF, Kaiser A, VandenBrooks JM (2009) Mysteries of oxygen and insect size. In: Morris S, Vosloo A, eds. 4th CPB Meeting in Africa: Mara 2008. Molecules to migration: the pressures of life. Bologna, Italy: Medimond Publishing. pp 293–302.

Harrison JF, Frazier MR, Henry JR, Kaiser A, Klok CJ, Rascón B (2006) Responses of terrestrial insects to hypoxia or hyperoxia. Respir Physiol Neurobiol 154: 4–17.

Harrison JF, Kaiser A, VandenBrooks JM (2010) Atmospheric oxygen level and the evolution of insect body size. Proc R Soc Lond B 277: 1937–1946.

Harrison JF, Haddad GG (2011) Effects of oxygen on growth and size: Synthesis of molecular, organismal, and evolutionary studies with Drosophila melanogaster. Annu Rev Physiol 73: 95–113.

Harsanyi JC (1953) Cardinal utility in welfare economics and in the theory of risk-taking. J Polit Econ 61: 434–435.

Harsanyi JC (1955) Cardinal welfare, individualistic ethics, and interpersonal comparisons of utility. J Polit Econ 63: 309–321.

Harshey RM (1994) Bees aren’t the only ones—Swarming in Gram-negative bacteria. Mol Microbiol 13: 389–394.

Hart BL, Hart LA (1992) Reciprocal allogrooming in impala, Aepyceros melampus. Anim Behav 44: 1073–1083.

Hart MM, Reader RJ (2002) Taxonomic basis for variation in the colonization strategy of arbuscular mycorrhizal fungi. New Phytol 153: 335–344.

Hartfelder K (2000) Insect juvenile hormone: from “status quo” to high society. Braz J Med Biol Res 33: 157–177.

Hartfelder K, Engels W (1998) Social insect polymorphism: hormonal regulation of plasticity in development and reproduction in the honeybee. Curr Topics Dev Biol 40: 45–77.

Hartley IR, Davies NB, Hatchwell BJ, Desrochers A, Nebel D, Burke T (1995) The polygynandrous mating system of the alpine accentor, Prunella collaris. II. Multiple paternity and parental effort. Anim Behav 49: 789–803.

Hartmann A, Wantia J, Torres JA, Heinze J (2003) Worker policing without genetic conflicts in a clonal ant. Proc Natl Acad Sci USA 100: 12836–12840.

Hartnett DC, Bazzaz FA (1983) Physiological integration among intraclonal ramets in Solidago canadensis. Ecology 64: 779–788.

Hartnett DC, Bazzaz FA (1985) The genet and ramet population dynamics of Solidago canadensis in an abandoned field. J Ecol 73: 407–413.

Hartung DK, Kirkton SD, Harrison JF (2004) Ontogeny of tracheal system structure: a light and electron-microscopy study of the metathoracic femur of the American locust, Schistocerca americana. J Morphol 262: 800–812.

Hartwell LH, Hopfield JJ, Leibler S, Murray AW (1999) From molecular to modular cell biology. Nature 402: C47–C52.

Hasiotis ST (2003) Complex ichnofossils of solitary and social soil organisms: understanding their evolution and roles in terrestrial paleoecosystems. Palaeogeogr Palaeocl 192: 259–320.

Hassett DJ, Ma JF, Elkins JG, McDermott TR, Ochsner UA, West SE, et al. (1999) Quorum sensing in Pseudomonas aeruginosa controls expression of catalase and superoxide dismutase genes and mediates biofilm susceptibility to hydrogen peroxide. Mol Microbiol 34: 1082–1093.

Hastings A (1983) Can spatial variation alone lead to selection for dispersal? Theor Popul Biol 24: 244–251.

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: 631–640.

Hatchwell BJ (2007) Avian reproduction: role of ecology in the evolution of cooperative breeding. Curr Biol 17: R845–R847.

Hatchwell BJ (2009) The evolution of cooperative breeding in birds: kinship, dispersal and life history. Phil Trans R Soc B Biol Sci 364: 3217–3227.

Hatchwell BJ, Komdeur J (2000) Ecological constraints, life history traits and the evolution of cooperative breeding. Anim Behav 59: 1079–1086.

Hatchwell BJ, Sharp SP, Simeoni M, McGowan A (2009) Factors influencing overnight loss of body mass in the communal roosts of a social bird. Funct Ecol 23: 367–372.

Hatchwell BJ, Gullett PR, Adams MJ (2014) Helping in cooperatively breeding long-tailed tits: a test of Hamilton’s rule. Phil Trans R Soc B 369: 20130565.

Hattori A, Sugime Y, Sasa C, Miyakawa H, Ishikawa Y, Miyazaki S, et al. (2013) Soldier morphogenesis in the damp-wood termite is regulated by the insulin signaling pathway. J Exp Zool B 320: 295–306.

Hauber ME, Sherman PW (2001) Self-referent phenotype matching: theoretical considerations and empirical evidence. Trends Neurosci 24: 609–616.

Hauber ME, Lacey EA (2005) Bateman’s principle in cooperatively breeding vertebrates: the effects of non-breeding alloparents on variability in female and male reproductive success. Integr Comp Biol 45: 903–914.

Hauert C, De Monte S, Hofbauer J, Sigmund K (2002) Volunteering as Red Queen mechanism for cooperation in public goods games. Science 296: 1129–1132.

Hauert C, Holmes M, Doebeli M (2006) Evolutionary games and population dynamics: maintenance of cooperation in public goods games. Proc Biol Sci 273: 3131–3132.

Hauert C, Traulsen A, Brandt H, Nowak MA, Sigmund K (2007) Via freedom to coercion: the emergence of costly punishment. Science 316: 1905–1907.

Haugen L (2001) Privation and uncertainty in the small nursery of Peruvian tadpoles: larval ecology shapes parental mating system. PhD thesis. Norman, OK: University of Oklahoma.

Hauser MD, Chen MK, Chen F, Chuang E (2003) Give unto others: genetically unrelated cotton-top tamarin monkeys preferentially give food to those who altruistically give food back. Proc R Soc Lond Ser B Biol Sci 270: 2363–2370.

Hausfater G, Blaffer Hrdy S (1984) Infanticide: comparative and evolutionary perspectives. New York, NY: Aldine Pub. Co.

Håvarstein LS, Martin B, Johnsborg O, Granadel C, Claverys JP (2006) New insights into pneumococcal fratricide: Relationship to clumping and identification of a novel immunity factor. Mol Microbiol 59: 1297–1307.

Hawkes K (1983) Kin selection and culture. Am Ethnologist 10: 345−363.

Hawn AT, Radford AN, du Plessis MA (2007) Delayed breeding affects lifetime reproductive success differently in male and female green woodhoopoes. Curr Biol 17: 844–849.

Hayano A, Yoshioka M, Tanaka M, Amano M (2004) Differentiation in the Pacific white-sided dolphin Lagenorhynchus obliquidens inferred from mitochondrial DNA and microsatellite analyses. Zool Sci 21: 989–999.

Hayashi K, Lopes SM, Tang F, Surani MA (2008) Dynamic equilibrium and heterogeneity of mouse pluripotent stem cells with distinct functional and epigenetic states. Cell Stem Cell 3: 391–401.

Hayden BY, Platt ML (2007) Temporal discounting predicts risk sensitivity in rhesus macaques. Curr Biol 17: 49–53.

Hayden-Hixon DM, Ferris CF (1991) Steroid specific regulation of agonistic responding in the anterior hypothalamus of male hamsters. Physiol Behav 50: 793–799.

He WM, Alpert P, Yu FH, Zhang LL, Dong M (2011) Reciprocal and coincident patchiness of multiple resources differentially affect benefits of clonal integration in two perennial plants. J Ecol 99: 1202–1210.

He X, Zhang J (2006) Toward a molecular understanding of pleiotropy. Genetics 173: 1885–1891.

Heams T (2012) Selection within organisms in the nineteenth century: Wilhelm Roux's complex legacy. Prog Biophys Mol Biol 110: 24–33.

Heath KD, Tiffin P (2009) Stabilizing mechanisms in a legume-rhizobium mutualism. Evolution 63: 652–662.

Heath KD, Stinchcombe JR (2014) Explaining mutualism variation: a new evolutionary paradox? Evolution 68: 309–317.

Hecht T, Appelbaum S (1988) Observations on intraspecific aggression and coeval sibling cannibalism by larval and juvenile Claias gariepinus (Clariidae: Pisces) under controlled conditions. J Zool 214: 21–44.

Hecht T, Pienaar AG (1993) A review of cannibalism and its implication in fish larviculture. J World Aquacult Soc 24: 246–261.

Heekeren HR, Marrett S, Bandettini PA, Ungerleider LG (2004) A general mechanism for perceptual decision-making in the human brain. Nature 431: 859–862.

Hefetz A, Katzav-Gozansky T (2004) Are multiple honeybee queen pheromones indicators for a queen-workers arms race? Apiacta 39: 44–52.

Heg D (2008) Reproductive suppression in female cooperatively breeding cichlids. Biol Lett 4: 606–609.

Heg D, Bachar Z, Brouwer L, Taborsky M (2004) Predation risk is an ecological constraint for helper dispersal in a cooperatively breeding cichlid. Proc Biol Sci 271: 2367–2374.

Heg D, Bachar Z, Taborsky M (2005) Cooperative breeding and group structure in the Lake Tanganyika cichlid Neolamprologus savoryi. Ethology 111: 1017–1043.

Heg D, Taborsky M (2010) Helper response to experimentally manipulated predation risk in the cooperatively breeding cichlid Neolamprologus pulcher. PLoS ONE 5: e10784.

Hein G, Singer T (2008) I feel how you feel but not always: the empathic brain and its modulation. Curr Opin Neurobiol 18: 153–158.

Heininger K (1999a) A unifying hypothesis of Alzheimer’s disease. I. Ageing sets the stage. Hum Psychopharmacol Clin Exp 14: 363–414.

Heininger K (1999b) A unifying hypothesis of Alzheimer’s disease. II. Pathophysiological processes. Hum Psychopharmacol Clin Exp 14: 525–581.

Heininger K (2000a) A unifying hypothesis of Alzheimer’s disease. III. Risk factors. Hum Psychopharmacol Clin Exp 15: 1–70.

Heininger K (2000b) A unifying hypothesis of Alzheimer’s disease. IV. Causation and sequence of events. Rev Neurosci 11: 213–328.

Heininger K (2001) The deprivation syndrome is the driving force of phylogeny, ontogeny and oncogeny. Rev Neurosci 12: 217–287.

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.

Heininger K (2013) The mutagenesis-selection-cascade theory of sexual reproduction. WebmedCentral REPRODUCTION 4: WMC004367.

Heininger K (2015) Duality of stochasticity and natural selection: a cybernetic evolution theory. WebmedCentral 6: WMC004796.

Heinrich B (1979) Bumblebee economics. Cambridge, MA: Harvard University Press.

Heinrich EC, Farzin M, Klok CJ, Harrison JF (2011) The effect of developmental stage on the sensitivity of cell and body size to hypoxia in Drosophila melanogaster. J Exp Biol 214: 1419–1427.

Heinrichs M, Baumgartner T, Kirschbaum C, Ehlert U (2003) Social support and oxytocin interact to suppress cortisol and subjective responses to psychosocial stress. Biol Psychiatry 54: 1389–1398.

Heinrichs M, von Dawans B, Domes G (2009) Oxytocin, vasopressin, and human social behavior. Front Neuroendocrinol 30: 548–557.

Heinsohn RG (1991) Slow learning of foraging skills and extended parental care in cooperatively breeding white-winged choughs. Am Nat 137: 864–881.

Heinsohn RG (1992) Cooperative enhancement of reproductive success in white-winged choughs. Evol Ecol 6: 97–114.

Heinsohn R, Cockburn A (1994) Helping is costly to young birds in cooperatively breeding white-winged choughs. Proc R Soc Lond Ser B 256: 293–298.

Heinsohn RG, Packer C (1995) Complex cooperative strategies in group-territorial African lions. Science 269: 1260–1217.

Heinsohn R, Legge S (1999) The cost of helping. Trends Ecol Evol 14: 53–57.

Heinze J (1993a) Habitat structure, dispersal strategies and queen number in two boreal Leptothorax ants. Oecologia 96: 32–39.

Heinze J (1993b) Queen-queen interactions in polygynous ants. In: Keller L, ed. Queen number and sociality in insects. Oxford, UK: Oxford University Press. pp 334–361.

Heinze J, Foitzik S Hippert A, Hölldobler B (1996) Apparent dear-enemy phenomenon and environment-based recognition cues in the ant Leptothorax nylanderi. Ethology 102: 510–522.

Heinze J, d’Ettorre P (2009) Honest and dishonest communication in social Hymenoptera. J Exp Biol 212: 1775–1779.

Heisenberg WK (1971) Physics and beyond. New York, NY: Harper & Row.

Helanterä H, Sundström L (2005) Worker reproduction in the ant Formica fusca. J Evol Biol 18: 162–171.

Helanterä H, Ratnieks FLW (2009) Sex allocation conflict in insect societies: who wins? Biol Lett doi:10.1098/rsbl.2009.0501.

Helbing D, Yu W (2009) The outbreak of cooperation among success-driven individuals under noisy conditions. Proc Natl Acad Sci USA 106: 3680–3685.

Helbing D, Szolnoki A, Perc M, Szabó G (2010) Evolutionary establishment of moral and double moral standards through spatial interactions. PLoS Comput Biol 6: e1000758.

Helling RB, Vargas CN, Adams J (1987) Evolution of Escherichia coli during growth in a constant environment. Genetics 116: 349–358.

Hellström K, Kytöviita MM, Tuomi J, Rautio P (2006). Plasticity of clonal integration in the perennial herb Linaria vulgaris after damage. Funct Ecol 20: 413–420.

Hellung-Larsen P, Lyhne I (1993) Certain fatty acids and steroids protect Tetrahymena from cell division stress caused by shaking. Microbios 75: 241–247.

Helms Cahan S, Fewell JH (2004) Division of labor and the evolution of task sharing in queen associations of the harvester ant Pogonomyrmex californicus. Behav Ecol Sociobiol 56: 9–17.

Helms Cahan S, Helms KR (2012) Relatedness does not explain geographic variation in queen cooperation in the seed-harvester ant Messor pergandei. Insectes Soc 59: 579–585.

Helms Cahan S, Gardner-Morse E (2013) The emergence of reproductive division of labor in forced queen groups of the ant Pogonomyrmex barbatus. J Zool 291: 12–22.

Hemelrijk CK (1996) Dominance interactions, spatial dynamics and emergent reciprocaty in a virtual world. In: Maes MJMP, Meyer JA, Pollack J, Wilson SW, eds. From Animals to Animats 4: Proceedings of the fourth international conference on simulation of adaptive behavior. Cambridge, MA: MIT Press/Bradford Books. pp 545–552.

Hemelrijk CK (1997) Cooperation without genes, games or cognition. In: Husbands P, Harvey I, eds. Fourth European Conference on Artificial Life. Brighton, UK: MIT Press.

Hemelrijk CK (1999) An individual-orientated model of the emergence of despotic and egalitarian societies. Proc R Soc Lond B 266: 361–369.

Hemelrijk CK (2002) Understanding social behaviour with the help of complexity science (invited article). Ethology 108: 655–671.

Hemelrijk C, ed. (2005) Self-organization and evolution of social systems. New York, NY: Cambridge University Press.

Hemelrijk CK, Hildenbrandt H (2012) Schools of fish and flocks of birds: their shape and internal structure by self-organization. Interface Focus 2: 726–737.

Hemelrijk CK, Puga-Gonzalez I (2012) An individual-oriented model on the emergence of support in fights, its reciprocation and exchange. PLoS ONE 7: e37271.

Hemker HC, Hemker PW (1969) General kinetics of enzyme cascades. Proc R Soc B 173: 411–420.

Hempel CG (1966) Philosophy of natural science. Englewood Cliffs, NJ: Prentice Hall.

Henderson G, Andersen JF, Phillips JK, Jeanne RL (1990) Internest aggression and identification of possible nestmate discrimination pheromones in polygynous ant Formica montana. J Chem Ecol 16: 2217–2228.

Henrich J (2004) Animal behaviour: inequity aversion in capuchins? Nature 428: 139.

Henrich J, McElreath R (2003) The evolution of cultural evolution. Evol Anthropol 12: 123–135.

Henry JP, Wang S (1998) Effects of early stress on adult affiliative behavior. Psychoneuroendocrinology 23: 863–875.

Henry JR, Harrison JF (2004) Plastic and evolved responses of larval tracheae and mass to varying atmospheric oxygen content in Drosophila melanogaster. J Exp Biol 207: 3559–3567.

Hepper PG, ed. (1991) Kin recognition. Cambridge, UK: Cambridge University Press.

Herbers JM (1986) Nest site limitation and facultative polygyny in the ant Leptothorax longispinosus. Behav Ecol Sociobiol 19: 115–122.

Herbers JM (1993) Ecological determinants of queen number in ants. In: Keller L, ed. Queen number and sociality in insects. Oxford, UK: Oxford University Press. pp 262–293.

Hermans EJ, Putman P, Van Honk J (2006) Testosterone administration reduces empathetic behavior: A facial mimicry study. Psychoneuroendocrinology 31: 859–866.

Herre EA (1993) Population structure and the evolution of virulence in nematode parasites of fig wasps. Science 259: 1442–1445.

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: 1501–1507.

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.

Herre EA, Wcislo WT (2011) In defence of inclusive fitness theory. Nature 471: E8–E9.

Hewitt SE, Macdonald DW, Dugdale HL (2009) Context-dependent linear dominance hierarchies in social groups of European badgers, Meles meles. Anim Behav 77: 161–169.

Hewstone M, Rubin M, Willis H (2002) Intergroup bias. Annu Rev Psychol 53: 575–604.

Hey JD (2001) Does repetition improve consistency? Exp Econ 4: 5–54.

Hey JD, Orme C (1994) Investigating generalisations of expected utility theory using experimental data. Econometrica 62: 1291–1326.

Hietakangas V, Cohen SM (2009) Regulation of tissue growth through nutrient sensing. Annu Rev Genet 43: 389–410.

Higley JD, Mehlman PT, Poland RE, Taub DM, Higley SB, Suomi SJ, et al. (1992) Cerebrospinal fluid monoamine and adrenal correlates of aggression in free-ranging rhesus monkeys. Arch Gen Psychiatry 49: 436–441.

Higley JD, Mehlman PT, Poland RE, Taub DM, Vickers J, Suomi SJ, et al. (1996) CSF testosterone and 5-HIAA correlate with different types of aggressive behaviors. Biol Psychiatry 40: 1067–1082.

Hilbe C, Sigmund K (2010) Incentives and opportunism: from the carrot to the stick. Proc R Soc B 277: 2427–2433.

Hilbe C, Traulsen A (2012) Emergence of responsible sanctions without second order free riders, antisocial punishment or spite. Sci Rep 2: 458.

Hildenbrandt H, Carere C, Hemelrijk CK (2010) Self-organized aerial displays of thousands of starlings: a model. Behav Ecol 21: 1349–1359.

Hill E, Jenkins J, Farmer L (2008) Family unpredictability, future discounting, and risk taking. J Socio Econ 37: 1381–1396.

Hill K, Hurtado AM (2009) Cooperative breeding in South American hunter-gatherers. Proc R Soc B 276: 3863–3870.

Hillegass MA, Waterman JM, Roth JD (2010) Parasite removal increases reproductive success in a social African ground squirrel. Behav Ecol 21: 696–700.

Hillesland KL, Lenski RE, Velicer GJ (2007) Ecological variables affecting predatory success in Myxococcus xanthus. Microb Ecol 53: 571–578.

Hilson JA, Kolmes SA, Nellis LF (1994) Fruiting body architecture, spore capsule contents, selfishness, and heterocytosis in the cellular slime mold, Dictyostelium discoideum. Ethol Ecol Evol 6: 529–535.

Hinchey MG, Sterritt R, Rouff C (2007) Swarms and swarm intelligence. IEEE Computer 40: 111–113.

Hinde RA (2002) Why good is good. The sources of morality. London, UK: Routledge.

Hindorff LA, Sethupathy P, Junkins HA, Ramos EM, Mehta JP, Collins FS, Manolio TA (2009) Potential etiologic and functional implications of genome-wide association loci for human diseases and traits. Proc Natl Acad Sci USA 106: 9362–9367.

Hines HM, Hunt JH, O’Connor TK, Gillespie JJ, Cameron SA (2007) Multigene phylogeny reveals eusociality evolved twice in vespid wasps. Proc Natl Acad Sci USA 104: 3295–3299.

Hirao A, Ehlers RU (2010) Influence of inoculum density on population dynamics and dauer juvenile yields in liquid culture of biocontrol nematodes Steinernema carpocapsae and S. feltiae (Nematoda: Rhabditida). Appl Microbiol Biotechnol 85: 507–515.

Hirsch BT, Stanton MA, Maldonado JE (2012) Kinship shapes affiliative social networks but not aggression in ring-tailed coatis. PLoS ONE 7: 37301–37309.

Hirschliefer J (1977) Economics from a biological perspective. J Law Econ 20: 1–53.

Hirsh AE, Fraser HB (2001) Protein dispensability and rate of evolution. Nature 411: 1046–1049.

Ho HI, Hirose S, Kuspa A, Shaulsky G (2013) Kin recognition protects cooperators against cheaters. Curr Biol 23: 1590–1595.

Hoang T (2004) The origin of hematopoietic cell type diversity. Oncogene 23: 7188–7198.

Hoare DJ, Couzin ID, Godin JGJ, Krause J (2004) Context-dependent group size choice in fish. Anim Behav 67: 155–164.

Hobbs HH, Brown MS, Goldstein JL (1992) Molecular genetics of the LDL receptor gene in familial hypercholesterolemia. Hum Mutat 1: 445–466.

Hochberg ME, van Baalen M (1998) Antagonistic coevolution over productivity gradients. Am Nat 152: 620–634.

Hochberg ME, Gomulkiewicz R, Holt RD, Thompson JN (2000) Weak sinks could cradle mutualistic symbioses – strong sources should harbour parasitic symbioses. J Evol Biol 13: 213–222.

Hochberg ME, van Baalen M (2000) A geographical perspective of virulence. In: Poulin R, Morans S, Skorping A, eds. Evolutionary biology of host–parasite relationships: theory meets reality. Amsterdam, The Netherlands: ElsevierScience Ltd. pp 81–96.

Hodge SJ, Thornton A, Flower TP, Clutton-Brock TH (2009) Food limitation increases aggression in juvenile meerkats. Behav Ecol 20: 930–935.

Hodgkin J (1998) Seven types of pleiotropy. Int J Dev Biol 42: 501–505.

Hodgson GM (2002) Darwinism in economics: from analogy to ontology. J Evol Econ 12: 259–281.

Hoeksema JD, Schwartz MW (2003) Expanding comparative–advantage biological market models: contingency of mutualism on partner's resource requirements and acquisition trade–offs. Proc R Soc Lond Ser B Biol Sci 270: 913–919.

Hoeksema JD, Chaudhary VB, Gehring CA, Johnson NCN, Karst J, Koide RT, et al. (2010) A meta-analysis of context-dependency in plant response to inoculation with mycorrhizal fungi. Ecol Lett 13: 394–407.

Hofbauer J, Sigmund K (1998) Evolutionary games and population dynamics. Cambridge, UK: Cambridge University Press.

Hofer H, East ML (2008) Siblicide in Serengeti spotted hyenas: a long-term study of maternal input and cub survival. Behav Ecol Sociobiol 62: 341–351.

Hoffman ML (1981) Is altruism part of human nature? J Pers Soc Psychol 40: 121–137.

Hoffmann KA (2011) Conflict and conflict-resolution in lower termite societies. PhD thesis. Osnabrück, Germany: University of Osnabrück.

Hoffmann K, Korb J (2011) Is there conflict over direct reproduction in lower termite colonies? Anim Behav 81: 265–274.

Hoffmann K, Foster KR, Korb J (2012) Nest value mediates reproductive decision making within termite societies. Behav Ecol 23: 1203–1208.

Hogan D, Kolter R (2002) Why are bacteria refractory to antimicrobials? Curr Opin Microbiol 5: 472–477.

Hogendoorn K, Velthuis HHW (1993) The sociality of Xylocopa pubescens: Does a helper really help? Behav Ecol Sociobiol 32: 247–257.

Hogendoorn K, Zammit J (2001) Benefits of cooperative breeding through increased colony survival in an allodapine bee. Insect Soc 48: 392–397.

Hogendoorn K, Watiniasih NL, Schwarz MP (2001) Extended alloparental care in the almost solitary bee Exoneurella eremophila (Hymenoptera: Apidae). Behav Ecol Sociobiol 50: 275–282.

Hogeweg P (1988) MIRROR beyond MIRROR, puddles of LIFE. In: Langton C, ed. Artificial life, SFI studies in the sciences of complexity. Redwood City, CA: Addison-Wesley. pp 297–316.

Hogg MA (1992) The social psychology of group cohesiveness. From attraction to social identity. New York, NY: Harvester Wheatsheaf.

Holbrook CT, Clark RM, Jeanson R, Bertram SM, Kukuk PF, Fewell JH (2009) Emergence and consequences of division of labor in associations of normally solitary sweat bees. Ethology 115: 301–310.

Holland B, Rice WR (1999) Experimental removal of sexual selection reverses intersexual antagonistic coevolution and removes a reproductive load. Proc Natl Acad Sci USA 96: 5083–5088.

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, Bronstein JL (2008) Mutualism. In: Jorgensen SE, Fath BD, eds. Encyclopedia of ecology, Vol. 3: Population dynamics. Oxford, UK: Elsevier. pp 2485–2491.

Holland JN, DeAngelis DL (2009) Consumer-resource theory predicts dynamic transitions between outcomes of interspecific interactions. Ecol Lett 12: 1357–1366.

Holland JN, DeAngelis DL (2010) A consumer-resource approach to the density-dependent population dynamics of mutualism. Ecology 91: 1286–1295.

Holland JN, Wang Y, Sun S, DeAngelis DL (2013) Consumer–resource dynamics of indirect interactions in a mutualism–parasitism food web module. Theor Ecol 6: 475–493.

Hölldobler B (1976) Recruitment behavior, home range orientation and territoriality in harvester ants, Pogonomyrmex. Behav Ecol Sociobiol 1: 3–44.

Hölldobler B, Carlin N (1989) Colony founding, queen control, and worker reproduction in the ant Aphaenogaster (=Novomessor) cockerelli. Psyche 96: 131-151.

Hölldobler B, Wilson EO (1990) The ants. Cambridge, MA: Harvard University Press.

Hölldobler B, Wilson EO (2009) The superorganism: the beauty, elegance and strangeness of insect societies. New York, NY: W. W. Norton and Company.

Hölldobler B, Wilson EO (2011) The leafcutter ants: civilization by instinct. New York, NY: WW Norton.

Hollén LI, Radford AN (2009) The development of alarm call behaviour in mammals and birds. Anim Behav 78: 791–800.

Holman L, Dreier S, d’Ettorre P (2010a) Selfish strategies and honest signalling: reproductive conflicts in ant queen associations. Proc R Soc B 277: 2007–2015.

Holman L, Jørgensen CG, Nielsen J, d’Ettorre P (2010b) Identification of an ant queen pheromone regulating worker sterility. Proc R Soc B Biol Sci 277: 3793–3800.

Holman L, van Zweden JS, Linksvayer TA, d’Ettorre P (2013) Crozier's paradox revisited: maintenance of genetic recognition systems by disassortative mating. BMC Evol Biol 13: 211.

Holman SD, Goy RW (1995) Experiential and hormonal correlates of care-giving in rhesus macaques. In: Pryce CR, Skuse RDMD, eds. Motherhood in human and nonhuman primates: biosocial determinants. Basel, Switzerland: Karger. pp 87–93.

Holme P, Newman MEJ (2006) Nonequilibrium phase transition in the coevolution of networks and opinions. Phys Rev E 74: 056108.

Holmes WG (1984) Sibling recognition in thirteen-lined ground squirrels: effects of genetic relatedness, rearing association, and olfaction. Behav Ecol Sociobiol 14: 225–233.

Holmes WG (1986) Identification of paternal half siblings by captive Belding’s ground squirrels. Anim Behav 34: 321–327.

Holmgren M, Scheffer M, Huston MA (1997) The interplay of facilitation and competition in plant communities. Ecology 78: 1966–1975.

Holmgren M, Scheffer M (2010) Strong facilitation in mild environments: the stress gradient hypothesis revisited. J Ecol 98: 1269–1275.

Holt RD (1985) Population-dynamics in 2-patch environments – some anomalous consequences of an optimal habitat distribution. Theor Popul Biol 28: 181–208.

Holt-Lunstad J, Smith TB, Layton JB (2010) Social relationships and mortality risk: a meta-analytic review. PloS Med 7: 1–20.

Holzer B, Kümmerli R, Keller L, Chapuisat M (2006) Sham nepotism as a result of intrinsic differences in brood viability in ants. Proc R Soc B 273: 2049–2052.

Honkanen T, Haukioja E (1994) Why does a branch suffer more after branch-wide than after tree-wide defoliation? Oikos 71: 441–450.

Hoogland JL (1985) Infanticide in prairie dogs: lactating females kill offspring of close kin. Science 230: 1037–1040.

Hoogland JL (1995) The black-tailed prairie dog: social life of a burrowing mammal. Chicago, IL: University of Chicago Press.

Hoover SER, Keeling CI, Winston ML, Slessor KN (2003) The effect of queen pheromones on worker honey bee ovary development. Naturwissenschaften 90: 477–480.

Hoover SER, Oldroyd BP, Wossler TC, Winston ML (2005a) Anarchistic queen honey bees have normal queen mandibular pheromones. Insect Soc 52: 6–10.

Hoover SER, Winston ML, Oldroyd BP (2005b) Retinue attraction and ovary activation: responses of wild type and anarchistic honey bees (Apis mellifera) to queen and brood pheromones. Behav Ecol Sociobiol 59: 278–284.

Hoover SER, Higo HA, Winston ML (2006) Worker honey bee ovary development: seasonal variation and the influence of larval and adult nutrition. J Comp Physiol B Biochem Syst Environ Physiol 176: 55–63.

Hori K, Matsumoto S (2010) Bacterial adhesion: from mechanism to control. Biochem Eng J 48: 424–434.

Horn HS (1968) The adaptive significance of colonial nesting in the Brewer’s blackbird (Euphagus cyanocephalus). Ecology 49: 682–694.

Horowitz A (2012) Fair is fine, but more is better: limits to inequity aversion in the domestic dog. Soc Justice Res 25: 195–212.

Hostetler CM, Ryabinin AE (2013) The CRF system and social behavior: a review. Front Neurosci 7: 92.

Hough SR, Laslett AL, Grimmond SB, Kolle G, Pera MF (2009) A continuum of cell states spans pluripotency and lineage commitment in human embryonic stem cells. PLoS ONE 4: e7708.

House JS, Landis KR, Umberson D (1988) Social relationships and health. Science 241: 540–545.

Howard RW, McDaniel CA, Blomquist GJ (1980) Chemical mimicry as an integrating mechanism: cuticular hydrocarbons of a termitophile and its host. Science 210: 431–433.

Howard RW, McDaniel CA, Blomquist GJ (1982) Chemical mimicry as an integrating mechanism for three termitophiles associated with Reticulitermes virginicus (Banks). Psyche 89: 157–168.

Howard RW, Blomquist GJ (2005) Ecological, behavioral, and biochemical aspects of insect hydrocarbons. Annu Rev Entomol 50: 371–393.

Hrdy SB (1999) Mother nature: A history of mothers, infants and natural selection. New York, NY: Pantheon.

Hrdy SB (2009) Mothers and others: The evolutionary origins of mutual understanding. Cambridge, MA: Harvard University Press.

Hsu HJ, Drummond-Barbosa D (2009) Insulin levels control female germline stem cell maintenance via the niche in Drosophila. Proc Natl Acad Sci USA 106: 1117–1121.

Huang HC, Mitchison TJ, Shi J (2010) Stochastic competition between mechanistically independent slippage and death pathways determines cell fate during mitotic arrest. PLoS One 5: e15724.

Huang W, Richards S, Carbone MA, Zhu D, Anholt RRH, Ayroles JF, et al. (2012) Epistasis dominates the genetic architecture of Drosophila quantitative traits. Proc Natl Acad Sci USA 109: 15553–15559.

Huang WS (2008) Predation risk of whole-clutch filial cannibalism in a tropical skink with maternal care. Behav Ecol 19: 1069–1074.

Huang ZY, Robinson GE (1992) Honeybee colony integration: worker–worker interactions mediate hormonally regulated plasticity in division of labor. Proc Natl Acad Sci USA 89: 11726–11729.

Huang ZY, Robinson GE, Borst DW (1994) Physiological correlates of division of labor among similarly aged honey bees. J Comp Physiol A 174: 731–739.

Huck UW, Bracken AC, Lisk RD (1983) Female-induced pregnancy block in the golden hamster. Behav Neural Biol 38: 190–193.

Huck UW, Lisk RD, Miller KS, Bethel A (1988) Progesterone levels and socially-induced implantation failure and fetal resorption in golden hamsters (Mesocricetus auratus). Physiol Behav 44: 321–326.

Hudson RE, Aukema JE, Rispe C, Roze D (2002) Altruism, cheating, and anticheater adaptations in cellular slime molds. Am Nat 160: 31–43.

Hudson R, Trillmich F (2008) Sibling competition and cooperation in mammals: challenges, developments and prospects. Behav Ecol Sociobiol 62: 299–307.

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: 257–262.

Hughes CR, Strassmann JE (1988) Age is more important than size in determining dominance among workers in the primitively eusocial wasp, Polistes instabilis. Behaviour 107: 1–14.

Hughes W, Lavery J (2004) Critical thinking. An introduction to the basic skills. Toronto, Canada: Broadview Press Ltd.

Hughes WOH, Ratnieks FLW, Oldroyd BP (2008a) Multiple paternity or multiple queens: two routes to greater intracolonial genetic diversity in the eusocial Hymenoptera. J Evol Biol 21: 1090–1095.

Hughes WOH, Oldroyd BP, Beekman M, Ratnieks FLW (2008b) Ancestral monogamy shows kin selection is key to the evolution of eusociality. Science 320: 1213–1216.

Hull LA (1982) Progress towards a unified theory of the mechanisms of carcinogenesis, III: Circumvention of proliferation controls. Med Hypotheses 8: 85–93.

Humphrey NK (1976) The social function of intellect. In: Bateson PPG, Hinde RA, eds. Growing points in ethology. Cambridge, UK: Cambridge University Press. pp 303–317.

Hunt GJ, Page RE Jr (1995) Linkage map of the honey bee, Apis mellifera, based on RAPD markers. Genetics 139: 1371–1382.

Hunt JH (2007) The evolution of social wasps. Oxford, UK: Oxford University Press.

Huntingford FA, Turner AK (1987) Animal conflict. London, UK: Chapman and Hall.

Hurst JL, Payne CE, Nevison CM, Marie AD, Humphries RE, Robertson DHL, et al. (2001) Individual recognition in mice mediated by major urinary proteins. Nature 414: 631–634.

Hurst LD, Smith NG (1999) Do essential genes evolve slowly? Curr Biol 9: 747–750.

Hüser J, Rechenmacher CE, Blatter LA (1998) Imaging the permeability pore transition in single mitochondria. Biophys J 74: 2129–2137.

Hutchings MJ (1999) Clonal plants as cooperative systems: benefits in heterogeneous environments. Plant Species Biol 14: 1–10.

Hutchings MJ, Bradbury IK (1986) Ecological perspectives on clonal perennial herbs. BioScience 36: 178–182.

Hutchings MJ, de Kroon H (1994) Foraging in plants: the role of morphological plasticity in resource acquisition. Adv Ecol Res 25: 159–238.

Hutchings MJ, Wijesinghe DK (1997) Patchy habitat, division of labour, and growth dividends in clonal plants. Trends Ecol Evol 12: 390–394.

Hwang J, Kim S, Lee D (2009) Temporal discounting and inter-temporal choice in rhesus monkeys. Front Behav Neurosci 3: 9.

Hwang YC, Flannagan SE, Clewell DB, Sedgley CM (2011) Bacteriocin-related siblicide in clinical isolates of enterococci. Probiotics Antimicro Prot 3: 57–61.

Idnurm A, Walton FJ, Floyd A, Heitman J (2008) Identification of the sex genes in an early diverged fungus. Nature 451: 193–196.

Igaki T (2009) Correcting developmental errors by apoptosis: lessons from Drosophila JNK signaling. Apoptosis 14: 1021–1028.

Iida H (2003) Small within clutch variance in spiderling body size as a mechanism for avoiding sibling cannibalism in the wolf spider Pardosa pseudoannulata (Araneae: Lycosidae). Popul Ecol 45: 1–6.

Iliopoulos D, Hintze A, Adami C (2010) Critical dynamics in the evolution of stochastic strategies for the iterated Prisoner’s Dilemma. PLoS Comput Biol 6: e1000948.

Illera JC, Diaz M (2006) Reproduction in an endemic bird of a semiarid island: a food-mediated process. J Avian Biol 37: 447–456.

Imhof LA, Fudenberg D, Nowak MA (2005) Evolutionary cycles of cooperation and defection. Proc Natl Acad Sci USA 102: 10797–10800.

Imhof LA, Nowak MA (2010) Stochastic evolutionary dynamics of direct reciprocity. Proc R Soc B Biol Sci 277: 463–468.

Inbar S, Katzav-Gozansky T, Hefetz A (2008) Kin composition effects on reproductive competition among queenless honeybee workers. Naturwissenschaften 95: 427–432.

Ingram KK, Pilko A, Heer J, Gordon DM (2013) Colony life history and lifetime reproductive success of red harvester ant colonies. J Anim Ecol 82: 540–550.

Innocent TM, West SA, Sanderson JL, Hyrkkanen N, Reece SE (2011) Lethal combat over limited resources: testing the importance of competitors and kin. Behav Ecol 22: 923–931.

Inouye Κ (1989) Control of cell type proportions by a secreted factor in Dictyostelium discoideum. Development 107: 605–609.

Insall R, Nayler O, Kay RR (1992) DIF-1 induces its own breakdown in Dictyostelium. EMBO J 11: 2849–2854.

Insel TR (2010) The challenge of translation in social neuroscience: a review of oxytocin, vasopressin, and affiliative behavior. Neuron 65: 768–779.

Insel TR, Shapiro LE (1992) Oxytocin receptor distribution reflects social organization in monogamous and polygamous voles. Proc Natl Acad Sci USA 89: 5981–5985.

Insel TR, Wang ZX, Ferris CF (1994) Patterns of brain vasopressin receptor distribution associated with social organization in microtine rodents. J Neurosci 14: 5381–5392.

Insel TR, Hulihan TJ (1995) A gender-specific mechanism for pair bonding: oxytocin and partner preference formation in monogamous voles. Behav Neurosci 109: 782–789.

Insel TR, Young LJ (2000) Neuropeptides and the evolution of social behaviour. Curr Opin Neurobiol 10: 784–789.

Insel TR, Young LJ (2001) The neurobiology of attachment. Nat Rev Neurosci 2: 129–136.

Ioannou CC, Guttal V, Couzin ID (2012) Predatory fish select for coordinated collective motion in virtual prey. Science 337: 1212–1215.

Irie Y, Parsek MR (2008) Quorum sensing and microbial biofilms. In: Romeo T, ed. Bacterial biofilms. Berlin, Germany: Springer-Verlag. pp 67–84.

Irwin RE, Brody AK (1998) Nectar robbing in Ipomopsis aggregata: effects on pollinator behavior and plant fitness. Oecologia 116: 519–527.

IsHak WW, Kahloon M, Fakhry H (2011) Oxytocin role in enhancing well-being: a literature review. J Affect Disord 130: 1–9.

Isler K, van Schaik CP (2009) The expensive brain: A framework for explaining evolutionary changes in brain size. J Hum Evolut 57: 392–400.

Ito F, Higashi S (1991) A linear dominance hierarchy regulating reproduction and polyethism of the queenless ant Pachycondyla sublaevis. Naturwissenschaften 78: 80–82.

Ito J, Kaneko K (2001) Spontaneous structure formation in a network of chaotic units with variable connection strengths. Phys Rev Lett 88: 028701.

Itô Y (1987) Role of pleometrosis in the evolution of eusociality in wasps. In: Itô Y, Brown JL, Kikkawa J, eds. Animal sociality: Theories and facts. Tokyo, Japan: Japan Sci Soc Press. pp 17–33.

Itô Υ (1989) The evolutionary biology of sterile soldier in aphids. Trends Ecol Evol 4: 69–73.

Ito Y (1993) Behaviour and social evolution of wasps: the communal aggregation hypothesis. New York, NY: Oxford University Press.

Jackson DW, Simecka JW, Romeo T (2002) Catabolite repression of Escherichia coli biofilm formation. J Bacteriol 184: 3406–3410.

Jackson JBC (1977) Competition on marine hard substrata: the adaptive significance of solitary and colonial strategies. Am Nat 111: 743–767.

Jallon JM (1984) A few chemical words exchanged by Drosophila during courtship and mating. Behav Genet 14: 441–478.

James R, Lazdunski C, Pattus F (1991) Bacteriocins, microcins and lantibiotics. NATO ASI series. New York, NY: Springer-Verlag.

Jan YN, Jan LY (1995) Maggot’s hair and bug’s eye: role of cell interactions and intrinsic factors in cell fate specification. Neuron 14: 1–5.

Jandér KC, Herre EA (2010) Host sanctions and pollinator cheating in the fig tree-fig wasp mutualism. Proc Biol Sci 277: 1481–1488.

Jandér KC, Herre EA, Simms EL (2012) Precision of host sanctions in the fig tree–fig wasp mutualism: consequences for uncooperative symbionts. Ecol Lett 15: 1362–1369.

Janson CH (1992) Evolutionary ecology of primate social structure. In: Smith EA, Winterhalder B, eds. Evolutionary ecology and human behavior. New York, NY: Walter de Gruyter. pp 95–130.

Janson CH (1998) Testing the predation hypothesis for vertebrate sociality: prospects and pitfalls. Behaviour 135: 389–410.

Janson CH, Goldsmith ML (1995) Predicting group size in primates: foraging costs and predation risks. Behav Ecol 6: 326–336.

Jarecki J, Johnson E, Krasnow MA (1999) Oxygen regulation of airway branching in Drosophila is mediated by Branchless FGF. Cell 99: 211–220.

Jarvis JUM, O’Riain MJ, Bennett NC, Sherman PW (1994) Mammalian eusociality: A family affair. Trends Ecol Evol 9: 47–51.

Jasienski M, Korzeniak U, Lomnicki A (1988) Ecology of kin and nonkin larval interactions in Tribolium beetles. Behav Ecol Sociobiol 22: 277–284.

Jasper H, Jones DL (2010) Metabolic regulation of stem cell behavior and implications for aging. Cell Metab 12: 561–565.

Jaspers P, Kangasjärvi J (2010) Reactive oxygen species in abiotic stress signaling. Physiol Plant 138: 405–413.

Jay SC (1966) Drifting of honeybees in commercial apiaries. II. Effect of various factors when hives are arranged in rows. J Apicult Res 5: 103–112.

Jay SC (1972) Ovary development of worker honeybees when separated from worker brood by various methods. Can J Zool 50: 661–664.

Jay SC, Jay DH (1993) The effect of kiwifruit (Actinidia deliciosa A Chev) and yellow flowered broom (Cytisus scoparius Link) pollen on the ovary development of worker honey bees (Apis mellifera L). Apidologie 24: 557–563.

Jeanne RL, Nordheim EV (1996) Productivity in a social wasp: Per capita output increases with swarm size. Behav Ecol 7: 43–48.

Jeanson R, Kukuk PF, Fewell JH (2005) Emergence of division of labour in halictine bees: contributions of social interactions and behavioural variance. Anim Behav 70: 1183–1193.

Jeanson R, Fewell JH (2008) Influence of the social context on division of labor in ant foundress associations. Behav Ecol 19: 567–574.

Jeanson R, Clark RM, Holbrook CT, Bertram SM, Fewell JH, Kukuk PF (2008) Division of labour and socially induced changes in response thresholds in associations of solitary halictine bees. Anim Behav 76: 593–602.

Jefferson KK (2004) What drives bacteria to produce a biofilm? FEMS Microbiol Lett 236: 163–173.

Jennions MD, Macdonald DW (1994) Cooperative breeding in mammals. Trends Ecol Evol 9: 89–93.

Jensen-Seaman MI, Furey TS, Payseur BA, Lu Y, Roskin KM, et al. (2004) Comparative recombination rates in the rat, mouse, and human genomes Genome Res 14: 528–538.

Jerne N (1955) The natural selection theory of antibody formation. Proc Natl Acad Sci USA 41: 849–857.

Jeschke JM, Tollrian R (2005) Effects of predator confusion on functional responses. Oikos111: 547–555.

Jeschke JM (2006) Density-dependent effects of prey defenses and predator offenses. J Theor Biol 242: 900–907.

Jeschke JM, Tollrian R (2007) Prey swarming: which predators become confused and why? Anim Behav 74: 387–393.

Jetz W, Rubenstein DR (2011) Environmental uncertainty and the global biogeography of cooperative breeding in birds. Curr Biol 21: 72–78.

Jia CX, Liu RR, Yang HX, Wang BH (2010) Effects of fluctuations on the evolution of cooperation in the prisoner's dilemma game. Europhys Lett 90: 30001.

Jin S, White E (2007) Role of autophagy in cancer: management of metabolic stress. Autophagy 3: 28–31.

Johansson J (2008) Evolutionary responses to environmental changes: how does competition affect adaptation? Evolution 62: 421–435.

Johnigk SA, Ehlers RU (1999) Endotokia matricida in hermaphrodites of Heterorhabditis spp. and the effect of the food supply. Nematology 1: 717–726.

Johns PM, Howard KJ, Breisch NL, Rivera A, Thorne BL (2009) Nonrelatives inherit colony resources in a primitive termite. Proc Natl Acad Sci USA 106: 17452–17456.

Johnson DDP, Kays R, Blackwell PG, Macdonald DW (2002) Does the resource dispersion hypothesis explain group living? Trends Ecol Evol 17: 563–570.

Johnson DDP, Stopka P, Macdonald DW (2004) Ideal flea constraints on group living: unwanted public goods and the emergence of cooperation. Behav Ecol 15: 181–186.

Johnson NC, Graham JH, Smith FA (1997) Functioning of mycorrhizal associations along the mutualism-parasitism continuum. New Phytol 135: 575–585.

Johnson PCD, Whitfield JA, Foster WA, Amos W (2002) Clonal mixing in the soldier-producing aphid Pemphigus spyrothecae (Hemiptera: Aphididae). Mol Ecol 11: 1525–1531.

Johnson RA (2004) Colony founding by pleometrosis in the semiclaustral seed-harvester ant Pogonomyrmex californicus (Hymenoptera: Formicidae). Anim Behav 68: 1189–1200.

Johnston C, Martin B, Fichant G, Polard P, Claverys JP (2014) Bacterial transformation: distribution, shared mechanisms and divergent control. Nat Rev Microbiol 12: 181–196.

Johnston LA (2009) Competitive interactions between cells: death, growth and geography. Science 324: 1679–1682.

Johnston RE (2003) Chemical communication in rodents: from pheromones to individual recognition. J Mammal 84: 1141–1162.

Johnstone RA (2000) Models of reproductive skew: a review and synthesis. Ethology 106: 5–26.

Johnstone RA, Bshary R (2008) Mutualism, market effects and partner control. J Evol Biol 21: 879–888.

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 JC, Myerscough MR, Graham S, Oldroyd BP (2004) Honey bee nest thermoregulation: Diversity promotes stability. Science 305: 402–404.

Jones ME, Mensch JA (1991) Behavioral correlates of male mating success in a multi-sire flock as determined by DNA fingerprinting. Poult Sci 70: 1493–1498.

Jones ME, Cockburn A, Hamede R, Hawkins C, Hesterman H, Lachish S, et al. (2008) Life-history change in disease-ravaged Tasmanian devil populations. Proc Natl Acad Sci USA 105: 10023–10027.

Jones TC, Parker PG (2000) Costs and benefits of foraging associated with delayed dispersal in the spider Anelosimus studiosus (Araneae, Theridiidae). J Arachnol 28: 61–69.

Jones TC, Parker PG (2002) Delayed juvenile dispersal benefits both mother and offspring in the cooperative spider Anelosimus studiosus (Araneae: Theridiidae). Behav Ecol 13: 142–148.

Jones TC, Riechert SE, Dalrymple SE, Parker PG (2007) Fostering model explains variation in levels of sociality in a spider system. Anim Behav 73: 195–204.

Jones TC, Riechert SE (2008) Patterns of reproductive success associated with social structure and microclimate in a spider system. Anim Behav 76: 2011–2019.

Jónsdóttir IS, Watson MA (1997) Extensive physiological integration: an adaptive trait in resource-poor environments? In: de Kroon H, van Groenendael J, eds. The ecology and evolution of clonal plants. Leiden, the Netherlands: Backhuys Publishers. pp 109–136.

Jordan IK, Rogozin IB, Wolf YI, Koonin EV (2002) Essential genes are more evolutionarily conserved than are nonessential genes in bacteria. Genome Res 12: 962–968.

Jordan KW, Craver KL, Magwire MM, Cubilla CE, Mackay TFC, et al. (2012) Genome-wide association for sensitivity to chronic oxidative stress in Drosophila melanogaster. PLoS ONE 7: e38722.

Jorde LB, Spuhler JN (1974) A statistical analysis of selected aspects of primate demography, ecology, and social behavior. J Anthropol Res 30: 199–224.

Jørgensen C, Fiksen Ø (2006) State-dependent energy allocation in cod (Gadus morhua). Can J Fish Aquat Sci 63: 186–199.

Jørgensen C, Ernande B, Fiksen Ø, Dieckmann U (2006) The logic of skipped spawning in fish. Can J Fish Aquat Sci 63: 200–211.

Jørgensen MM, Bross P, Gregersen N (2003) Protein quality control in the endoplasmic reticulum. APMIS Suppl 2003: 86–91.

Julien B, Kaiser AD, Garza A (2000) Spatial control of cell differentiation in Myxococcus xanthus. Proc Natl Acad Sci USA 97: 9098–9103.

Julian GE, Fewell JH (2004) Genetic variation and task specialization in the desert leaf-cutter ant, Acromyrmex versicolor. Anim Behav 68: 1–8.

Jutsum AR, Saunders TS, Cherrett JM (1979) Intraspecific aggression in the leaf-cutting ant Acromyrmex octospinosus. Anim Behav 27: 839–844.

Kaatz HH, Hildebrandt H, Engels W (1992) Primer effect of queen pheromone on juvenile hormone biosynthesis in adult worker honey bees. J Comp Physiol B Biochem Syst Environ Physiol 162: 588–592.

Kaatz H, Eichmüller S, Kreissl S (1994) Stimulatory effect of octopamine on juvenile hormone biosynthesis in honey bees (Apis mellifera): physiological and immunocytochemical evidence. J Insect Physiol 40: 865–872.

Kacelnik A, Bateson M (1996) Risky theories – the effects of variance on foraging decisions. Am Zool 36: 402–434.

Kaiser A, Klok CJ, Socha JJ, Lee WK, Quinlan MC, Harrison JF (2007) Increase in tracheal investment with beetle size supports hypothesis of oxygen limitation on insect gigantism. Proc Natl Acad Sci USA 104: 13198–13203.

Kaiser D (1996) Bacteria also vote. Science 272: 1598–1599.

Kaiser D, Losick R (1993) How and why bacteria talk to each other. Cell 73: 873–885.

Kalamatianos T, Kallo I, Goubillon ML, Coen CW (2004) Cellular expression of V1a vasopressin receptor mRNA in the female rat preoptic area: effects of oestrogen. J Neuroendocrinol 16: 525–533.

Kalantaridou SN, Makrigiannakis A, Zoumakis E, Chrousos GP (2004) Stress and the female reproductive system. J Reprod Immunol 62: 61–68.

Kalmar T, Lim C, Hayward P, Munoz-Descalzo S, Nichols J, et al. (2009) Regulated fluctuations in nanog expression mediate cell fate decisions in embryonic stem cells. PLoS Biol 7: e1000149.

Kamenšek S, Podlesek Z, Gillor O, Žgur-Bertok D (2010) Genes regulated by the Escherichia coli SOS repressor LexA exhibit heterogenous expression. BMC Microbiol 10: 283.

Kammenga JE, Doroszuk A, Riksen JA, Hazendonk E, Spiridon L, Petrescu AJ, et al. (2007) A Caenorhabditis elegans wild type defies the temperature-size rule owing to a single nucleotide polymorphism in tra-3. PLoS Genet 3: e34.

Kano T (1992) The last ape: pygmy chimpanzee behavior and ecology. Stanford, CA: Stanford University Press.

Kaplan H, Hill K, Hawkes K, Hurtado AM (1984) Food sharing among the Ache hunter-gatherers of eastern Paraguay. Curr Anthropol 25: 113–115.

Kaplan H, Hill K (1985) Food sharing among Ache foragers: Tests of explanatory hypotheses. Curr Anthropol 26: 223–245.

Kappeler PM (1993) Variation in social structure: the effects of sex and kinship on social interactions in three lemur species. Ethology 93: 125–145.

Kappeler PM, Fichtel C (2011) Female reproductive competition in Eulemur rufifrons: eviction and reproductive restraint in a plurally breeding Malagasy primate. Mol Ecol 21: 685–698.

Kappeler PM, Barrett L, Blumstein DT, Clutton-Brock TH (2013) Constraints and flexibility in mammalian social behaviour: introduction and synthesis. Phil Trans R Soc B 368: 20120337. http://dx.doi.org/10.1098/rstb.2012.0337

Kapur S, Remington G (1996) Serotonin–dopamine interaction and its relevance to schizophrenia. Am J Psychiat 153: 436−476.

Karakashian S, Milkman R (1967) Colony fusion compatibility types in Botryllus schlosseri. Biol Bull 133: 473.

Karban R (1986) Interspecific competition between folivorous insects on Erigeron glaucus. Ecology 67: 1063–1072.

Karczmarski L, Wursig B, Gailey G, Larson KW, Vanderli C (2005) Spinner dolphins in a remote Hawaiian atoll: social grouping and population structure. Behav Ecol 16: 675–685.

Karlin S, Matessi C (1983) Kin selection and altruism. Proc R Soc Lond B 219: 327–353.

Karp X, Greenwald I (2003) Post-transcriptional regulation of the E/Daughterless ortholog HLH-2, negative feedback, and birth order bias during the AC/VU decision in C. elegans. Genes Dev 17: 3100–3111.

Karsai I, Wenzel JW (1998) Productivity, individual-level and colony-level flexibility, and organization of work as consequences of colony size. Proc Natl Acad Sci USA 95: 8665–8669.

Karvonen A, Seehausen O (2012) The role of parasitism in adaptive radiations—when might parasites promote and when might they constrain ecological speciation? Int J Ecol 2012: 280169.

Kashi YD, King D, Soller M (1997) Simple sequence repeats as a source of quantitative genetic variation. Trends Genet 13: 74–78.

Kashi Y, King DG (2006) Simple sequence repeats as advantageous mutators in evolution. Trends Genet 22: 253–259.

Kasuya E (1981) Internidal drifting of workers in the Japanese paper wasp, Polistes chinensis antennalis (Vespidae: Hymenoptera). Insect Soc 28: 343–346.

Katsikopopulos KV, King AJ (2010) Swarm intelligence in animal groups: When can a collective out-perform an expert? PLoS ONE 5: e15505.

Katsnelson E, Motro U, Feldman MW, Lotem A (2012) Evolution of learned strategy choice in a frequency-dependent game. Proc R Soc B Biol Sci 279: 1176–1184.

Katzav-Gozansky T (2006) The evolution of honeybee multiple queen-pheromones–a consequence of a queen-worker arms race. Braz J Morphol Sci 23: 287–294.

Katzav-Gozansky T, Soroker V, Francke W, Hefetz A (2003) Honeybee egg-laying workers mimic a queen signal. Insect Soc 50: 20–23.

Katzav-Gozansky T, Boulay R, Soroker V, Hefetz A (2004) Queen–signal modulation of worker pheromonal composition in honeybees. Proc R Soc B Biol Sci 271: 2065–2069.

Katzir G (1983) Relationships between social structure and response to novelty in captive jackdaws, Corvus monedula L. II. Response to novel palatable food. Behaviour 87: 183–208.

Kauffman SA (1969) Metabolic stability and epigenesis in randomly constructed genetic nets. J Theor Biol 22: 437–467.

Kauffman SA (1993) The origins of order. Oxford, UK: Oxford University Press.

Kautz S, Lumbsch HT, Ward PS, Heil M (2009) How to prevent cheating: A digestive specialization ties mutualistic plant-ants to their ant-plant partners. Evolution 63: 839–853.

Kavaliers M, Colwell DD, Choleris E, Agmo A, Muglia LJ, Ogawa S, Pfaff DW (2003) Impaired discrimination of and aversion to parasitized male odors by female oxytocin knockout mice. Genes Brain Behav 2: 220–230.

Kavaliers M, Choleris E, Pfaff DW (2005) Recognition and avoidance of the odors of parasitized conspecifics and predators: differential genomic correlates. Neurosci Biobehav Rev 29: 1347–1359.

Kawata T, Early A, Williams J (1996) Evidence that a combined activator–repressor protein regulates Dictyostelium stalk cell differentiation. EMBO J 15: 3085–3092.

Kay RR, Jermyn KA (1983) A possible morphogen controlling differentiation in Dictyostelium. Nature 303: 242–244.

Kay RR, Flatman P, Thompson CRL (1999) DIF signalling and cell fate. Sem Cell Dev Biol 10: 577–585.

Kay RR, Thompson CRL (2001) Cross-induction of cell types in Dictyostelium: evidence that DIF-1 is made by prespore cells. Development 128: 4959–4966.

Kearns DB (2008) Division of labour during Bacillus subtilis biofilm formation. Mol Microbiol 67: 229–231.

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: 357–369.

Kearsey MJ (1965) The interaction of food supply and competition in two lines of Drosophila melanogaster. Heredity 20: 169–181.

Keeling PJ, Burger G, Durnford DG, Lang BF, Lee RW, Pearlman RE, Roger AJ, Gray MW (2005) The tree of eukaryotes. Trends Ecol Evol 20: 670–676.

Keeping MG, Crewe R (1987) The ontogeny and evolution of foundress associations in Belonogaster petiolata (Hymenoptera: Vespidae). In: Eder J, Rembold H, eds. Chemistry and biology of social insects. München, Germany: Verlag J. Pepeny. pp 383–384.

Kéfi S, van Baalen M, Rietkerk M, Loreau M (2008) Evolution of local facilitation in arid ecosystems. Am Nat 172: E1–E17.

Keller EF, Segel LA (1971) Model for chemotaxis. J Theor Biol 30: 225–234.

Keller L, ed. (1993) Queen number and sociality in insects. Oxford, UK: Oxford University Press.

Keller L (1995) Social life: the paradox of multiple-queen colonies. Trends Ecol Evol 10: 355–360.

Keller L (1997) Indiscriminate altruism: unduly nice parents and siblings. Trends Ecol Evol 12: 99–103.

Keller L, ed. (1999) Levels of selection in evolution. Princeton, NJ: Princeton University Press.

Keller L, Passera L (1989) Influence of the number of queens on nestmate recognition and attractiveness of queens to workers in the Argentine ant, Iridomyrmex humilis (Mayr). Anim Behav 37: 733–740.

Keller L, Nonacs P (1993) The role of queen pheromones in social insects: queen control or queen signal? Anim Behav 45: 787–794.

Keller L, Reeve HK (1994) Partitioning of reproduction in animal societies. Trends Ecol Evol 9: 98–102.

Keller L, Ross KG (1998) Selfish genes: a green beard in the red fire ant. Nature 394: 573–575.

Keller L, Surette MG (2006) Communication in bacteria: an ecological and evolutionary perspective. Nat Rev Microbiol 4: 249–258.

Kelley SE (1989) Experimental studies of the evolutionary significance of sexual reproduction. V. A field test of the sib-competition lottery hypothesis. Evolution 43: 1054–1065.

Kellner K, Barth B, Heinze J (2010) Colony fusion causes within-colony variation in a parthenogenetic ant. Behav Ecol Sociobiol 64: 737–746.

Kellner K, Heinze J (2011) Absence of nepotism in genetically heterogeneous colonies of a clonal ant. Ethology 117: 556–564.

Kelly JK (1992) Restricted migration and the evolution of altruism. Evolution 46: 1492–1495.

Kelly JK (1994) The effect of scale dependent processes on kin selection: mating and density regulation. Theor Popul Biol 46: 32–57.

Kelly JK (1996) Kin selection in the annual plant Impatiens capensis. Am Nat 147: 899–918.

Kemp AC, Woodcock M (1995) The hornbills: bucerotiformes. Oxford, UK: Oxford University Press.

Kempenaers B, Sheldon BC (1996) Why do male birds not discriminate between their own and extra-pair offspring? Anim Behav 51: 1165–1173.

Kendrick KM (2000) Oxytocin, motherhood and bonding. Exp Physiol 85 Spec No: 111S–124S.

Kendrick KM, Keverne EB, Baldwin BA (1987) Intracerebroventricular oxytocin stimulates maternal behaviour in the sheep. Neuroendocrinology 46: 56–61.

Kennedy D, Norman C (2005) What don't we know? Science 309: 75.

Kennedy J, Kennedy JF, Eberhart RC (2001) Swarm intelligence. San Diego, CA: Academic Press.

Kennedy JS (1966) Some outstanding questions in insect behaviour. In: Haskell PT, ed. Insect behaviour. Symposium of the Royal Entomological Society of London, No. 3. London, UK: Royal Entomological Society. pp 97–112.

Kent DS, Simpson JA (1992) Eusociality in the beetle Austroplatypus incompertus (Coleoptera: Curculionidae). Naturwissenschaften 79: 86–87.

Kenward RE (1978) Hawks and doves: Factors affecting success and selection in goshawk attacks on woodpigeons. J Anim Ecol 47: 449–460.

Kerr B, Feldman MW (2003) Carving the cognitive niche: optimal learning strategies in homogeneous and heterogeneous environments. J Theor Biol 220: 169–188.

Kerr B, Godfrey-Smith P, Feldman MW (2004) What is altruism? Trends Ecol Evol 19: 135–140.

Keverne EB, Meller RE, Eberhart A (1982) Dominance and subordination: concepts or physiological states? In: Chiarelli AB, Corruccini RS, eds. Advanced views in primate biology. Berlin, Germany: Springer-Verlag. pp 81–94.

Khamis AM, Kamel MS, Salichs MA (2006) Cooperation: concepts and general typology. In: SMC'06. IEEE International Conference on Systems, Man and Cybernetics (Vol. 2). IEEE. pp 1499–1505.

Khan AI, Dinh DM, Schneider D, Lenski RE, Cooper TF (2011) Negative epistasis between beneficial mutations in an evolving bacterial population. Science 332: 1193–1196.

Khare A, Shaulsky G (2006) First among equals: competition between genetically identical cells. Nat Rev Genet 7: 577–583.

Khare A, Santorelli LA, Strassmann JE, Queller DC, Kuspa A, et al. (2009) Cheater-resistance is not futile. Nature 461: 980–982.

Khila A, Abouheif E (2010) Evaluating the role of reproductive constraints in ant social evolution. Phil Trans R Soc B Biol Sci 365: 617–630.

Kibert CJ, Thiele L, Peterson A, Monroe M (2011) The ethics of sustainability. http://www.cce.ufl.edu/wp-content/uploads/2012/08/Ethics%20of%20Sustainability%20Textbook.pdf.

Kidner J, Moritz RAF (2013) The Red Queen process does not select for high recombination rates in haplodiploid hosts. Evol Biol 40: 377–384.

Kiers E, Rousseau R, West S, Denison R (2003) Host sanctions and the legume–rhizobium mutualism. Nature 425: 78–81.

Kiers ET, Rousseau RA, Denison RF (2006) Measured sanctions: legume hosts detect quantitative variation in rhizobium cooperation and punish accordingly. Evol Ecol Res 8: 1077–1086.

Kiers ET, Denison RF, Kawakita A, Herre EA (2011a) The biological reality of host sanctions and partner fidelity. Proc Natl Acad Sci USA 108: E7.

Kiers ET, Duhamel M, Beesetty Y, Mensah JA, Franken O, Verbruggen E, et al. (2011b) Reciprocal rewards stabilize cooperation in the mycorrhizal symbiosis. Science 333: 880–882.

Khila A, Abouheif E (2010) Evaluating the role of reproductive constraints in ant social evolution. Phil Trans R Soc B 365: 617–630.

Kikuchi Y, Hosokawa T, Fukatsu T (2011) An ancient but promiscuous host–symbiont association between Burkholderia gut symbionts and their heteropteran hosts. ISME J 5: 446–460.

Kikuta N, Tsuji K (1999) Queen and worker policing in the monogynous and monandrous ant, Diacamma sp. Behav Ecol Sociobiol 46: 180–189.

Killingback T, Doebeli M, Knowlton N (1999) Variable investment, the continuous prisoner’s dilemma, and the origin of cooperation. Proc R Soc Ser B 266: 1723–1728.

Killingback T, Doebeli M (2002) The continuous prisoner’s dilemma and the evolution of cooperation through reciprocal altruism with variable investment. Am Nat 160: 421–438.

Killingback T, Bieri J, Flatt T (2006) Evolution in group-structured populations can resolve the tragedy of the commons. Proc R Soc Lond B 273: 1477–1481.

Kils U (1986) Verhaltensphysiologische Untersuchungen an pelagischen Schwärmen, Schwarmbildung als Strategie zur Orientierung in Umweltgradienten, Bedeutung der Schwarmbildung in der Aquakultur. Ber Inst Meeresk Kiel 163: 1–168; as cited in Pitcher TJ, Parrish JK (1986) Functions of shoaling behavior in teleosts. In: Pitcher TJ, ed. Behaviour of teleost fishes. London, UK: Chapman & Hall. pp 416–418.

Kim K, Krafft B, Choe J (2005) Cooperative prey capture by young subsocial spiders I. Functional value. Behav Ecol Sociobiol 59: 92–100.

Kim PS, Lee PP, Levy D (2011) A theory of immunodominance and adaptive regulation. Bull Math Biol 73: 1645–1665.

Kimble J (1981) Alterations in cell lineage following laser ablation of cells in the somatic gonad of Caenorhabditis elegans. Dev Biol 87: 286–300.

Kimble J, Hirsh D (1979) The postembryonic cell lineages of the hermaphrodite and male gonads in Caenorhabditis elegans. Dev Biol 70: 396–417.

Kimura M (1989) The neutral theory of molecular evolution and the world view of the neutralists. Genome 31: 24–31.

King AJ, Cowlishaw G (2009) Leaders, followers, and group decision-making. Commun Integr Biol 2: 147–150.

King DG (1994) Triple repeat DNA as a highly mutable regulatory mechanism. Science 263: 595–596.

King JL (1967) Continuously distributed factors affecting fitness. Genetics 55: 483–492.

Kingma SA, Hall ML, Peters A (2011) Multiple benefits drive helping behavior in a cooperatively breeding bird: an integrated analysis. Am Nat 177: 486–495.

Kingma SA, Santema P, Taborsky M, Komdeur J (2014) Group augmentation and the evolution of cooperation. Trends Ecol Evol 29: 476–484.

Kingsland SE (1985) Modelling nature. Chicago, IL: University of Chicago Press.

Kingsolver JG, Pfennig DW (2004) Individual-level selection as a cause of Cope’s rule of phyletic size increase. Evolution 58: 1608–1612.

Kingsolver J, Huey RB (2008) Size, temperature and fitness. Three rules. Evol Ecol Res 10: 251–268.

Kinross JM, Darzi AW, Nicholson JK (2011) Gut microbiome-host interactions in health and disease. Genome Med 3: 14.

Kirsch P, Esslinger C, Chen Q, Mier D, Lis S, Siddhanti S, et al. (2005) Oxytocin modulates neural circuitry for social cognition and fear in humans. J Neurosci 25: 11489–11493.

Kirschner S, Kleineidam CJ, Zube C, Rybak J, Grünewald B, et al. (2006) Dual olfactory pathway in the honeybee, Apis mellifera. J Comp Neurol 499: 933–952.

Kisdi É, Jacobs FJA, Geritz SAH (2001) Red Queen evolution by cycles of evolutionary branching and extinction. Selection 2: 161–176.

Kisielow P, von Boehmer H (1965) Development and selection of T cells: facts and puzzles. Adv Immunol 58: 87–209.

Kitade O, Hayashi Y, Kikuchi Y, Kawarasaki S (2004) Distribution and composition of colony founding associations of a subterranean termite, Reticulitermes kanmonensis. Entomol Sci 7: 1–8.

Kitano H (2002) Computational system biology. Nature 420: 206–210.

Kjaer JB, Hjarvard BM, Jensen KH, Hansen-Møller J, Naesby-Larsen O (2004) Effects of haloperidol, a dopamine D2 receptor antagonist, on feather pecking behaviour in laying hens. Appl Anim Behav Sci 86: 77–91.

Klaenhammer TR (1988) Bacteriocins of lactic acid bacteria. Biochimie 70: 337–349.

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.

Klepzig KD, Moser JC, Lombardero ML, Hofstetter RW, Ayres MP (2001) Symbiosis and competition: Complex interactions among beetles, fungi, and mites. Symbiosis 30: 83–96.

Kleerebezem M, de VosWM, Kuipers OP (1999) The lantibiotics nisin and subtilin act as extracellular regulators of their own synthesis. In: Dunny GM, Winans SC, eds. Cell-cell signaling in bacteria. Washington, DC: American Society for Microbiology. pp 159–174.

Kleerebezem M, Quadri LEN (2001) Peptide pheromone-dependent regulation of antimicrobial peptide production in Gram-positive bacteria: a case of multicellular behaviour. Peptides 22: 1579–1596.

Kleerebezem M (2004) Quorum sensing control of lantibiotic production; nisin and subtilin autoregulate their own biosynthesis. Peptides 25: 1405–1414.

Kleineidam C, Rössler W (2009) Adaptations in the olfactory system of social hymenoptera. In: Gadau J, Fewell J, eds. Organization of insect societies: from genome to sociocomplexity. Cambridge, MA: Harvard University Press. pp 195–219.

Kleizen B, Braakman I (2004) Protein folding and quality control in the endoplasmic reticulum. Curr Opin Cell Biol 16: 343–349.

Kleszczynska A, Sokolowska E, Kulczykowska E (2012) Variation in brain arginine vasotocin (AVT) and isotocin (IT) levels with reproductive stage and social status in males of three-spined stickleback (Gasterosteus aculeatus). Gen Comp Endocrinol 175: 290–296.

Klimeš L (2008) Clonal splitters and integrators in harsh environments of the Trans-Himalaya. Evol Ecol 22: 351–367.

Klimešová J, Doležal J, Sammul M (2011) Evolutionary and organismic constraints on the relationship between spacer length and environmental conditions in clonal plants. Oikos 120: 1110–1120.

Klinkhamer PGL, Kubo T, Iwasa Y (1997) Herbivores and the evolution of the semelparous perennial life-history of plants. J Evol Biol 10: 529–550.

Klobuchar EA, Deslippe RJ (2002) A queen pheromone induces workers to kill sexual larvae in colonies of the red imported fire ant (Solenopsis invicta). Naturwissenschaften 89: 302–304.

Klok CJ, Harrison JF (2009) Atmospheric hypoxia limits selection for large body size in insects. PLoS ONE 4: e3876.

Klok CJ, Kaiser A, Lighton JR, Harrison JF (2010) Critical oxygen partial pressures and maximal tracheal conductances for Drosophila melanogaster reared for multiple generations in hypoxia or hyperoxia. J Insect Physiol 56: 461–469.

Klug H, Lindström K, St Mary CM (2006) Parents benefit from eating offspring: density-dependent egg survivorship compensates for filial cannibalism. Evolution 60: 2087–2095.

Klug H, Bonsall MB (2007) When to care for, abandon, or eat your offspring: the evolution of parental care and filial cannibalism. Am Nat 170: 886–901.

Klug H, Bonsall MB (2010) Life history and the evolution of parental care. Evolution 64: 823–835.

Klug H, Alonzo SH, Bonsall MB (2012) Theoretical foundations of parental care. In: Royle NJ, Smiseth PT, Kölliker M, eds. The evolution of parental care. Oxford, UK: Oxford University Press. pp 21–39.

Knight RL (1984) Responses of nesting ravens to people in areas of different human densities. Condor 86: 1345–1346.

Knutson B, Panksepp J (1996) Effects of fluoxetine on play dominance in juvenile rats. Aggress Behav 22: 297–307.

Kocher SD, Richard FJ, Tarpy DR, Grozinger CM (2008) Genomic analysis of post-mating changes in the honey bee queen (Apis mellifera). BMC Genomics 9: 232.

Kocher SD, Richard FJ, Tarpy DR, Grozinger CM (2009) Queen reproductive state modulates pheromone production and queen-worker interactions in honeybees. Behav Ecol 20: 1007–1014.

Kocher SD, Grozinger CM (2011) Cooperation, conflict, and the evolution of queen pheromones. J Chem Ecol 37: 1263–1275.

Kodric-Brown A (1986) Satellite and sneakers: Opportunistic male breeding tactics in pupfish (Cyprinodon pecosensis). Behav Ecol Sociobiol 19: 415–432.

Koenig WD, Pitelka FA, CarmenWJ, Mumme RL, Stanback MT (1992) The evolution of delayed dispersal in cooperative breeders. Q Rev Biol 67: 111–150.

Koenig WD, Mumme RL, Stanback MT, Pitelka FA (1995) Patterns and consequences of egg destruction among joint-nesting acorn woodpeckers. Anim Behav 50: 607–621.

Koenig WD, Dickinson JL, eds. (2004) Ecology and evolution of cooperative breeding in birds. New York, NY: Cambridge University Press.

Koeslag JH, Terblanche E (2003) Evolution of cooperation: cooperation defeats defection in the cornfield model. J Theor Biol 224: 399–410.

Köhler T, Buckling A, Van Delden C (2009) Cooperation and virulence of clinical Pseudomonas aeruginosa populations. Proc Natl Acad Sci USA 106: 6339–6344.

Koide RT (1991) Density-dependent response to mycorrhizal infection in Abutilon theophrasti Medic. Oecologia 85: 389–395.

Kokko H, Sutherland WJ (1998) Optimal floating and queuing strategies: consequences for density dependence and habitat loss. Am Nat 152: 354–366.

Kokko H, Lindström J, Alatalo RV, Rintamäki PT (1998) Queuing for territory positions in the lekking black grouse (Tetrao tetrix). Behavioral Ecology 9: 376–383.

Kokko H, Johnstone RA (1999) Social queuing in animal societies: a dynamic model of reproductive skew. Proc R Soc Lond B Biol Sci 266: 571–578.

Kokko H, Johnstone RA, Clutton-Brock TH (2001a) The evolution of cooperative breeding through group augmentation. Proc R Soc Lond Ser B Biol Sci 268: 187–196.

Kokko H, Sutherland WJ, Johnstone RA (2001b) The logic of territory choice: implications for conservation and source-sink dynamics. Am Nat 157: 459–463.

Kokko H, Ekman J (2002) Delayed dispersal as a route to breeding: territorial inheritance, safe havens, and ecological constraints. Am Nat 160: 468–484.

Kölliker M, Vancassel M (2007) Maternal attendance and the maintenance of family groups in common earwigs (Forficula auricularia): a field experiment. Ecol Entomol 32: 24–27.

Kolmer K, Heinze J (2000) Rank orders and division of labour among unrelated cofounding ant queens. Proc R Soc Lond B 267: 1729–1734.

Kolmes SA (1985) An ergonomic study of Apis mellifera (Hymenoptera:Apidae). J Kans Entomol Soc 58: 413–421.

Kolter R, Moreno F (1992) Genetics of ribosomally synthesized peptide antibiotics. Annu Rev Microbiol 46: 141–163.

Kolter R, Siegele DA, Tormo AA (1993) The stationary phase of the bacterial life cycle. Annu Rev Microbiol 47: 855–874.

Komdeur J (1992) Importance of habitat saturation and territory quality for the evolution of cooperative breeding in the Seychelles warbler. Nature 358: 493–495.

Komdeur J (1994) Experimental evidence for helping and hindering by previous offspring in the cooperative-breeding Seychelles warbler Acrocephalus sechellensis. Behav Ecol Sociobiol 34: 175–186.

Komdeur J, Huffstadt A, Prast W, Castle G, Mileto R, Wattel J (1995) Transfer experiments of Seychelles warblers to new islands: changes in dispersal and helping behaviour. Anim Behav 49: 695–708.

Komdeur J, Hatchwell BJ (1999) Kin recognition: function and mechanism in avian societies. Trends Ecol Evol 14: 237–241.

Komdeur J, Edelaar P (2001) Evidence that helping at the nest does not result in territory inheritance in the Seychelles warbler. Proc R Soc Lond B 268: 2007–2012.

Komdeur J, Richardson DS, Burke T (2004) Experimental evidence that kin discrimination in the Seychelles warbler is based on association and not on genetic relatedness. Proc R Soc Lond Ser B Biol Sci 271: 963–969.

Komdeur J, Richardson DS, Hatchwell B (2008) Kin-recognition mechanisms in cooperative breeding systems: ecological causes and behavioral consequences of variation. In: Korb J, Heinze J, eds. Ecology of social evolution. Berlin, Germany: Springer-Verlag. pp 175–193.

Komers PE, Curman GP (2000) The effect of demographic characteristics on the success of ungulate re-introduction. Biol Conserv 93: 187–194.

Kondo S (1988) Altruistic cell suicide in relation to radiation hormesis. Int J Radiat Biol 53: 95–102.

Kong A, Barnard J, Gudbjartsson DF, Thorleiffson G, Jonsdottir G, et al. (2004) Recombination rate and reproductive success in humans. Nat Genet 36: 1203–1206.

Kooijman SALM, Auger P, Poggiale JC, Kooi BW (2003) Quantitative steps in symbiogenesis and the evolution of homeostasis. Biol Rev 78: 435–463.

Koolhaas JM, Van Den Brink THC, Roozendaal B, Boorsma F (1990) Medial amygdala and aggressive behavior: interaction between testosterone and vasopressin. Aggressive Behav 16: 223–229.

Kopachik W, Oohata A, Dhokia B, Brookman JJ, Kay RR (1983) Dictyostelium mutants lacking DIF, a putative morphogen. Cell 33: 397–403.

Kopp M,Hermisson J (2006)The evolution of genetic architecture under frequency-dependent disruptive selection.Evolution 60: 1537–1550.

Korb J (2006) Limited food induces nepotism in drywood termites. Biol Lett 2: 364–366.

Korb J, Heinze J (2004) Multilevel selection and social evolution of insect societies. Naturwissenschaften 91: 291–304.

Korb J, Schneider K (2007) Does kin structure explain the occurrence of workers in a lower termite? Evol Ecol 21: 817–828.

Korb J, Heinze J, eds. (2008) Ecology of social evolution. Berlin, Germany: Springer-Verlag.

Korb J, Weil T, Hoffmann K, Foster KR, Rehli M (2009) A gene necessary for reproductive suppression in termites. Science 324: 758.

Korobkova E, Emonet T, Vilar JM, Shimizu TS, Cluzel P (2004) From molecular noise to behavioural variability in a single bacterium. Nature 428: 574–578.

Koschwanez JH, Foster KR, Murray AW (2011) Sucrose utilization in budding yeast as a model for the origin of undifferentiated multicellularity. PLoS Biol 9: e1001122.

Kosfeld M, Heinrichs M, Zak PJ, Fischbacher U, Fehr E (2005) Oxytocin increases trust in humans. Nature 435: 673–676.

Kossinets G, Watts DJ (2006) Empirical analysis of an evolving social network. Science 311: 88–90.

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: 2309–2317.

Kotrschal A, Taborsky B (2010) Environmental change enhances cognitive abilities in fish. PLoS Biol 8: e1000351.

Kou R, Chou SY, Chen SC, Huang ZY (2009) Juvenile hormone and the ontogeny of cockroach aggression. Horm Behav 56: 332–338.

Koyama NF, Caws C, Aureli F (2006) Interchange of grooming and agonistic support in chimpanzees. Int J Primatol 27: 1293–1309.

Koyama S, Takagi T, Martin SJ, Yoshida T, Takahashi J (2009) Absence of reproductive conflict during queen rearing in Apis cerana. Insect Soc 56: 171–175.

Krackow S, König B (2008) Microsatellite length polymorphisms associated with dispersal-related agonistic onset in male wild house mice (Mus musculus domesticus). Behav Ecol Sociobiol 62: 813–820.

Kraemer SA, Toups MA, Velicer GJ (2010) Natural variation in developmental life-history traits of the bacterium Myxococcus xanthus. FEMS Microbiol Ecol 73: 226–233.

Kraemer SA, Velicer, GJ (2014) Social complementation and growth advantages promote socially defective bacterial isolates. Proc R Soc B Biol Sci 281: 20140036.

Krams I, B?rzi?š A, Krama T, Wheatcroft D, Igaune K, Rantala MJ (2010) The increased risk of predation enhances cooperation. Proc R Soc B 277: 513–518.

Krasnec MO, Breed MD (2012) Eusocial evolution and the recognition systems in social insects. In: López-Larrea C, ed. Sensing in nature. Dordrecht, the Netherlands: Springer. pp 78–92.

Krause ET, Krüger O, Kohlmeier P, Caspers BA (2012) Olfactory kin recognition in a songbird. Biol Lett 8: 327–329.

Krause J, Godin JGJ (1995) Predator preferences for attacking particular prey group sizes: consequences for predator hunting success and prey predation risk. Anim Behav 50: 465–473.

Krause J, Ruxton GD (2002) Living in groups. Oxford, UK: Oxford University Press.

Krause J, Ruxton GD, Krause S (2010) Swarm intelligence in animals and humans. Trends Ecol Evol 25: 28–34.

Kravitz EA (2000) Serotonin and aggression: insights gained from a lobster model system and speculations on the role of amine neurons in a complex behavior. J Comp Physiol A 186: 221–238.

Kravitz EA, Huber R (2003) Aggression in invertebrates. Curr Opin Neurobiol 13: 736–743.

Krebs CJ, Lambin X, Wolff JO (2007) Social behavior and self-regulation in murid rodents. In: Wolff JO, Sherman P, eds. Rodent societies: an ecological & evolutionary perspective. Cambridge, UK: Cambridge University Press. Pp 173–184.

Krebs JR, Davies NB (1993) An introduction to behavioural ecology. 3rd edn. Oxford, UK: Blackwell Scientific Publications.

Kreft JU (2004a) Conflicts of interest in biofilms. Biofilms 1: 265–276.

Kreft JU (2004b) Biofilms promote altruism. Microbiology 150: 2751–2760.

Kreps DM, Milgrom P, Roberts J, Wilson R (1982) Rational cooperation in the finitely repeated prisoners' dilemma. J Econ Theor 27: 245–252.

Krieger MJ, Ross KG (2002) Identification of a major gene regulating complex social behavior. Science 295: 328–332.

Krink T (2000) Cooperation and selfishness in strategies for resource management. Spill Sci Technol B 6: 165–171.

Kroes I, Lepp PW, Relman DA (1999) Bacterial diversity within the human subgingival crevice. Proc Natl Acad Sci USA 96: 14547–14552.

Kronauer DJC, Schöning C, Pedersen JS, Boomsma JJ, Gadau J (2004) Extreme queen-mating frequency and colony fission in African army ants. Mol Ecol 13: 2381–2388.

Kronauer DJC, Berghoff SM, Powell S, Denny AJ, Edwards KJ, Franks NR, Boomsma JJ (2006) A reassessment of the mating system characteristics of the army ant Eciton burchellii. Naturwissenschaften 93: 402–406.

Kronauer DJC, Johnson RA, Boomsma JJ (2007) The evolution of multiple mating in army ants. Evolution 61: 413–422.

Kronauer DJC, Schöning C, d’Ettorre P, Boomsma JJ (2010) Colony fusion and worker reproduction after queen loss in army ants. Proc Biol Sci 277: 755–763.

Kropotkin P (1902) Mutual aid: a factor in evolution. London, UK: William Heinemann.

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: 2111–2120.

Krüger O (2005) Age at first breeding and fitness in goshwaks Accipter gentilis. J Anim Ecol 74: 266–273.

Kruk MR, Halasz J, Meelis W, Haller J (2004) Fast positive feedback between the adrenocortical stress response and a brain mechanism involved in aggressive behavior. Behav Neurosci 118: 1062–1070.

Krupp DB, DeBruine LM, Jones BC (2011) Cooperation and conflict in the light of kin recognition systems. In: Salmon CA, Shackelford TK, eds. The Oxford Handbook of Evolutionary Family Psychology. Oxford, UK: Oxford University Press. pp 345–362.

Kruuk H (1972) The spotted hyaena: a study of predation and social behavior. Chicago, IL: University of Chicago Press.

Kruuk H, Macdonald DW (1985) Group territories of carnivores: empires and enclaves. In: Sibly RM, Smith RH, eds. Behavioural ecology: ecological consequences of adaptive behaviour. Oxford, UK: Blackwell Scientific Publications. pp 521–536.

Kryger P, Moritz RFA (1997) Lack of kin recognition in swarming honeybees (Apis mellifera). Behav Ecol Sociobiol 40: 271–276.

Kubohara Y, Kikuchi H, Matsuo Y, Oshima Y, Homma Y (2013) Mitochondria are the target organelle of differentiation-inducing factor-3, an anti-tumor agent isolated from Dictyostelium discoideum. PLoS ONE 8: e72118.

Kucharski R, Maleszka J, Foret S, Maleszka R (2008) Nutritional control of reproductive status in honeybees via DNA methylation. Science 319: 1827–1830.

Kuhn TS (1970) The structure of scientific revolutions, 2nd edn. Chicago, IL: University of Chicago Press.

Kuipers OP, Beerthuyzen MM, de Ruyter PGGA, Luesink EJ, de Vos WM (1995) Autoregulation of nisin biosynthesis in Lactococcus lactis by signal transduction. J Biol Chem 270: 27299–27304.

Kukuk PF (1992) Cannibalism in social bees. In: Elgar MA, Crespi BJ, eds. Cannibalism: ecology and evolution among diverse taxa. Oxford, UK: Oxford University Press. pp 214–237.

Kukuk PF (1994) Replacing the terms “primitive” and “advanced”. New modifiers for the term “eusocial”. Anim Behav 47: 1475–1478.

Kukuk PF, Sage GK (1994) Reproductivity and relatedness in a communal halictine bee Lasioglossum (Chilalictus) hemichalceum. Insect Soc 41: 443–455.

Kukuk PF, Ward SA, Jozwiak A (1998) Mutualistic benefits generate an unequal distribution of risky activities among unrelated group members. Naturwissenschaften 85: 445–449.

Kummel M, Salant SW (2006) The economics of mutualisms: Optimal utilization of mycorrhizal mutualistic partners by plants. Ecology 87: 892–902.

Kummer H (1971) Spacing mechanisms in social behavior. In: Eisenberg JF, Dillon WS, eds. Man and beast: Comparative social behavior. Washington, DC: Smithsonian Institution Press. pp 219–234.

Kümmerli R, Griffin AS, West SA, Buckling A, Harrison F (2009a) Viscous medium promotes cooperation in the pathogenic bacterium Pseudomonas aeruginosa. Proc Biol Sci 276: 3531–3538.

Kümmerli R, Gardner A, West SA, Griffin AS (2009b) Limited dispersal, budding dispersal, and cooperation: an experimental study. Evolution 63: 939–949.

Kümmerli R, Jiricny N, Clarke LS, West SA, Griffin AS (2009c) Phenotypic plasticity of a cooperative behaviour in bacteria. J Evol Biol 22: 589–598.

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.

Kümmerli R, Brown SP (2010) Molecular and regulatory properties of a public good shape the evolution of cooperation. Proc Natl Acad Sci USA 107: 18921–18926.

Kun Á, Oborny B (2003) Survival and competition of clonal plant populations in spatially and temporally heterogeneous habitats. Community Ecol 4: 1–20.

Kun Á, Dieckmann U (2013) Resource heterogeneity can facilitate cooperation. Nat Commun 4: 2453.

Kuperman MN, Risau-Gusman S (2008) The effect of topology on the spatial ultimatum game. Eur Phys J B 62: 233–238.

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: 385–390.

Kupiec JJ (1997) A Darwinian theory for the origin of cellular differentiation. Mol Gen Genet 255: 201–208.

Kupke T, Götz F (1996) Post-translational modifications of lantibiotics. Antonie Van Leeuwenhoek 69: 139–150.

Kurakin A (2007) Self-organization versus Watchmaker: ambiguity of molecular recognition and design charts of cellular circuitry. J Mol Recognit 20: 205–214.

Kuussaari M, Saccheri I, Camara M, Hanski I (1998) Allee effect and population dynamics in the Glanville fritillary butterfly. Oikos 82: 384–392.

Kuzdzal-Fick JJ, Foster KR, Queller DC, Strassmann JE (2007) Exploiting new terrain: an advantage to sociality in the slime mold Dictyostelium discoideum. Behav Ecol 18: 433–437.

Kuzdzal-Fick JJ, Fox SA, Strassmann JE, Queller DC (2011) High relatedness is necessary and sufficient to maintain multicellularity in Dictyostelium. Science 334: 1548–1551.

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.

Lacey EA, Sherman PW (1991) Social organization of naked mole-rat colonies: evidence for divisions of labor. In: Sherman PW, Jarvis JUM, Alexander RD, eds. The biology of the naked mole-rat. Princeton, NJ: Princeton University Press. pp 195–208.

Lacey EA, Sherman PW (1997) Cooperative breeding in naked mole rats: implications for vertebrate and invertebrate sociality. In: Solomon NG, French JA, eds. Cooperative breeding in mammals. Cambridge, UK: Cambridge University Press. pp 267–301.

Lacey EA, Sherman PW (2007) The ecology of sociality in rodents. In: Wolff JO, Sherman PW, eds. Rodent societies: an ecological and evolutionary perspective. Chicago, IL: University of Chicago Press. pp 243–254.

Lack DL (1968) Ecological adaptations for breeding in birds. London, UK: Methuen.

Lackey BR, Gray SLL, Hendricks DM (2000) Physiological basis for use of insulin-like growth factors in reproductive applications: a review. Theriogenology 53: 1147–1156.

LaFever L, Drummond-Barbosa D (2005) Direct control of germline stem cell division and cyst growth by neural insulin in Drosophila. Science 309: 1071–1073.

Lafon-Cazal M, Baehr J (1988) Octopaminergic control of corpora allata activity in an insect. Experientia 44: 895–896.

Lahav S, Soroker V, Vander Meer RK, Hefetz A (1998) Nestmate recognition in the ant Cataglyphis niger: do queens matter? Behav Ecol Sociobiol 43: 203–212.

Lahav S, Soroker V, Hefetz A, Vander Meer RK (1999) Direct behavioral evidence for hydrocarbons as ant recognition discriminators. Naturwissenschaften 86: 246–249.

Lahti K, Koivula K, Rytkönen S, Mustonen T, Welling P, Pravosudov VV, Orell M (1998) Social influences on food caching in willow tits: a field experiment. Behav Ecol 9: 122–129.

Lam KW, Jeffreys AJ (2007) Processes of de novo duplication of human alpha-globin genes. Proc Natl Acad Sci USA 104: 10950–10955.

Lambin X, Krebs CJ (1993) Influence of female relatedness on the demography of Townsends vole populations in spring. J Anim Ecol 62: 536–550.

Lambin X, Yoccoz NG (1998) The impact of population kin-structure on nestling survival in Townsend’s voles, Microtus townsendii. J Anim Ecol 67: 1–16.

Lambin X, Aars J, Piertney SB (2001) Dispersal, intraspecific competition, kin competition and kin facilitation: a review of the empirical evidence. In: Clobert J, Danchin E, Dhondt AA, Nichols JD, eds. Dispersal. Oxford, UK: Oxford University Press. pp 110–122.

Lamm C, Batson CD, Decety J (2007) The neural substrate of human empathy: effects of perspective-taking and cognitive appraisal. J Cogn Neurosci 19: 42–58.

Landa K, Benner B, Watson MA, Gartner J (1992) Physiological integration for carbon in mayapple (Podophyllum peltatum), a clonal perennial herb. Oikos 63: 348–356.

Langer P, Hogendoorn K, Keller L (2004) Tug-of-war over reproduction in a social bee. Nature 428: 844–847.

Langer P, Nowak MA, Hauert C (2008) Spatial invasion of cooperation. J Theor Biol 250: 634–641.

Langergraber KE, Mitani JC, Vigilant L (2007) The limited impact of kinship on cooperation in wild chimpanzees. Proc Natl Acad Sci USA 104: 7786–7790.

Langergraber KE, Mitani JC, Vigilant L (2009) Kinship and social bonds in female chimpanzees (Pan troglodytes). Am J Primatol 71: 840–851.

Langmore NE, Stevens M, Maurer G, Heinsohn R, Hall ML, Peters A, Kilner RM (2011) Visual mimicry of host nestlings by cuckoos. Proc R Soc B 278: 2455–2463.

Langton CG (1997) Artificial life: An overview. Cambridge, MA: MIT Press.

Laranzo-Perea C, Castro CSS, Harrison R, Araujo A, Arruda MF, Snowdon CT (2000) Behavioural and demographic changes following the loss of the breeding female in cooperatively breeding marmosets. Behav Ecol Sociobiol 48: 137–146.

Larracuente AM, Sackton TB, Greenberg AJ, Wong A, Singh ND, et al. (2008) Evolution of protein-coding genes in Drosophila. Trends Genet 24: 114–123.

Larrere M, Couillaud F (1993) Role of juvenile hormone biosynthesis in dominance status and reproduction of the bumblebee, Bombus terrestris. Behav Ecol Sociobiol 33: 335–338.

Lattorff HMG, Moritz RFA (2008) Recombination rate and AT-content show opposite correlations in mammalian and other animal genomes. Evol Biol 35: 146–149.

Latty T, Beekman M (2011) Irrational decision-making in an amoeboid organism: transitivity and context-dependent preferences. Proc R Soc B Biol Sci 278: 307–312.

Lavalli KL, Herrnkind WF (2009) Collective defense by spiny lobster (Panulirus argus) against triggerfish (Balistes capriscus): effects of number of attackers and defenders. N Z J Mar Freshwater Res 43: 15–28.

Lavial F, Acloque H, Bertocchini F, Macleod DJ, Boast S, Bachelard E, et al. (2007) The Oct4 homologue PouV and Nanog regulate pluripotency in chicken embryonic stem cells. Development 134: 3549–3563.

Law R, Marrow P, Dieckmann U (1997) On evolution under asymmetric competition. Evol Ecol 11: 485–501.

Lawlor LR, Maynard Smith J (1976) The coevolution and stability of competing species. Am Nat 110: 79–99.

Lawrance EC (1991) Poverty and the rate of time preference: evidence from panel data. J Political Econ 99: 54–77.

Lazarus J (1979) Early warning function of flocking in birds: experimental study with captive quelea. Anim Behav 27: 855–865.

Leach CK, Ashworth JM, Garrod DR (1973) Cell sorting out during the differentiation of mixtures of metabolically distinct populations of Dictyostelium discoideum. J Embryol Exp Morphol 29: 647–661.

Leadbeater E, Chittka L (2007a) Social learning in insects—from miniature brains to consensus building. Curr Biol 17: R703–R713.

Leadbeater E, Chittka L (2007b) The dynamics of social learning in an insect model, the bumblebee (Bombus terrestris). Behav Ecol Sociobiol 61: 1789–1796.

Leadbeater E, Carruthers JM, Green JP, van Heusden J, Field J (2010) Unrelated helpers in a primitively eusocial wasp: Is helping tailored towards direct fitness? PLoS ONE 5: e11997.

Leadbeater E, Carruthers JM, Green JP, Rosser NS, Field J (2011) Nest inheritance is the missing source of direct fitness in a primitively eusocial insect. Science 333: 874–876.

Lease HM, Wolf BO, Harrison JF (2006) Intraspecific variation in tracheal volume in the American locust, Schistocerca americana, measured by a new inert gas method. J Exp Biol 209: 3476–3483.

LeBoeuf AC, Benton R, Keller L (2013) The molecular basis of social behavior: models, methods and advances. Curr Opin Neurobiol 23: 3–10.

Ledoux MN, Winston ML, Higo H, Keeling CI, Slessor KN, LeConte Y (2001) Queen and pheromonal factors influencing comb construction by simulated honey bee (Apis mellifera L.) swarms. Insect Soc 48: 14–20.

Lee B (2009) The role of negative regulators in coordination of the Myxococcus xanthus developmental program. Doctoral dissertation. Marburg, Germany: Philipps-Universität Marburg.

Lee HH, Molla MN, Cantor CR, Collins JJ (2010) Bacterial charity work leads to population-wide resistance. Nature 467: 82–86.

Lee JY, Yao TP (2010) Quality control autophagy: A joint effort of ubiquitin, protein deacetylase and actin cytoskeleton. Autophagy 6: 555–557.

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: 969–980.

Lee RB (1979) The !Kung San: Men, women, and work in a foraging society. Cambridge, UK: Cambridge University Press.

Lefebvre D, Ménard N, Pierre JS (2003) Modelling the influence of demographic parameters on group structure in social species with dispersal asymmetry and group fission. Behav Ecol Sociobiol 53: 402–410.

Lefebvre D, Pierre J (2007) Demographic consequences of drift in contiguous hives of Bombus terrestris. J Econ Entomol 100: 1756–1763.

Lefebvre L (2000) Feeding innovations and their cultural transmission in bird populations. In: Heyes CM, Huber L, eds. The evolution of cognition. Cambridge, MA: MIT Press. pp 311–328.

Le Gac M, Plucain J, Hindre T, Lenski RE, Schneider D (2012) Ecological and evolutionary dynamics of coexisting lineages during a long-term experiment with Escherichia coli. Proc Natl Acad Sci USA 109: 9487–9492.

Le Galliard JF, Ferrière R, Dieckmann U (2003) The adaptive dynamics of altruism in spatially heterogeneous populations. Evolution 57: 1–17.

Leggett HC, El Mouden C, Wild G, West S (2012) Promiscuity and the evolution of cooperative breeding. Proc R Soc B Biol Sci 279: 1405–1411.

Lehmann J, Boesch B (2009) Sociality of the dispersing sex: the nature of social bonds in West African female chimpanzees, Pan troglodytes. Anim Behav 77: 377–387.

Lehmann L, Keller L (2006) The evolution of cooperation and altruism–a general framework and a classification of models. J Evol Biol 19: 1365–1376.

Lehmann L, Keller L, Sumpter DJT (2007a) The evolution of helping and harming on graphs: the return of the inclusive fitness effect. J Evol Biol 20: 2284–2295.

Lehmann L, Keller L, West S, Roze D (2007b) Group selection and kin selection: two concepts but one process. Proc Natl Acad Sci USA 104: 6736–6739.

Lehmann L, Ravigné V, Keller L (2008) Population viscosity can promote the evolution of altruistic sterile helpers and eusociality. Proc R Soc B Biol Sci 275: 1887–1895.

Lehmann L, Rousset F (2010) How life history and demography promote or inhibit the evolution of helping behaviours. Phil Trans R Soc B 365: 2599–2617.

Leigh EG (1971) Adaptation and diversity. San Francisco, CA: Freeman, Cooper and Company.

Leigh EG Jr (1977) How does selection reconcile individual advantage with the good of the group? Proc Natl Acad Sci USA 74: 4542–4546.

Leigh EG Jr (1983) When does the good of the group override the advantage of the individual? Proc Natl Acad Sci USA 80: 2985–2989.

Leigh EG Jr (1991) Genes, bees and ecosystems: the evolution of a common interest among individuals. Trends Ecol Evol 6: 257–262.

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: 2507–2528.

Leigh EG Jr, Rowell TE (1995) The evolution of mutualism and other forms of harmony at various levels of biological organization. Ecologie 26: 131–158.

Leighton M, Leighton DR (1982) The relationship of size of feeding aggregate to size of food patch: howler monkeys (Alouatta palliata) feeding in Trichilia cipo fruit trees on Barro Colorado Island. Biotropica 14: 81–90.

Leimar O, Hammerstein P (2001) Evolution of cooperation through indirect reciprocity. Proc R Soc Lond B 268: 745–753.

Leimar O, Hammerstein P (2010) Cooperation for direct fitness benefits. Phil Trans R Soc B Biol Sci 365: 2619–2626.

Leinders-Zufall T, Brennan P, Widmayer P, Maul-Pavicic A, Jäger M, Li XH, et al. (2004) MHC Class I peptides as chemosensory signals in the vomeronasal organ. Science 306: 1033–1037.

Leisler B, Winkler H, Wink M (2002) Evolution of breeding systems in acrocephaline warblers. Auk 119: 379–390.

Leisner M, Stingl K, Frey E, Maier B (2008) Stochastic switching to competence. Curr Opin Microbiol 11: 553–559.

Leitão M, Guedes Á, Yamamoto ME, de Araújo Lopes F (2013) Do people adjust career choices according to socioeconomic conditions? An evolutionary analysis of future discounting. Psychol Neurosci 6: 383–390.

Lekberg Y, Hammer EC, Olsson PA (2010) Plants as resource islands and storage units - adopting the mycocentric view of arbuscular mycorrhizal networks. FEMS Microbiol Ecol 74: 336–345.

Lekberg Y, Koide RT (2014) Integrating physiological, community, and evolutionary perspectives on the arbuscular mycorrhizal symbiosis. Botany 92: 241–251.

Lema SC, Nevitt GA (2004) Exogenous vasotocin alters aggression during agonistic exchanges in male Armagosa pupfish (Cyprinodon nevadensis amargosae). Horm Behav 46: 628–637.

Le Maho Y (1977) The emperor penguin: a strategy to live and breed in the cold. Am Sci 65: 680–693.

Lemche N (1979) Andurarum and misarum: Comments on the problem of social edicts and their application in the ancient Near East. J Near Eastern Stud 38: 11–22.

Le Moli F, Mori A, Grasso DA (1992) Nestmate and conspecific non-nestmate recognition in Formica cunnicularia Latr.: the effect of diet difference. In: Billen J, ed. Biology and evolution of social insects. Leuven, Belgium: Leuven University Press. pp 161–165.

Lendenmann M, Thonar C, Barnard RL, Salmon Y, Werner RA, Frossard E, Jansa J (2011) Symbiont identity matters: carbon and phosphorus fluxes between Medicago truncatula and different arbuscular mycorrhizal fungi. Mycorrhiza 21: 689–702.

Lenoir A, D’Ettorre P, Errard C, Hefetz A (2001) Chemical ecology and social parasitism in ants. Annu Rev Entomol 46: 573–599.

Lenz M (1985) Is inter- and intraspecific variability of lower termite neotenic numbers due to adaptive thresholds for neotenic elimination? – Considerations from studies on Porotermes adamsoni (Froggatt) (Isoptera: Termopsidae). In: Watson JAL, Okot-Kotber BM, Noirot C, eds. Caste differentiation in social insects. London, UK: Pergamon Press. pp 136–145.

Lesica P, Young TP (2005) A demographic model explains life-history variation in Arabis fecunda. Funct Ecol 19: 471–477.

Lett C, Auger P, Gaillard JM (2004) Continuous cycling of grouped vs. solitary strategy frequencies in a predator–prey model. Theor Popul Biol 65: 263–270.

Lett C, Semeria M, Thiebault A, Tremblay Y (2014) Effects of successive predator attacks on prey aggregations. Theor Ecol 7: 239–252.

Leung TLF, Poulin R (2008) Parasitism, commensalism, and mutualism: exploring the many shades of symbioses. Vie Milieu 58: 107–115.

Levin SA, Cohen D, Hastings A (1984) Dispersal strategies in patchy environments. Theor Popul Biol 26: 165–191.

Levine DK (1998) Modeling altruism and spitefulness in experiments. Rev Econ Dynam 1: 593–622.

Levine RA, Campbell DT (1972) Ethnocentrism: Theories of conflict, ethnic attitudes and group behavior. New York, NY: John Wiley.

Levinton JS (2001) Genetics, paleontology and macroevolution. Cambridge, MA: Cambridge University Press.

Levy F, Melo AI, Galef G, Madden M, Fleming AS (2003) Complete maternal deprivation affects social but not spatial learning in adult rats. Dev Psychobiol 43: 177–191.

Lewis DH (1985) Symbiosis and mutualism: crisp concepts and soggy semantics. In: Boucher DH, ed. The biology of mutualism. London, UK: Croom Helm. pp 29–39.

Lewis K (2010) Persister cells. Annu Rev Microbiol 64: 357–372.

Lewontin RC (1974) The genetic basis of evolutionary change. Cambridge, MA: Harvard University Press.

Ley RE, Turnbaugh P, Klein S, Gordon J (2006a) Microbial ecology: Human gut microbes associated with obesity. Nature 444: 1022–1023.

Ley RE, Harris JK, Wilcox J, Spear JR, Miller SR, Bebout BM, et al. (2006b) Unexpected diversity and complexity of the Guerrero Negro hypersaline microbial mat. Appl Environ Microbiol 72: 3685–3695.

Li J, Wang L, Wu X, Fang O, Wang L, Lu C, et al. (2013) Polygenic molecular architecture underlying non-sexual cell aggregation in budding yeast. DNA Res 20: 55–66.

Li P, Duan H (2014) Robustness of cooperation on scale-free networks in the evolutionary prisoner’s dilemma game. Europhys Lett 105: 48003.

Li R, Montpetit A, Rousseau M, Wu SYM, Greenwood CM, Spector TD, et al. (2013) Somatic point mutations occurring early in development: a monozygotic twin study. J Med Genet doi:10.1136/jmedgenet-2013-101712.

Li SI, Purugganan MD (2011) The cooperative amoeba: Dictyostelium as a model for social evolution. Trends Genet 27: 48–54.

Li SI, Buttery NJ, Thompson CR, Purugganan MD (2014) Sociogenomics of self vs. non-self cooperation during development of Dictyostelium discoideum. BMC Genomics 15: 616.

Li W, Baker NE (2007) Engulfment is required for cell competition. Cell 129: 1215–1225.

Li YC, Korol AB, Fahima T, Nevo E (2004) Microsatellites within genes: structure, function, and evolution. Mol Biol Evol 21: 991–1007.

Li YH, Tian X (2012) Quorum sensing and bacterial social interactions in biofilms. Sensors 12: 2519–2538.

Li ZF, Li X, Liu H, Liu X, Han K, Wu ZH, et al. (2011) Genome sequence of the halotolerant marine bacterium Myxococcus fulvus HW-1. J Bacteriol 193: 5015–5016.

Liang D, Silverman J (2000) “You are what you eat”: diet modifies cuticular hydrocarbons and nestmate recognition in the Argentine ant, Linepithema humile. Naturwissenschaften 87: 412–416.

Liang D, Blomquist GJ, Silverman J (2001) Hydrocarbon-released nestmate aggression in the Argentine ant, Linepithema humile, following encounters with insect prey. Comp Biochem Physiol B Biochem Mol Biol 129: 871–882.

Liang Q, Conte N, Skarnes WC, Bradley A (2008) Extensive genomic copy number variation in embryonic stem cells. Proc Natl Acad Sci USA 105: 17453–17456.

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: 1119–1128.

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: 2072–2080.

Liao CM (2013) Risk taking begets risk taking: evidence from casino openings and investor portfolios. Unpublished working paper. University of Toronto.

Liao JC, Beal DN, Lauder GV, Triantafyllou MS (2003) Fish exploiting vortices decrease muscle activity. Science 302: 1566–1569.

Libbrecht R, Schwander T, Keller L (2011) Genetic components to caste allocation in a multiple-queen ant species. Evolution 65: 2907–2922.

Lichocki P, Tarapore D, Keller L, Floreano D (2012) Neural networks as mechanisms to regulate division of labor. Am Nat 179: 391–400.

Liebert AE, Starks PT (2006) Taming of the skew: transactional models fail to predict reproductive partitioning in the paper wasp Polistes dominulus. Anim Behav 71: 913–923.

Liebig J (2010) Hydrocarbon profiles indicate fertility and dominance status in ant, bee, and wasp colonies. In: Blomquist GJ, Bagnères AG,eds. Insect hydrocarbons: biology, biochemistry, and chemical ecology. Cambridge, UK: Cambridge University Press. pp 254–281.

Liebig J, Peeters C, Hölldobler B (1999) Worker policing limits the number of reproductives in a ponerine ant. Proc R Soc Lond B Biol Sci 266: 1865–1870.

Liebig J, Peeters C, Oldham NJ, Markstädter C, Hölldobler B (2000) Are variations in cuticular hydrocarbons of queens and workers a reliable signal of fertility in the ant Harpegnathos saltator? Proc Natl Acad Sci USA 97: 4124–4131.

Liebig J, Eliyahu D, Brent CS (2009) Cuticular hydrocarbon profiles indicate reproductive status in the termite Zootermopsis nevadensis. Behav Ecol Sociobiol 63: 1799–1807.

Liermann K, Hilborn R (1997) Depensation in fish stocks: A hierarchic Bayesian meta-analysis. Can J Fish Aquat Sci 54: 1976–1984.

Liersch S, Schmid-Hempel P (1998) Genetic variation within social insect colonies reduces parasite load. Proc R Soc Lond Ser B Biol Sci 265: 221–225.

Ligon JD, Ligon SH (1990) Green woodhoopoe: life history traits and sociality. In: Stacey PB, Koenig WD, eds. Cooperative breeding in birds. Cambridge, UK: Cambridge University Press. pp 33–65.

Lim MM, Wang Z, Olazábal DE, Ren X, Terwilliger EF, Young LJ (2004) Enhanced partner preference in a promiscuous species by manipulating the expression of a single gene. Nature 429: 754–757.

Lim MM, Young LJ (2006) Neuropeptidergic regulation of affiliative behavior and social bonding in animals. Horm Behav 50: 506–517.

Lima AAM, Lyerly DM, Wilkins TD, Innes DJ, Guerrant RL (1988) Effects of Clostridium difficile toxins A and B in rabbit small and large intestine in vivo and on cultured cells in vitro. Infect Immun 56: 582–588.

Lima SL (1994) Collective detection of predatory attack by birds in the absence of alarm signals. J Avian Biol 25: 319–326.

Lima SL (1995) Collective detection of predatory attack by social foragers: fraught with ambiguity? Anim Behav 50: 1097–1108.

Lima SL, Dill LM (1990) Behavioral decisions made under the risk of predation: a review and prospectus. Can J Zool 68: 619–640.

Lin HR, Winston ML (1998) The role of nutrition and temperature in the ovarian development of the worker honey bee (Apis mellifera). Can Entomol 130: 883–891.

Lin Ν, Michener CD (1972) Evolution of sociality in insects. Q Rev Biol 47: 131–159.

Linardi S, McConnell MA (2011) No excuses for good behavior: Volunteering and the social environment. J Public Econ 95: 445–454.

Lindenfors P, Froberg L, Nunn CL (2004) Females drive primate social evolution. Proc R Soc Lond Ser B Suppl 271: S101–S103.

Lindquist S (1986) The heat-shock response. Annu Rev Biochem 55: 1151–1191.

Linksvayer TA, Wade MJ (2005) The evolutionary origin and elaboration of sociality in the aculeate hymenoptera: maternal effects, sib-social effects, and heterochrony. Q Rev Biol 80: 317–336.

Linksvayer TA, Wade MJ (2009) Genes with social effects are expected to harbor more sequence variation within and between species. Evolution 63: 1685–1696.

Linnoila VM, Virkkunen M (1992) Aggression, suicidality, and serotonin. J Clin Psychiat 53: 46−51.

Lion S, Gandon S (2009) Habitat saturation and the spatial evolutionary ecology of altruism. J Evol Biol 22: 1487–1502.

Lion S, Gandon S (2010) Life history, habitat saturation and the evolution of fecundity and survival altruism. Evolution 64: 1594–1606.

Liow LH, Van Valen L, Stenseth NC (2011) Red Queen: from populations to taxa and communities. Trends Ecol Evol 26: 349–358.

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: 2960–2966.

Liu H, Li Y, Guo G (2014) Gene by social-environment interaction for youth delinquency and violence: thirty-nine aggression-related genes. Soc Forces doi: 10.1093/sf/sou086.

Livernois J (1987) The redistributional effect of lotteries: Evidence from Greyhound. Public Financ Quart 15: 339–351.

Lo N, Hayashi Y, Kitade O (2009) Should environmental caste determination be assumed for termites? Am Nat 173: 848–853.

Locke M (1958) The co-ordination of growth in the tracheal system of insects. Q J Microsc Sci 99: 373–391.

Lockey KH (1988) Lipids of the insect cuticle: origin, composition and function. Comp Biochem Phys B 89: 595–645.

Loewenstein G, Elster J, eds. (1992) Choice over time. New York, NY: Russell Sage.

Loker ES (2012) Macroevolutionary immunology: a role for immunity in the diversification of animal life. Front Immunol 3: 25.

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: 526–539.

Lombardo MP (2008) Access to mutualistic endosymbiotic microbes: an underappreciated benefit of group living. Behav Ecol Sociobiol 62: 479–497.

Lonstein JS, Morrell JI (2007) Neuroendocrinology and neurochemistry of maternal motivation and behavior. In: Lajtha A, Blaustein JD, eds. Handbook of neurochemistry and molecular neurobiology, 3rd edn. Berlin, Germany: Springer-Verlag. pp 195–245.

Lopez D, Vlamakis H, Kolter R (2009) Generation of multiple cell types in Bacillus subtilis. FEMS Microbiol Rev 33: 152–163.

Lopez-Pascua LDC, Buckling A (2008) Increasing productivity accelerates host-parasite coevolution. J Evol Biol 21: 853–860.

López-Suárez C, Toro MA, Garcia C (1993) Genetic heterogeneity increases viability in competing groups of Drosophila hydei. Evolution 47: 977–981.

Lopez-Vaamonde C, Koning JW, Brown RM, Jordan WC, Bourke AFG (2004) Social parasitism by male-producing reproductive workers in a eusocial insect. Nature 430: 557–560.

Lopez-Vaamonde C, Brown RM, Lucas ER, Pereboom JJ, Jordan WC, Bourke AF (2007) Effect of the queen on worker reproduction and new queen production in the bumble bee Bombus terrestris. Apidologie 38: 171–180.

López-Vargas K (2014) Risk attitudes and fairness: theory and experiment. http://econweb.umd.edu/~davis/ev entpapers/LopezAttitudes.pdf

Lorberbaum JP, Bohning DE, Shastri A, Sine LE (2002) Are there really no evolutionarily stable strategies in the iterated prisoner’s dilemma? J Theor Biol 214: 155–169.

Loreau M (1998) Ecosystem development explained by competition within and between material cycles. Proc R Soc Lond B Biol Sci 265: 33–38.

Losick R, Desplan C (2008) Stochasticity and cell fate. Science 320: 65–68.

Lott DF (1984) Intraspecific variation in the social systems of wild vertebrates. Behaviour 88: 266–325.

Lott DF (1991) Intraspecific variation in the social systems of wild vertebrates. New York, NY: Cambridge University Press.

Loudon C (1988) Development of Tenebrio molitor in low oxygen levels. J Insect Physiol 34: 97–103.

Loudon C (1989) Tracheal hypertrophy in mealworms: design and plasticity in oxygen supply systems. J Exp Biol 147: 217–235.

Lourdais O, Brischoux F, Shine R, Bonnet X (2005) Adaptive maternal cannibalism in snakes (Epicrates cenchria maurus Boidae). Biol J Linn Soc 84: 767–774.

Love OP, Salvante KG, Dale J, Williams TD (2008) Specific variability in the immune system across life-history stages. Am Nat 172: E99–112.

Lowen SB, Ozaki T, Kaplan E, Saleh BEA, Teich MC (2001) Fractal features of dark, maintained, and driven neural discharges in the cat visual system. Methods 24: 377–394.

Lozada M, D’Adamo P, Fuentes MA (2011) Beneficial effects of human altruism. J Theor Biol 289: 12–16.

Lozano S, Arenas A, Sánchez A (2008) Mezoscopic structure conditions the emergence of cooperation on social networks. PLoS ONE 3: e1892.

Lubin Y, Bilde T (2007) The evolution of sociality in spiders. Adv Stud Behav 37: 83–145.

Lucas ER, Field J (2011) Assured fitness returns in a social wasp with no worker caste. Proc R Soc B Biol Sci 278: 2991–2995.

Lucas JR, Creel SR, Waser PM (1996) How to measure inclusive fitness, revisited. Anim Behav 51: 225–228.

Lucas JR, Creel SR, Waser PM (1997) Dynamic optimization and cooperative breeding: an evaluation of future fitness effects. In: Solomon NG, French JA, eds. Cooperative breeding in mammals. Cambridge, UK: Cambridge University Press. pp 171–198.

Luce K, Weil AC, Osiewacz HD (2010) Mitochondrial protein quality control systems in aging and disease. Adv Exp Med Biol 694: 108–125.

Lucia KE, Keane B, Hayes LD, Lin YK, Schaefer RL, Solomon NG (2008) Philopatry in prairie voles: an evaluation of the habitat saturation hypothesis. Behav Ecol 19: 774–783.

Luciani MF, Kubohara Y, Kikuchi H, Oshima Y, Golstein P (2009) Autophagic or necrotic cell death triggered by distinct motifs of the differentiation factor DIF-1. Cell Death Differ 16: 564–570.

Lukas D, Reynolds V, Boesch C, Vigilant L (2005) To what extent does living in a group mean living with kin? Mol Ecol 14: 2181–2196.

Lundvall D, Svanbäck R, Persson L, Byström P (1999) Size-dependent predation in piscivores: interactions between predator foraging and prey avoidance abilities. Can J Fish Aquat Sci 56: 1285–1292.

Lutz A, Brefczynski-Lewis J, Johnstone T, Davidson R (2008) Regulation of the theme neural circuitry of emotion by compassion meditation: effects of the meditative expertise. PLoS ONE 3: e1897.

Lynch M, Lande R (1993) Evolution and extinction in response to environmental change. In: Kareiva PM, Kingsolver JG, Huey RB, eds. Biotic interactions and global change. Sunderland, MA: Sinauer Associates. pp 234–250.

Lynch M, Spitze K (1994) Evolutionary genetics of Daphnia. In: Real LA, ed. Ecological Genetics. Princeton, NJ: Princeton University Press. pp 109–128.

Lynd LR, Weimer PJ, Van Zyl WH, Pretorius IS (2002) Microbial cellulose utilization: fundamentals and biotechnology. Microbiol Mol Biol Rev 66: 506–577.

Lynn A, Ashley T, Hassold T (2004) Variation in human meiotic recombination. Annu Rev Genomics Hum Genet 5: 317–349.

Lyon BE (1993) Conspecific brood parasitism as a flexible female reproductive tactic in American coots. Anim Behav 46: 911–928.

Lyon P (2007) From quorum to cooperation: lessons from bacterial sociality for evolutionary theory. Stud Hist Philos Biol Biomed Sci 38: 820–833.

Lyon P (2011) To be or not to be: Where is self-preservation in evolutionary theory? In: Calcott B, Sterelny K, eds. The major transitions in evolution revisited. Cambridge, MA: The MIT Press. pp 105–125.

Lysenko ES, Ratner AJ, Nelson AL, Weiser JN (2005) The role of innate immune responses in the outcome of interspecies competition for colonization of mucosal surfaces. PLoS Pathog 1: e1.

Maamar H, Raj A, Dubnau D (2007) Noise in gene expression determines cell fate in Bacillus subtilis. Science 317: 526–529.

MacArthur BD, Ma'ayan A, Lemischka IR (2009) Systems biology of stem cell fate and cellular reprogramming. Nat Rev Mol Cell Biol 10: 672–681.

MacArthur R(1970) Species packing and competitive equilibrium for many species. Theor Popul Biol 1: 1–11.

MacArthur R, Levins R (1964) Competition, habitat selection, and character displacement in a patchy environment. Proc Natl Acad Sci USA 51: 1207–1210.

MacArthur R, Wilson EO (1967) The theory of island biogeography. Princeton, NJ: Princeton University Press.

MacColl ADC (2009)Parasites may contribute to ‘magic trait’ evolution in the adaptive radiation of three-spined sticklebacks, Gasterosteus aculeatus (Gasterosteiformes: Gasterosteidae). Biol J Linn Soc 96: 425–433.

MacColl ADC, Chapman SM (2010) Parasites can cause selection against migrants following dispersal between environments. Funct Ecol 24: 847–856.

Macdonald DW (1979) The flexible social system of the golden jackal (Canis aureus). Behav Ecol Sociobiol 5: 17–38.

Macdonald DW (1983) The ecology of carnivore social behaviour. Nature 301: 379–384.

Mace R (2000) Evolution ecology of human life history. Anim Behav 59: 1–10.

Macedo RHF, Bianchi CA (1997) When birds go bad: circumstantial evidence for infanticide in the communal South-American Guira Cuckoo. Ethol Ecol Evol 9: 45–54.

Macedo RHF, Cariello M, Muniz L (2001) Context and frequency of infanticide in communally breeding Guira Cuckoos. Condor 103: 170–175.

MacGregor B (2014) Why inequality undermines societies - an evolutionary perspective. Contributoria https://www.contributoria.com/issue/2014-04/52f24e1ddf80ad6e4d000002.

Machado CA, Herre EA, McCafferty S, Bermingham E (1996) Molecular phylogenies of fig pollinating and non-pollinating wasps and the implications for the origin and evolution of the fig-fig wasp mutualism. J Biogeogr 23: 521–530.

Machmer MM, Ydenberg RC (1998) The relative roles of hunger and size asymmetry in sibling aggression between nestling ospreys, Pandion haliaetus. Can J Zool 76: 181–186.

Mackay TFC (2001) Quantitative trait loci in Drosophila. Nat Rev Genet 2: 11–20.

Mackay TFC, Anholt RRH (2006) Of flies and man: Drosophila as a model for human complex traits. Annu Rev Genomics Hum Genet 7: 339–367.

MacLean RC, Gudelj I (2006) Resource competition and social conflict in experimental populations of yeast. Nature 441: 498–501.

MacLean RC, Fuentes-Hernandez A, Greig D, Hurst LD, Gudelj I (2010a) A mixture of “cheats” and “co-operators” can enable maximal group benefit. PLoS Biol 8: e1000486.

MacLean RC, Perron GG, Gardner A (2010b) Diminishing returns from beneficial mutations and pervasive epistasis shape the fitness landscape for rifampicin resistance in Pseudomonas aeruginosa. Genetics 186: 1345–1354.

MacNeill A (2009) http://evolutionlist.blogspot.com/2009/03/are-mechanisms-that-produce-phenotypic.html

Madden JR, Clutton-Brock TH (2011) Experimental peripheral administration of oxytocin elevates a suite of cooperative behaviours in a wild social mammal. Proc R Soc B 278: 1189–1194.

Madden JR, Nielsen JF, Clutton-Brock TH (2012) Do networks of social interactions reflect patterns of kinship? Curr Zool 58: 319–328.

Maderspacher F (2011) Asexuality: the insects that stick with it. Curr Biol 21: R495–497.

Madsen T, Shine R (1998) Quantity or quality? Determinants of maternal reproductive success in tropical pythons (Liasis fuscus). Proc R Soc B Biol Sci 265: 1521–1525.

Maestripieri D, Hoffman CL, Anderson GM, Carter CS, Higley JD (2009) Mother–infant interactions in free-ranging rhesus macaques: relationships between physiological and behavioral variables. Physiol Behav 96: 613–619.

Mágori K, Oborny B, Dieckmann U, Meszéna G (2003) Cooperation and competition in heterogeneous environments: the evolution of resource sharing in clonal plants. Evol Ecol Res 5: 787–817.

Magrath RD, Whittingham LA (1997) Subordinate males are more likely to help if unrelated to the breeding female in cooperatively breeding white-browed scrubwrens. Behav Ecol Sociobiol 41: 185–192.

Magurran AE (1990) The adaptive significance of schooling as an anti-predator defence in fish. Ann Zool Fenn 27: 51–66.

Magurran AE, Oulton W, Pitcher TJ (1985) Vigilant behaviour and shoal size in minnows. Z Tierpsychol 67: 167–178.

Magurran AE, Pitcher TJ (1987) Provenance, shoal size and the sociobiology of predator-evasion behaviour in minnow shoals. Proc R Soc Lond Ser B 229: 439–465.

Magyar G, Kun Á, Oborny B, Stuefer JF (2007) Importance of plasticity and decision-making strategies for plant resource acquisition in spatio-temporally variable environments. New Phytol 174: 182–193.

Maherali H, Klironomos JN (2007) Influence of phylogeny on fungal community assembly and ecosystem functioning. Science 316: 1746–1748.

Majeed H, Ghazaryan L, Herzberg M, Gillor O (2014) Bacteriocin expression in sessile and planktonic populations of Escherichia coli. J Antibiot doi: 10.1038/ja.2014.84.

Major PF (1978) Predator-prey interactions in two schooling fishes, Caranx ignobilis and Stolephorus purpureus. Anim Behav 26: 760–777.

Maklakov AA (2002) Snake-directed mobbing in a cooperative breeder: anti-predator behavior or self-advertisement for the formation of dispersal coalitions? Behav Ecol Sociobiol 52: 372–378.

Malcolm JR, Marten K (1982) Natural selection and the communal rearing of pups in African wild dogs (Lycaon pictus). Behav Ecol Sociobiol 10: 1–13.

Malka O, Shnieor S, Hefetz A, Katzav-Gozansky T (2007) Reversible royalty in worker honeybees (Apis mellifera) under the queen influence. Behav Ecol Sociobiol 61: 465–473.

Maloof JE, Inouye DW (2000) Are nectar robbers cheaters or mutualists. Ecology 81: 2651–2661.

Manceau M, Domingues VS, Mallarino R, Hoekstra HE (2011) The developmental role of Agouti in color pattern evolution. Science 331: 1062–1065.

Mangel M (1990) Dynamic information in uncertain and changing worlds. J Theor Biol 146: 317–332.

Manhes P, Velicer GJ (2011) Experimental evolution of selfish policing in social bacteria. Proc Natl Acad Sci USA 108: 8357–8362.

Manica A (2002) Filial cannibalism in teleost fish. Biol Rev 77: 261–277.

Manica A (2004) Parental fish change their cannibalistic behaviour in response to the cost-to-benefit ratio of parental care. Anim Behav 67: 1015–1021.

Manser MB (1999) Response of foraging group members to sentinel calls in suricates, Suricata suricatta. Proc R Soc B 266: 1013–1019.

Marco DE, Carbajal JP, Cannas S, Pérez-Arnedo R, Hidalgo-Perea A, Olivares J, Ruiz-Sainz JE, Sanjuán J (2009) An experimental and modelling exploration of the host-sanction hypothesis in legume-rhizobia mutualism. J Theor Biol 259: 423–33.

Marconato A, Bisazza A, Fabris M (1993) The cost of parental care and egg cannibalism in the river bullhead, Cottus gobio L (Pisces, Cottidae). Behav Ecol Sociobiol 32: 229–237.

Marée AF, Hogeweg P (2001) How amoeboids self-organize into a fruiting body: Multicellular coordination in D. discoideum. Proc Natl Acad Sci USA 98: 3879–3883.

Margulis L (1970) Origin of eukaryotic cells. New Haven, CT: Yale University Press.

Margulis L (1981) Symbiosis in cell evolution. San Francisco, CA: Freeman.

Margulis L (1998) Symbiotic planet: a new look at evolution. Amherst, MA: Basic Books.

Margulis SW, Nabong M, Alaks G, Walsh A, Lacy RC (2005) Effects of early experience on subsequent parental behaviour and reproductive success in oldfield mice, Peromyscus polionotus. Anim Behav 69: 627–634.

Marino J, Sillero-Zubiri C, Johnson PJ, Macdonald DW (2012) Ecological bases of philopatry and cooperation in Ethiopian wolves. Behav Ecol Sociobiol 66: 1005–1015.

Maris M, te Boekhorst R (1996) Exploiting physical constraints: heap formation through behavioral error in a group of robots. IEEE/RSJ International Conference on intelligent robots and systems IROS 1996, Senri Life Center, Osaka, Japan, 3: 1655–1661.

Marques JT, Carthew RW (2007) A call to arms: coevolution of animal viruses and host innate immune responses. Trends Genet 23: 359–364.

Marras S, Killen SS, Lindström J, McKenzie DJ, Steffensen JF, Domenici P (2014) Fish swimming in schools save energy regardless of their spatial position. Behav Ecol Sociobiol DOI 10.1007/s00265-014-1834-4.

Marshall JAR (2011) Group selection and kin selection: formally equivalent approaches. Trends Ecol Evol 26: 325–332.

Martin FPJ, Wang Y, Sprenger N, Yap IKS, Lundstedt T, Lek P, et al. (2008) Probiotic modulation of symbiotic gut microbial-host metabolic interactions in a humanized microbiome mouse model. Mol Syst Biol 4: 157.

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: 71–75.

Martin MJ, Pérez-Tomé JM, Toro MA (1988) Competition and genotypic variability in Drosophila melanogaster. Heredity 60: 119–123.

Martin OY, Hosken DJ (2003) The evolution of reproductive isolation through sexual conflict. Nature 423: 979–982.

Martin S, Helanterä H, Drijfhout F (2008) Colony-specific hydrocarbons identify nest mates in two species of Formica ant. J Chem Ecol 34: 1072–1080.

Martin SJ, Beekman M, Wossler TC, Ratnieks FLW (2002) Parasitic Cape honeybee workers, Apis mellifera capensis, evade policing. Nature 415: 163–165.

Martin SJ, Carruthers JM, Williams PH, Drijfhout FP (2010) Host specific social parasites (Psithyrus) indicate chemical recognition system in bumblebees. J Chem Ecol 36: 855–863.

Martindale JL, Holbrook NJ (2002) Cellular response to oxidative stress: signaling for suicide and survival. J Cell Physiol 192: 1–15.

Maruo T, Sakamoto H, Iranfar N, Fuller D, Morio T, Urushihara H, Tanaka Y, Maeda M, Loomis WF (2004) Control of cell type proportioning in Dictyostelium discoideum by differentiation-inducing factor as determined by in situ hybridization. Eukaryot Cell 3: 1241–1248.

Marusyk A, Porter CC, Zaberezhnyy V, DeGregori J (2010) Irradiation selects for p53-deficient hematopoietic progenitors. PLoS Biol 8: e1000324.

Marx K (1875/1970) Critique of the Gotha program. In: Marx/Engels Selected Works, Vol. 3. Moscow, USSR: Progress Publishers. pp 13–30.

Maryanovich M, Gross A (2013) A ROS rheostat for cell fate regulation. Trends Cell Biol 23: 129–134.

Mas F, Kölliker M (2008) Maternal care and offspring begging in social insects: chemical signalling, hormonal regulation and evolution. Anim Behav 76: 1121–1131.

Mas F, Haynes KF, Kölliker M (2009) A chemical signal of offspring quality affects maternal care in a social insect. Proc R Soc Lond Ser B Biol Sci 276: 2847–2853.

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: 1106–1110.

Masse E, Gottesman S (2002) A small RNA regulates the expression of genes involved in iron metabolism in Escherichia coli. Proc Natl Acad Sci USA 99: 4620–4625.

Massen JJ, Van den Berg LM, Spruijt BM, Sterck EH (2012) Inequity aversion in relation to effort and relationship quality in long-tailed Macaques (Macaca fascicularis). Am J Primatol 74: 145–156.

Mateo JM (2002) Kin-recognition abilities and nepotism as a function of sociality. Proc R Soc B 269: 721–727.

Mateo JM (2003) Kin recognition in ground squirrels and other rodents. J Mammal 84: 1163–1181.

Mateo JM (2004) Recognition systems and biological organization: the perception component of social recognition. Ann Zool Fenn 41: 729–745.

Matsuda T, Takahashi-Yanaga F, Yoshihara T, Maenaka K, Watanabe Y, Miwa Y, et al. (2010) Dictyostelium differentiation-inducing factor-1 binds to mitochondrial malate dehydrogenase and inhibits its activity. J Pharmacol Sci 112: 320–326.

Matsuura K, Himuro C, Yokoi T, Yamamoto Y, Vargo EL, Keller L (2010) Identification of a pheromone regulating caste differentiation in termites. Proc Natl Acad Sci USA 107: 12963–12968.

Mattila HR, Seeley TD (2007) Genetic diversity in honey bee colonies enhances productivity and fitness. Science 317: 362–364.

Mattila HR, Burke KM, Seeley TD (2008) Genetic diversity within honeybee colonies increases signal production by waggle-dancing foragers. Proc R Soc B Biol Sci 275: 809–816.

Mattila HR, Rios D, Walker-Sperling VE, Roeselers G, Newton IL (2012) Characterization of the active microbiotas associated with honey bees reveals healthier and broader communities when colonies are genetically diverse. PloS ONE 7: e32962.

Matzinger P (2002) The danger model: a renewed sense of self. Science 296: 301–305.

Maughan H, Birky CW Jr, Nicholson WL (2009) Transcriptome divergence and the loss of plasticity in Bacillus subtilis after 6,000 generations of evolution under relaxed selection for sporulation. J Bacteriol 191: 428–433.

Maxson SC (1999) Aggression: Concepts and methods relevant to genetic analysis in mice and humans. In: Jones BC, Mormede P, eds. Neurobehavioral genetics. Methods and applications. Boca Raton, FL: CRC Press. pp 293–300.

Maxson SC (2009) The genetics of offensive aggression in mice. In: Kim KY, ed. Handbook of behavior genetics. New York, NY: Springer. pp 301–316.

Maxson SC, Canastar A (2003) Conceptual and methodological issues in the genetics of mouse agonistic behavior. Horm Behav 44: 258–262.

Mayack C, Naug D (2011) A changing but not an absolute energy budget dictates risk-sensitive behaviour in the honeybee. Anim Behav 82: 595–600.

Maye A, Hsieh C-h, Sugihara G, Brembs B (2007) Order in spontaneous behavior. PLoS ONE 2: e443.

Maynard Smith J (1964) Group selection and kin selection. Nature 201: 1145–1147.

Maynard Smith J (1978) The evolution of sex. Cambridge UK: Cambridge University Press.

Maynard Smith J (1982) Evolution and the theory of games. Cambridge, UK: Cambridge University Press.

Maynard Smith J (1984) Game theory and the evolution of behaviour. Behav Brain Sci 7: 95–125.

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) Evolutionary genetics. Oxford, UK: Oxford University Press.

Maynard Smith J, Szathmáry E (1995) The major transitions in evolution. Oxford, UK: W.H. Freeman/Spektrum.

Mayr E (1961) Cause and effect in biology. Science 134: 1501–1506.

Mayr E (1992) The idea of teleology. J Hist Ideas 53: 117–135.

Mazmanian SK, Liu CH, Tzianabos AO, Kasper DL (2005) An immunomodulatory molecule of symbiotic bacteria directs maturation of the host immune system. Cell 122: 107–118.

Mazur A, Booth A (1998) Testosterone and dominance in men. Behav Brain Sci 21: 353–363, discussion 363–397.

McAuliffe K, Thornton A (2015) The psychology of cooperation in animals: an ecological approach. J Zool 295: 23–35.

McCall C, Mitchell-Olds T, Waller DM (1989) Fitness consequences of outcrossing in Impatiens capensis: tests of the frequency-dependent and sib-competition models. Evolution 43: 1075–1084.

McCall C, Singer T (2012) The animal and human neuroendocrinology of social cognition, motivation and behavior. Nat Neurosci 15: 681–688.

McCauley DM, Wade MJ (1980) Group selection: the genetic and demographic basis for the phenotypic differentiation of small populations of Tribolium castaneum. Evolution 34: 813–821.

McClellan AJ, Tam S, Kaganovich D, Frydman J (2005) Protein quality control: chaperones culling corrupt conformations. Nat Cell Biol 7: 736–741.

McCole SD, Claney K, Conte JC, Anderson R, Hagberg JM (1990) Energy expenditure during bicycling. J Appl Physiol 68: 748–753.

McCorquodale DB (1988) Relatedness among nestmates in a primitively social wasp, Cerceris antipodes (Hymenoptera: Sphecidae). Behav Ecol Sociobiol 23: 401–406.

McCorquodale DB (1989) Nest defense in single- and multifemale nests of Cerceris antipodes (Hymenoptera: Sphecidae). J Insect Behav 2: 267–275.

McCreadie JW, Beard CE, Adler PH (2005) Context-dependent symbiosis between black flies (Diptera: Simuliidae) and trichomycete fungi (Harpellales: Legeriomycetaceae). Oikos 108: 362–370.

McCullough ME, Pedersen EJ, Schroder JM, Tabak BA, Carver CS (2012) Harsh childhood environmental characteristics predict exploitation and retaliation in humans. Proc R Soc B 20122104.

McDonald DB, Potts WK (1994) Cooperative display and relatedness among males in a lek-mating bird. Science 266: 1030–1032.

McDonald PG, Wright J (2011) Bell miner provisioning calls are more similar among relatives and are used by helpers at the nest to bias their effort towards kin. Proc R Soc B 278: 3403–3411.

McFall-Ngai MJ (2002) Unseen forces: the influence of bacteria on animal development. Dev Biol 242: 1–14.

McGowan A, Hatchwell BJ, Woodburn RJW (2003) The effect of helping behaviour on the survival of juvenile and adult long-tailed tits Aegithalos caudatus. J Anim Ecol 72: 491–499.

McGowan A, Sharp SP, Simeoni M, Hatchwell BJ (2006) Competing for position in the communal roosts of long-tailed tits. Anim Behav 72: 1035–1043.

McGraw L, Székely T, Young LJ (2010) Pair bonds and parental behaviour. In: Székely T, Moore AJ, Komdeur J, eds. Social behaviour: genes, ecology and evolution. Cambridge, UK: Cambridge University Press. pp 271–301.

McKechnie AE, Lovegrove BG (2001) Thermoregulation and the energetic significance of clustering behavior in the white-backed mousebird (Colius colius). Physiol Biochem Zool 74: 238–249.

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: 5792–5797.

McNamara JM, Houston AI (1992) Evolutionarily stable levels of vigilance as a function of group size. Anim Behav 43: 641–658.

McNamara JM, Barta Z, Houston AI (2004) Variation in behaviour promotes cooperation in the Prisoner’s Dilemma game. Nature 428: 745–748.

McNamara JM, Leimar O (2010) Variation and the response to variation as a basis for successful cooperation. Phil Trans R Soc Lond B Biol Sci 365: 2627–2633.

McNamara JM, Trimmer PC, Eriksson A, Marshall JA, Houston AI (2011) Environmental variability can select for optimism or pessimism. Ecol Lett 14: 58–62.

McPeek MA, Holt RD (1992) The evolution of dispersal in spatially and temporally varying environments. Am Nat 140: 1010–1027.

McRory JE, Sherwood NM (1997) Ancient divergence of insulin and insulin-like growth factor. DNA Cell Biol 16: 939–949.

McShea WJ (1990) Social tolerance and proximate mechanisms of dispersal among winter groups of meadowvoles, Microtus pennsylvanicus. Anim Behav 39: 346–351.

Mehdiabadi NJ, Reeve HK, Mueller UG (2003) Queens versus workers: sex-ratio conflict in eusocial Hymenoptera. Trends Ecol Evol 18: 88–93.

Mehlis M, Bakker TCM, Frommen JG (2008) Smells like sib spirit: kin recognition in three-spined sticklebacks (Gasterosteus aculeatus) is mediated by olfactory cues. Anim Cogn 11: 643–650.

Meier P, Finch A, Evan G (2000) Apoptosis in development. Nature 407: 796–801.

Melathopoulos AP, Winston ML, Pettis JS, Pankiw T (1996) Effect of queen mandibular pheromone on initiation and maintenance of queen cells in the honey bee (Apis mellifera L.). Can Entomol 128: 263–272.

Melbourne BA, Hasting A (2008) Extinction risk depends strongly on factors contributing to stochasticity. Nature 454: 100–103.

Melis AP, Semmann D (2010) How is human cooperation different? Phil Trans R Soc B 365: 2663–2674.

Mendes P (1997) Biochemistry by numbers: simulation of biochemical pathways with Gepasi 3. Trends Biochem Sci 22: 361–363.

Merkling T, Agdere L, Albert E, Durieux R, Hatch SA, Danchin E, Blanchard P (2014) Is natural hatching asynchrony optimal? An experimental investigation of sibling competition patterns in a facultatively siblicidal seabird. Behav Ecol Sociobiol 68: 309–319.

Merlo LM, Pepper JW, Reid BJ, Maley CC (2006) Cancer as an evolutionary and ecological process. Nat Rev Cancer 6: 924–935.

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.

Merrett NR (1994) Reproduction in the North-Atlantic ichthyofauna and the relationship between fecundity and species sizes. Environ Biol Fish 41: 207–245.

Merryweather JW, Fitter AH (1998) Patterns of arbuscular mycorrhiza colonization of the roots of Hyacinthoides non-scripta after disruption of soil mycelium. Mycorrhiza 8: 87–91.

Mery F, Kawecki TJ (2002) Experimental evolution of learning ability in fruit flies. Proc Natl Acad Sci USA 99: 14274–14279.

Messick DM, Brewer MB (1983) Solving social dilemmas: A review. In: Wheeler L, Shaver P, eds. Review of personality and social psychology. Vol 4. Beverly Hills, CA: Sage. pp 11−44.

Messier F (1985) Solitary living and extraterritorial movements of wolves in relation to social-status and prey abundance. Can J Zool 63: 239–245.

Messina FJ (1991) Life-history variation in a seed beetle: adult egg-laying vs. larval competitive ability. Oecologia 85: 447–455.

Mesterton-Gibbons M, Dugatkin LA (1992) Cooperation among unrelated individuals: evolutionary factors. Q Rev Biol 67: 267–281.

Metcalf RA, Whitt GS (1977) Relative inclusive fitness in the social wasp, Polistes metricus. Behav Ecol Sociobiol 2: 353–360.

Metheny JD, Kalcounis-Rueppell MC, Willis CK, Kolar KA, Brigham RM (2008) Genetic relationships between roost-mates in a fission–fusion society of tree-roosting big brown bats (Eptesicus fuscus). Behav Ecol Sociobiol 62: 1043–1051.

Metzgar LH (1967) An experimental comparison of screech owl predation on resident and transient white-footed mice (Peromyscus leucopus). J Mammal 48: 387–391.

Meunier J, Kölliker M (2012) When it is costly to have a caring mother: food limitation erases the benefits of parental care in earwigs. Biol Lett 8: 547–550.

Meyer JM, Neely A, Stintzi A, Georges C, Holder IA (1996) Pyoverdine is essential for virulence of Pseudomonas aeruginosa. Infect Immun 64: 518–523.

Meyer JR, Kassen R (2007) The effects of competition and predation on diversification in a model adaptive radiation. Nature 446:432–435.

Meznar ER, Gadau J, Koeniger N, Rueppell O (2010) Comparative linkage mapping suggests a high recombination rate in all honey bees. J Hered 101: S118–S126.

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: 301–311.

Michaelson J (1987) Cell selection in development. Biol Rev Camb Philos Soc 62: 115–139.

Michaelson J (1993) Cellular selection in the genesis of multicellular organization. Lab Invest 69: 136–151.

Michaud JP, Grant AK (2004) Adaptive significance of sibling egg cannibalism in Coccinellidae: comparative evidence from three species. Ann Entomol Soc Am 97: 710–719.

Michener CD (1964) Reproductive efficiency in relation to colony size in hymenopterous societies. Insect Soc 11: 317–342.

Michener CD (1969) Comparative social behavior of bees. Annu Rev Entomol 14: 299–342.

Michener CD (1974) The social behavior of bees. Cambridge, MA: Belknap Press.

Michener CD (1990) Reproduction and castes in social halictine bees. In: Engels W, ed. Social insects: an evolutionary approach to castes and reproduction. Berlin, Germany: Springer-Verlag. pp 77–121.

Michener CD (2007) The bees of the world. 2nd edn. Baltimore, MD: Johns Hopkins University Press.

Michener CD, Rettenmeyer CW (1956) The ethology of Andrena erythronii with comparative data on other species (Hymenoptera, Andrenidae). Univ Kans Sci Bull 37: 645–684.

Michener CD, Brothers DJ (1974) Were workers of eusocial Hymenoptera initially altruistic or oppressed? Proc Natl Acad Sci USA 71: 671–674.

Michener GR (1983) Kin identification, matriarchies, and the evolution of sociality in ground-dwelling sciurids. In: Eisenberg JF, Kleiman DG, eds. Advances in the study of mammalian behavior. Lawrence, Kansas: The American Society of Mammalogists. pp 528–572.

Michod RE (1996) Cooperation and conflict in the evolution of individuality. II. Conflict mediation. Proc R Soc Lond B 263: 813–822.

Michod RE (1999) Darwinian dynamics. Princeton, NJ: Princeton University Press.

Michod RE (2005) On the transfer of fitness from the cell to the multicellular organism. Biol Philos 20: 967–987.

Michod RE, Roze D (2001) Cooperation and conflict in the evolution of multicellularity. Heredity 86: 1–7.

Michopoulos V, Checchi M, Sharpe D, Wilson ME (2011) Estradiol effects on behavior and serum oxytocin are modified by social status and polymorphisms in the serotonin transporter gene in female rhesus monkeys. Horm Behav 59: 528–535.

Michopoulos V, Higgins M, Toufexis D, Wilson ME (2012) Social subordination produces distinct stress-related phenotypes in female rhesus monkeys. Psychoneuroendocrinology 37: 1071–1085.

Michor F, Iwasa Y, Nowak MA (2004) Dynamics of cancer progression. Nat Rev Canc 4: 197–205.

Miczek KA, Maxson SC, Fish EW, Faccidomo S (2001) Aggressive behavioural phenotypes in mice. Behav Brain Res 125: 167–181.

Miczek KA, Fish EW, Joseph F, de Almeida RM (2002) Social and neural determinants of aggressive behavior: pharmacotherapeutic targets at serotonin, dopamine and γ-aminobutyric acid systems. Psychopharmacology 163: 434–458.

Miczek KA, Fish EW (2005) Monoamines, GABA, glutamate and aggression. In: Nelson RJ, ed. Biology of aggression. New York, NY: Oxford University Press. pp 114–150.

Mikics E, Kruk MR, Haller J (2004) Genomic and nongenomic effects of glucocorticoids on aggressive behavior in male rats. Psychoneuroendocrinology 29: 618–635.

Mikkelsen R, Wardman P (2003) Biological chemistry of reactive oxygen and nitrogen and radiation-induced signal transduction mechanisms. Oncogene 22: 5734–5754.

Mikolajczak M, Pinon N, Lane A, de Timary P, Luminet O (2010) Oxytocin not only increases trust when money is at stake, but also when confidential information is in the balance. Biol Psychol 85: 182–184.

Milinski M (2006) The major histocompatibility complex, sexual selection, and mate choice. Annu Rev Ecol Evol Syst 37: 159–186.

Milinski M, Semmann D, Krambeck HJ (2002) Reputation helps solve the ‘tragedy of the commons’. Nature 415: 424–426.

Miller AD, Roxburgh SH, Shea K (2011) How frequency and intensity shape diversity–disturbance relationships. Proc Natl Acad Sci USA 108: 5643–5648.

Miller DG (1998a) Consequences of communal gall occupation and a test for kin discrimination in the aphid Tamalia coweni (Cockerell) (Homoptera: Aphididae). Behav Ecol Sociobiol 43: 95–103.

Miller DG (1998b) Life history, ecology and communal gall occupation in the manzanita leaf-gall aphid, Tamalia coweni (Cockerell) (Homoptera: Aphididae). J Nat Hist 32: 351–366.

Miller DG (2004) The ecology of inquilinism in communally parasitic Tamalia aphids (Hemiptera: Aphididae). Ann Entomol Soc Am 97: 1233–1241.

Miller G, Shulaev V, Mittler R (2008) Reactive oxygen signaling and abiotic stress. Physiol Plant 133: 481–489.

Miller GF (1997) Protean primates: The evolution of adaptive unpredictability in competition and courtship. In: Whiten A, Byrne R, eds. Machiavellian intelligence II: Extensions and evaluations. Cambridge, UK: Cambridge University Press. pp 312–340.

Miller KD (1996) Synaptic economics: competition and cooperation in synaptic plasticity. Neuron 17: 371–374.

Miller PL (1966) The supply of oxygen to the active flight muscles of some large beetles. J Exp Biol 45: 285–304.

Milo R, Shen-Orr S, Itzkovitz S, Kashtan N, Chklovskii D, Alon U (2002) Network motifs: simple building blocks of complex networks. Science 298: 824–827.

Minakata H (2010) Oxytocin/vasopressin and gonadotropin-releasing hormone from cephalopods to vertebrates. Ann NY Acad Sci 1200: 33–42.

Minckley RL, Wcislo WT, Yanega D, Buchmann SL (1994) Behavior and phenology of a specialist bee (Dieunomia) and sunflower (Helianthus) pollen availability. Ecology 75: 1406–1419.

Minocherhomji S, Tollefsbol TO, Singh KK (2012) Mitochondrial regulation of epigenetics and its role in human diseases. Epigenetics 7: 326–334.

Miralles R, Moya A, Elena SF (1997) Is group selection a factor in modulating the virulence of RNA viruses? Genet Res 69: 165–172.

Miramontes O, DeSouza O (2014) Social evolution: new horizons. arXiv preprint arXiv:1404.6267.

Mirimin L, Banguera-Hinestroza E, Dillane E, Hoelzel AR, Cross TF, Rogan E (2011) Insights into genetic diversity, parentage, and group composition of Atlantic white-sided dolphins (Lagenorhynchus acutus) off the west of Ireland based on nuclear and mitochondrial genetic markers. J Hered 102: 79–87.

Mirouze N, Prepiak P, Dubnau D (2011) Fluctuations in spo0A transcription control rare developmental transitions in Bacillus subtilis. PLoS Genet 7: e1002048.

Mirth CK, Riddiford LM (2007) Size assessment and growth control: how adult size is determined in insects. Bioessays 29: 344–355.

Mitani JC (2006) Reciprocal exchange in chimpanzees and other primates. In: Kappeler P, van Schaik C, eds. Cooperation in primates: mechanisms and evolution. Heidelberg, Germany: Springer-Verlag. pp 101–113.

Mitani JC (2009) Male chimpanzees form enduring and equitable social bonds. Anim Behav 77: 633–640.

Mitchell J (2005) Queue selection and switching by false clown anemonefish Amphiprion ocellaris. Anim Behav 69: 643–652.

Mitchell RJ (1994) Effects of floral traits, pollinator visitation and plant size on Ipomopsis aggregata fruit production. Am Nat 143: 870–889.

Mitchell SD (2009) Unsimple truths: science, complexity and policy. Chicago, IL: University of Chicago Press.

Mitsui K, Tokuzawa Y, Itoh H, Segawa K, Murakami M, Takahashi K, et al. (2003) The homeoprotein Nanog is required for maintenance of pluripotency in mouse epiblast and ES cells. Cell 113: 631–642.

Mittal S, Kroos L (2009a) A combination of unusual transcription factors binds cooperatively to control Myxococcus xanthus developmental gene expression. Proc Natl Acad Sci USA 106: 1965–1970.

Mittal S, Kroos L (2009b) Combinatorial regulation by a novel arrangement of FruA and MrpC2 transcription factors during Myxococcus xanthus development. J Bacteriol 191: 2753–2763.

Mittelbach GG (1981) Foraging efficiency and body size: a study of optimal diet and habitat use by bluegills. Ecology 62: 1370–1386.

Mittelbach GG (2012) Community ecology. Sunderland, MA: Sinauer Associates, Inc.

Mitteldorf J, Wilson DS (2000) Population viscosity and the evolution of altruism. J Theor Biol 204: 481–496.

Mittler R (2002) Oxidative stress, antioxidants and stress tolerance. Trends Plant Sci 7: 405–410.

Miyanaga R, Maeta Y, Sakagami SF (1999) Geographical variation of sociality and size-linked color patterns in Lasioglossum (Evylaeus) apristum (Vachal) in Japan (Hymenoptera, Halictidae). Insect Soc 46: 224–232.

Miyanaga Y, Matsuoka S, Yanagida T, Ueda M (2007) Stochastic signal inputs for chemotactic response in Dictyostelium cells revealed by single molecule imaging techniques. Biosystems 88: 251–260.

Mock DW, Forbes LS (1992) Parent-offspring conflict: a case of arrested development? Trends Ecol Evol 7: 409–413.

Mock DW, Forbes LS (1995) The evolution of parental optimism. Trends Ecol Evol 10: 130–134.

Mock DW, Parker GA (1997) The evolution of sibling rivalry. Oxford, UK: Oxford University Press.

Mock DW, Parker GA, Schwagmeyer PL (1998) Game theory, sibling rivalry, and parent-offspring conflict. In: Dugatkin LA, Reeve HK, eds. Game theory and animal behavior. Oxford, UK: Oxford University Press. pp 146–167.

Modlmeier AP, Foitzik S (2011) Productivity increases with variation in aggression among group members in Temnothorax ants. Behav Ecol 22: 1026–1032.

Modlmeier AP, Liebmann JE, Foitzik S (2012) Diverse societies are more productive: a lesson from ants. Proc R Soc B Biol Sci 279: 2142–2150.

Moehlman PD (1979) Jackal helpers and pup survival. Nature 277: 382–383.

Moehlman PD (1986) Ecology of cooperation in canids. In: Rubenstein DI, Wrangham RW, eds. Ecological aspects of social evolution: birds and mammals. Princeton, NJ: Princeton University Press. pp 64–86.

Mohammedi A, Paris A, Crauser D, Le Conte Y (1998) Effect of aliphatic esters on ovary development of queenless bees (Apis mellifera L). Naturwissenschaften 85: 455–458.

Mohammedi A, Le Conte Y (2000) Do environmental conditions exert an effect on nest-mate recognition in queen rearing honey bees? Insect Soc 47: 307–312.

Møller AP, Pagel M (1998) Developmental stability and signalling among cells. J Theor Biol 193: 497–506.

Möller LM, Beheregaray LB, Harcourt RG, Krützen M (2001) Alliance membership and kinship in wild male bottlenose dolphins (Tursiops aduncus) of southeastern Australia. Proc R Soc Lond Ser B Biol Sci 268: 1941–1947.

Momeni B, Waite AJ, Shou W (2013) Spatial self-organization favors heterotypic cooperation over cheating. eLife 2: e00960.

Monds RD, O’Toole GA (2009) Developmental models of microbial biofilms: ten years of paradigm up for review. Trends Microbiol 17: 73–87.

Monnin T (2006) Chemical recognition of reproductive status in social insects. Ann Zool Fenn 43: 515–530.

Monnin T, Peeters C (1997) Cannibalism of subordinates’ eggs in the monogynous queenless ant Dinoponera quadriceps. Naturwissenschaften 84: 499–502.

Monnin T, Peeters C (1998) Monogyny and the regulation of worker mating in the queenless ant Dinoponera quadriceps. Anim Behav 55: 299–306.

Monnin T, Peeters C (1999) Dominance hierarchy and reproductive conflicts among subordinates in a monogynous queenless ant. Behav Ecol 10: 323–332.

Monnin T, Ratnieks FLW (1999) Reproduction versus work in queenless ants: when to join a hierarchy of hopeful reproductives? Behav Ecol Sociobiol 46: 413–422.

Montague PR (1996) The resource consumption principle: attention and memory in volumes of neural tissue. Proc Natl Acad Sci USA 93: 3619–3623.

Moore AJ, Brodie ED 3rd, Wolf JB (1997) Interacting phenotypes and the evolutionary process: I. Direct and indirect genetic effects of social interactions. Evolution 51: 1352–1362.

Mooring MS, Hart BL (1992) Animal grouping for protection from parasites: selfish herd and encounter-dilution effects. Behaviour 123: 173–193.

Moosa MM, Ud-Dean SMM (2011) The role of dominance hierarchy in the evolution of social species. J Theor Soc Behav 41: 203–208.

Moran NA (2007) Symbiosis as an adaptive process and source of phenotypic complexity. Proc Natl Acad Sci 104 (Suppl 1): 8627–8633.

Moran NA, Wernegreen JJ (2000) Lifestyle evolution in symbiotic bacteria: insights from genomics. Trends Ecol Evol 15: 321–326.

Morand-Ferron J, Giraldeau LA, Lefebvre L (2007) Wild Carib grackles play a producer-scrounger game. Behav Ecol 18: 916–921.

Morandini V, Ferrer M (2014) Sibling aggression and brood reduction: a review. Ethol Ecol Evol [ahead-of-print] 1-15. DOI:10.1080/03949370.2014.880161.

Morata G, Ripoll P (1975) Minutes: mutants of Drosophila autonomously affecting cell division rate. Dev Biol 42: 211–221.

Morata G, Martin FA (2007) Cell competition: the embrace of death. Dev Cell 13: 1–2.

Moreira da Silva F, Marques A, Chaveiro A (2010) Reactive oxygen species: a double-edged sword in reproduction. Open Vet Sci J 4: 127–133.

Moreno E (2008) Is cell competition relevant to cancer? Nat Rev Cancer 8: 141–147.

Moreno E, Basler K, Morata G (2002) Cells compete for decapentaplegic survival factor to prevent apoptosis in Drosophila wing development. Nature 416: 755–759.

Moreno E, Basler K (2004) dMyc transforms cells into super-competitors. Cell 117: 117–129.

Moretz JA, Martins EP, Robison BD (2007) The effects of early and adult social environment on zebrafish (Danio rerio) behavior. Env Biol Fish 80: 91–101.

Morgan AD, Buckling A (2004) Parasites mediate the relationship between host diversity and disturbance frequency. Ecol Lett 7: 1029–1034.

Morgan MJ, Godin JGJ (1985) Antipredator benefits of schooling behaviour in a cyprinodontid fish, the banded killifish (Fundulus diaphanus). Z Tierpsychol 70: 236–246.

Mori Y (1999) A note on swimming group size in captive African penguins (Spheniscus demersus) in relation to weather conditions. Appl Anim Behav Sci 62: 359–364.

Morin PJ, Lawler SP, Johnson FA (1990) Ecology and breeding phenology of larval Hyla andersonii: the disadvantages of breeding late. Ecology 71: 1590–1598.

Morris DC, Schwarz MP, Crespi BJ (2002) Pleometrosis in phyllode-glueing thrips (Thysanoptera: Phlaeothripidae) on Australian Acacia. Biol J Linn Soc 75: 467–474.

Morris DW (1991) On the evolutionary stability of dispersal to sink habitats. Am Nat 137: 907–911.

Morris JJ, Lenski RE, Zinser ER (2012) The Black Queen hypothesis: evolution of dependencies through adaptive gene loss. mBio 3: e00036-12.

Moses RA, Millar JS (1994) Philopatry and mother–daughter associations in bushy-tailed woodrats – space use and reproductive success. Behav Ecol Sociobiol 35: 131–140.

Moss CJ, Lee PC (2011) Female reproductive strategies and social relationships. In: Moss CJ, Croze H, Lee PC, eds. The Amboseli elephants. Chicago, IL: University of Chicago Press. pp 205–213.

Motro U (1983) Optimal rates of dispersal 3. Parent offspring conflict. Theor Popul Biol 23: 159–168.

Mottley K, Giraldeau LA (2000) Experimental evidence that group foragers can converge on predicted producer-scrounger equilibria. Anim Behav 60: 341–350.

Moussaid M, Garnier S, Theraulaz G, Helbing D (2009) Collective information processing and pattern formation in swarms, flocks, and crowds. Topics Cogn Sci 1: 469–497.

Moussaid M, Niriaska P, Garnier S, Helbing D, Theraulaz G (2010) The walking behaviour of pedestrian social groups and its impact on crowd dynamics. PLoS ONE 5: e10047.

Moyed HS, Bertrand KP (1983) hipA, a newly recognized gene of Escherichia coli K-12 that affects frequency of persistence after inhibition of murein synthesis. J Bacteriol 155: 768–775.

Moyer KE (1968) Kinds of aggression and their physiological basis. Comm Behav Biol 2: 65–87.

Mrak P, Podlesek Z, van Putten JPM, Zgur-Bertok D (2007) Heterogeneity in expression of the Escherichia coli colicin K activity gene cka is controlled by the SOS system and stochastic factors. Mol Genet Genomics 277: 391–401.

Msadek T (1999) When the going gets tough: survival strategies and environmental signaling networks in Bacillus subtilis. Trends Microbiol 7: 201–207.

Mueller L, Guo P, Ayala F (1991) Density-dependent natural selection and trade-offs in life history traits. Science 253: 433–435.

Mueller UG (1991) Haplodiploidy and the evolution of facultative sex ratios in a primitively eusocial bee. Science 254: 442–444.

Mueller UG, Ishak H, Lee JC, Sen R, Gutell RR (2010) Placement of attine ant-associated Pseudonocardia in a global Pseudonocardia phylogeny (Pseudonocardiaceae, Actinomycetales): A test of two symbiont-association models. Anton Leeuw Int J G 98: 195–212.

Mukai H, Watanabe H (1975) Distribution of fusion incompatibility types in natural populations of compound ascidian, Botryllus primigenus. Proc Jpn Acad 51: 44–47.

Mulec J, Podlesek Z, Mrak P, Kopitar A, Ihan A, Zgur-Bertok D (2003) A cka–gfp transcriptional fusion reveals that the colicin K activity gene is induced in only 3 percent of the population. J Bacteriol 185: 654–659.

Mull JF, MacMahon JA (1997) Spatial variation in rates of seed removal by harvester ants (Pogonomyrmex occidentalis) in a shrub-steppe ecosystem. Am Midl Nat 138: 1–13.

Müller AE, Thalmann U (2000) Origin and evolution of primate social organisation: a reconstruction. Biol Rev Camb Philos Soc 75: 405–435.

Müller GB (2007) Evo–devo: extending the evolutionary synthesis. Nat Rev Genet 8: 943–949.

Mumme RL (1997) A bird’s eye view of mammalian cooperative breeding. In: Solomon NG, French JA, eds. Cooperative breeding in mammals. Cambridge, UK: Cambridge University Press. pp 364–388.

Mumme RL, Koenig WD, Pitelka FA (1983) Reproductive competition in the communal acorn woodpecker—sisters destroy each others eggs. Nature 306: 583–584.

Mumme RL, Koenig WD, Ratnieks FLW (1989) Helping behaviour, reproductive value, and the future component of indirect fitness. Anim Behav 38: 331–343.

Munkvold L, Kjøller R, Vestberg M, Rosendahl S, Jakobsen I (2004) High functional diversity within species of arbuscular mycorrhizal fungi. New Phytol 164: 357–364.

Munson EL, Pfaller MA, Doern GV (2002) Modification of dienes mutual inhibition test for epidemiological characterization of Pseudomonas aeruginosa isolates. J Clin Microbiol 40: 4285–4288.

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: 903–910.

Muraille E (2013) Redefining the immune system as a social interface for cooperative processes. PLoS Pathog 9: e1003203.

Murase Y, Shimada T, Ito N, Rikvold PA (2010) Effects of demographic stochasticity on biological community assembly on evolutionary time scales. Phys Rev E 81: 041908.

Murphy CM, Breed MD (2007) A predictive distribution map for the giant tropical ant, Paraponera clavata. J Insect Sci 7: 1–10.

Murphy K (2012) Janeway’s immunobiology, 8th edn. New York, NY: Garland Science.

Murphy TF, Kirkham C, Sethi S, Lesse AJ (2005) Expression of a peroxiredoxin-glutaredoxin by Haemophilus influenzae in biofilms and during human respiratory tract infection. FEMS Immunol Med Microbiol 44: 81–89.

Murray MG, Gerrard RJ (1984) Conflict in the neighbourhood: models where close relatives are in direct competition. J Theor Biol 111: 237–246.

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: 105–137.

Mutti NS, Dolezal AG, Wolschin F, Mutti JS, Gill KS, Amdam GV (2011) IRS and TOR nutrient-signaling pathways act via juvenile hormone to influence honey bee caste fate. J Exp Biol 214: 3977–3984.

Myerscough MR, Oldroyd BP (2004) Simulation models of the role of genetic variability in social insect task allocation. Insect Soc 51: 146–152.

Myles TG (1988) Resource inheritance in social evolution from termites to man. In: Slobodchikoff CN, ed. The ecology of social behavior. San Diego, CA: Academic. pp 379–423.

Nadell CD, Xavier JB, Foster KR (2009) The sociobiology of biofilms. FEMS Microbiol Rev 33: 206–224.

Nadell CD, Bucci V, Drescher K, Levin SA, Bassler BL, Xavier JB (2013) Cutting through the complexity of cell collectives. Proc R Soc B 280: 20122770.

Nadin BM, Mah CS, Scharff JR, Ratner AI (2000) The regulative capacity of prespore amoebae as demonstrated by fluorescence-activated cell sorting and green fluorescent protein. Dev Biol 217: 173–178.

Nair VD, Yuen T, Olanow CW, Sealfon SC (2004) Early single cell bifurcation of pro- and antiapoptotic states during oxidative stress. J Biol Chem 279: 27494–27501.

Naisbit RE, Jiggins CD, Mallet J (2001) Disruptive sexual selection against hybrids contributes to speciation between Heliconius cydno and Heliconius melpomene. Proc R Soc Lond Ser B Biol Sci 268: 1849–1854.

Nakagawa S, Waas J, Miyazaki M (2001) Heart rate changes reveal that little blue penguin chicks Eudyptula minor can use vocal signatures to discriminate familiar from unfamiliar chicks. Behav Ecol Sociobiol 50: 180–188.

Nakamaru M, Matsuda H, Iwasa Y (1997) The evolution of cooperation in a lattice-structured population. J Theor Biol 184: 65–81.

Nakamura RR (1980) Plant kin selection. Evol Theory 5: 113–117.

Nakanishi A, Nishino H, Watanabe H, Yokohari F, Nishikawa M (2010) Sex-specific antennal sensory system in the ant Camponotus japonicus: glomerular organizations of antennal lobes. J Comp Neurol 518: 2186–2201.

Nam KB, Simeoni M, Sharp SP, Hatchwell BJ (2010) Kinship affects investment by helpers in a cooperatively breeding bird. Proc R Soc B 277: 3299–3306.

Narbonne P, Roy R (2006) Regulation of germline stem cell proliferation downstream of nutrient sensing. Cell Div 1: 29–38.

Nariya H, Inouye M (2008) MazF, an mRNA interferase, mediates programmed cell death during multicellular Myxococcus development. Cell 132: 55–66.

Narula J, Devi SN, Fujita M, Igoshin OA (2012) Ultrasensitivity of the Bacillus subtilis sporulation decision. Proc Natl Acad Sci USA 109: E3513–E3522.

Nash DR, Boomsma JJ (2008) Communication between hosts and social parasites. In: d’Ettorre P, Hughes DP, eds. Sociobiology of communication: an interdisciplinary perspective. Oxford, UK: Oxford University Press. pp 55–79.

Naumov GI, Naumova ES, Sancho ED, Korhola MP (1996) Polymeric SUC genes in natural populations of Saccharomyces cerevisiae. FEMS Microbiol Lett 135: 31–35.

Navarro Llorens JM, Tormo A, Martinez-Garcia E (2010) Stationary phase in gramnegative bacteria. FEMS Microbiol Rev 34: 476–495.

Naviaux RK (2008) Mitochondrial control of epigenetics. Cancer Biol Ther 7: 1191–1193.

Neander K (1991) The teleological notion of ‘function’. Australas J Philos 69: 454–468.

Nedelcu AM, Michod RE (2006) The evolutionary origin of an altruistic gene. Mol Biol Evol 23: 1460–1464.

Nedelcu AM, Driscoll WW, Durand PM, Herron MD, Rashidi A (2011) On the paradigm of altruistic suicide in the unicellular world. Evolution 65: 3–20.

Neff BD (2003a) Decisions about parental care in response to perceived paternity. Nature 422: 716–719.

Neff BD (2003b) Paternity and condition affect cannibalistic behavior in nest-tending bluegill sunfish. Behav Ecol Sociobiol 54: 377–384.

Neff BD, Gross MR (2001) Dynamic adjustment of parental care in response to perceived paternity. Proc R Soc Lond B 268: 1559–1565.

Neff BD, Cargnelli LM (2004) Relationships between condition factors, parasite load and paternity in bluegill sunfish, Lepomis macrochirus. Environ Biol Fish 71: 297–304.

Neill StSRJ, Cullen JM (1974) Experiments on whether schooling by their prey affects the hunting behaviour of cephalopod and fish predators. J Zool 172: 549–569.

Nelson RJ, Chiavegatto S (2001) Molecular basis of aggression. Trends Neurosci 24: 713–719.

Nelson RJ, Trainor BC (2007) Neural mechanisms of aggression. Nat Rev Neurosci 8: 536–546.

Nentwig W (1985) Social spiders catch larger prey: a study of Anelosimus eximius (Araneae: Theridiidae). Behav Ecol Sociobiol 17: 79–85.

Nettle D (2010) Dying young and living fast: variation in life history across English neighborhoods. Behav Ecol 21: 387–395.

Neuhauser C, Fargione JE (2004) A mutualism-parasitism continuum model and its application to plant mycorrhizae interactions. Ecol Model 177: 337–352.

Neumann ID (2008) Brain oxytocin: a key regulator of emotional and social behaviours in both females and males. J Neuroendocrinol 20: 858–865.

Neumann ID (2009) The advantage of social living: Brain neuropeptides mediate the beneficial consequences of sex and motherhood. Front Neuroendocrinol 30: 483–496.

Neumann P, Moritz RFA, Mautz D (2000) Colony evaluation is not affected by drifting of drone and worker honeybees (Apis mellifera L.) at a performance testing apiary. Apidologie 31: 67–79.

Neumann P, Radloff SE, Moritz RFA, Hepburn HR, Reece SL (2001) Social parasitism by honeybee workers (Apis mellifera capensis Escholtz): host finding and resistance of hybrid host colonies. Behav Ecol 12: 419–428.

Newman C (2000) The demography and parasitology of the WythamWoods badger population. PhD thesis. Oxford, UK: Oxford University.

Newman RA (1989) Developmental plasticity of Scaphiopus couchii tadpoles in an unpredictable environment. Ecology 70: 1175–1787.

Newsham KK, Fitter AH, Watkinson AR (1995) Multi-functionality and biodiversity in arbuscular mycorrhizas. Trends Ecol Evol 10: 407–411.

Ng WL, Bassler BL (2009) Bacterial quorum-sensing network architectures. Annu Rev Genet 43: 197–222.

Nicolis G (1995) Introduction to nonlinear science. Cambridge, UK: Cambridge University Press.

Nicolis G, Prigogine I (1977) Self-organization in nonequilibrium systems: from dissipative structures to order through fluctuations. New York, NY: John Wiley & Sons.

Niehuis O, Gibson JD, Rosenberg MS, Pannebakker BA, Koevoets T, et al. (2010) Recombination and its impact on the genome of the haplodiploid parasitoid wasp Nasonia. PLoS ONE 5: e8597.

Niemitz C (2002) A theory on the evolution of the habitual orthograde human bipedalism – The "Amphibische Generalistentheorie". Anthropol Anz 60: 3–66.

Nijhout HF (1994) Insect hormones. Princeton, NJ: Princeton University Press.

Nilsson J, D’Hertefeldt T (2008) Origin matters for level of resource sharing in the clonal herb Aegopodium podagraria. Evol Ecol 22: 437–448.

Nisenbaum LK (2002) The ultimate chip shot: can microarray technology deliver for neuroscience? Genes Brain Behav 1: 27–34.

Nishiguchi MK, Nair VS (2003) Evolution of symbiosis in the Vibroonaceae: A combined approach using molecules and physiology. Int J Syst Evol Microbiol 53: 2019–2026.

Noble JC, Marshall C (1983) The population biology of plants with clonal growth. II. The nutrient strategy and modular physiology of Carex arenaria. J Ecol 71: 865–877.

Noë R (1990) A veto game played by baboons: a challenge to the use of the Prisoner’s Dilemma as a paradigm for reciprocity and cooperation. Anim Behav 39: 78–90.

Noë R (2001) Biological markets: partner choice as the driving force behind the evolution of mutualisms. In: Noë R, van Hooff JARAM, Hammerstein P, eds. Economics in nature: Social dilemmas, mate choice and biological markets. UK: Cambridge University Press. pp 93–118.

Noë R (2006) Cooperation experiments: Coordination through communication versus acting apart together. Anim Behav 71: 1–18.

Noë R, Van Schaik CP, Van Hoof JARAM (1991) The market effect: an explanation for pay-off asymmetries among collaborating animals. Ethology 87: 97–118.

Noë R, Hammerstein P (1994) Biological markets: Supply and demand determine the effect of partner choice in cooperation, mutualism and mating. Behav Ecol Sociobiol 35: 1–11.

Noë R, Hammerstein P (1995) Biological markets. Trends Ecol Evol 10: 336–339.

Noë R, Van Hoof JARAM, Hammerstein P (2001) Economics in nature: social dilemmas, mate choice and biological markets. Cambridge, UK: Cambridge University Press.

Nonacs P (1988) Queen number in colonies of social Hymenoptera as a kin-selected adaptation. Evolution 42: 566–580.

Nonacs P (2011) Monogamy and high relatedness do not preferentially favor the evolution of cooperation. BMC Evol Biol 11: 58.

Nonacs P, Reeve HK (1995) The ecology of cooperation in wasps: causes and consequences of alternative reproductive decisions. Ecology 76: 953–967.

Nonacs P, Liebert AE, Starks PT (2006) Transactional skew and assured fitness return models fail to predict patterns of cooperation in wasps. Am Nat 167: 467–480.

Nonacs P, Kapheim KM (2007) Social heterosis and the maintenance of genetic diversity. J Evol Biol 20: 2253–2265.

Nonacs P, Kapheim KM (2008) Social heterosis and the maintenance of genetic diversity at the genome level. J Evol Biol 21: 631–635.

Noonan KM (1981) Individual strategies of inclusive-fi tness-maximizing in Polistes fuscatus foundresses. In: Alexander RD, Tinkle DW, eds. Natural selection and social behavior: recent research and theory. New York, NY: Chiron. pp 18–44.

Noriyuki S, Osawa N, Nishida T (2011) Prey capture performance in hatchlings of two sibling Harmonia ladybird species in relation to maternal investment through sibling cannibalism. Ecol Entomol 36: 282–289.

Noriyuki S, Kawatsu K, Osawa N (2012) Factors promoting maternal trophic egg provisioning in non-eusocial animals. Popul Ecol 54: 455–465.

Norman GJ, Hawkley LC, Cole SW, Berntson GG, Cacioppo JT (2012) Social neuroscience: The social brain, oxytocin, and health. Soc Neurosci 7: 18–29.

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.

Nosil P (2004) Reproductive isolation caused by visual predation on migrants between divergent environments. Proc R Soc Lond Ser B Biol Sci 271: 1521–1528.

Nosil P, Crespi BJ (2006) Experimental evidence that predation promotes divergence in adaptive radiation. Proc Natl Acad Sci USA 103: 9090–9095.

Nowak M (1990) Stochastic strategies in the Prisoner’s Dilemma. Theor Popul Biol 38: 93–112.

Nowak MA (2006a) Five rules for the evolution of cooperation. Science 314: 1560–1563.

Nowak MA (2006b) Evolutionary dynamics: exploring the equations of life. Cambridge, MA: Belknap/Harvard University Press.

Nowak M, Sigmund K (1989) Oscillations in the evolution of reciprocity. J Theor Biol 137: 21–26.

Nowak MA, Sigmund K (1990) The evolution of stochastic strategies in the Prisoner’s Dilemma. Acta Appl Math 20: 247–265.

Nowak MA, May RM (1992) Evolutionary games and spatial chaos. Nature 359: 826–829.

Nowak M, Sigmund K (1992) Tit for tat in heterogeneous populations. Nature 355: 250–252.

Nowak M, Sigmund K (1993a) Chaos and the evolution of cooperation. Proc Natl Acad Sci USA 90: 5091–5094.

Nowak M, Sigmund K (1993b) A strategy of win-stay, lose-shift that outperforms tit-for-tat in the Prisoner's Dilemma game. Nature 364: 56–58.

Nowak M, Bonhoeffer S, May R (1994) Spatial games and the maintenance of cooperation. Proc Natl Acad Sci USA 91: 4877–4881.

Nowak MA, Sigmund K, Elsedy E (1995) Automata, repeated games and noise. J Math Biol 33: 703–722.

Nowak MA, Sigmund K (1998) Evolution of indirect reciprocity by image scoring. Nature 393: 573–577.

Nowak MA, Sigmund K (2004) Evolutionary dynamics of biological games. Science 303: 793–799.

Nowak MA, Sasaki A, Taylor C, Fudenberg D (2004) Emergence of cooperation and evolutionary stability in finite populations. Nature 428: 646–650.

Nowak MA, Sigmund K (2005) Evolution of indirect reciprocity. Nature 437: 1291–1298.

Nowak MA, Tarnita CE, Wilson EO (2010) The evolution of eusociality. Nature 466: 1057–1062.

Nowak MA, Tarnita CE, Wilson EO (2011) Nowak et al. reply. Nature 471: E9–E10.

Nowak MA, Highfield R (2011) SuperCooperators: altruism, evolution, and why we need each other to succeed. New York, NY: Free Press.

Nowell PC (1976) The clonal evolution of tumor cell populations. Science 194: 23–28.

Nudds TD (1978) Convergence of group size strategies by mammalian social carnivores. Am Nat 112: 957–960.

Numan M, Insel TR (2003) The neurobiology of parental behavior. New York, NY: Springer.

Nunes S (2007) Dispersal and philopatry. In: Wolff JO, Sherman PW, eds. Rodent societies: an ecological & evolutionary perspective. Chicago, IL: University of Chicago Press. pp 150–162.

Nunney L (1985) Group selection, altruism, and structured-deme models. Am Nat 126: 212–230.

Nunney L (1999) Lineage selection: natural selection for long-term benefit. In: Keller L, ed. Levels of selection in evolution. Princeton, NJ: Princeton University Press. pp 238–252.

Obin MS (1986) Nestmate recognition cues in laboratory and field colonies of Solenopsis invicta Buren (Hymenoptera: Formicidae): effect of environment and role of cuticular hydrocarbons. J Chem Ecol 12: 1965–1975.

Obin MS, Vander Meer RK (1988) Source of nestmate recognition cues in the imported fire ant Solenopsis invicta Buren (Hymenoptera: Formicidae). Anim Behav 36: 1361–1370.

Oborny B, Kun Á, Czárán T, Bokros S (2000) The effect of clonal integration on plant competition for mosaic habitat space. Ecology 81: 3291–3304.

O’Brien H, Miadlikowska J, and Lutzoni F (2005) Assessing host specialization in symbiotic cyanobacteria associated with four closely related species of the lichen Peltigera. Eur J Phycol 40: 363–378.

Ochsner UA, Vasil AI, Johnson Z, Vasil ML (1999) Pseudomonas aeruginosafur overlaps with a gene encoding a novel outer membrane lipoprotein, OmlA. J Bacteriol 181: 1099–1109.

Ochsner UA, Wilderman PJ, Vasil AI, Vasil ML (2002) GeneChip® expression analysis of the iron starvation response in Pseudomonas aeruginosa: identification of novel pyoverdine biosynthesis genes. Mol Microbiol 45: 1277–1287.

O’Connell LA, Hofmann HA (2011) Genes, hormones, and circuits: an integrative approach to study the evolution of social behavior. Front Neuroendocrinol 32: 320–335.

O’Connor RJ (1978) Brood reduction in birds: selection for fratricide, infanticide and suicide. Anim Behav 26: 79–96.

O’Donnell S (1996) RAPD markers suggest genotypic effects on forager specialization in a eusocial wasp. Behav Ecol Sociobiol 38: 83–88.

O’Donnell S, Jeanne RL (1990) Forager specialization and the control of nest repair in Polybia occidentalis Olvier (Hymenoptera: Vespidae). Behav Ecol Sociobiol 27: 359–364.

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: 507–520.

Ogunniyi AD, Grabowicz M, Briles DE, Cook J, Paton JC (2007) Development of a vaccine against invasive pneumococcal disease based on combinations of virulence proteins of Streptococcus pneumoniae. Infect Immun 75: 350–357.

Ohba SY, Hidaka K, Sasaki M (2006) Notes on paternal care and sibling cannibalism in the giant water bug, Lethocerus deyrolli (Heteroptera: Belostomatidae). Entomol Sci 9: 1–5.

Ohgushi T (1991) Lifetime fitness and evolution of reproductive pattern in the herbivorous lady beetle. Ecology 72: 1110–2122.

Ohtsuki H, Iwasa Y (2004) How should we define goodness?—reputation dynamics in indirect reciprocity. J Theor Biol 231: 107–120.

Ohtsuki H, Iwasa Y (2006) The leading eight: Social norms that can maintain cooperation by indirect reciprocity. J Theor Biol 239: 435–444.

Ohtsuki H, Hauert C, Lieberman E, Nowak MA (2006) A simple rule for the evolution of cooperation on graphs and social networks. Nature 441: 502–505.

Okada Y, Miyazaki S, Miyakawa H, Ishikawa A, Tsuji K, Miura T (2010) Ovarian development and insulin-signaling pathways during reproductive differentiation in the queenless ponerine ant Diacamma sp. J Insect Physiol 56: 288–295.

Okajima R (2008) The controlling factors limiting maximum body size of insects. Lethaia 41: 423–430.

Okasha S (2002) Genetic relatedness and the evolution of altruism. Philos Sci 69: 138–149.

Okasha S (2005) Maynard Smith on the levels of selection question. Biol Philos 20: 989–1010.

Okasha S (2012) Social justice, genomic justice and the veil of ignorance: Harsanyi meets Mendel. Econ Philos 28: 43–71.

Okasha S, Binmore K, eds. (2012) Evolution and rationality: decisions, co-operation, and strategic behaviour. Cambridge, UK: Cambridge University Press.

Oldroyd BP, Smolenski AJ, Cornuet J-M, Crozier RH (1994) Anarchy in the beehive. Nature 371: 749.

Oldroyd BP, Thexton EG, Lawler SH, Crozier RH (1997) Population demography of Australian feral bees (Apis mellifera). Oecologia 111: 381–387.

Oldroyd BP, Fewell JH (2007) Genetic diversity promotes homeostasis in insect colonies. Trends Ecol Evol 22: 408–413.

Oliver ER, Saunders TL, Tarle SA, Glaser T (2004) Ribosomal protein L24 defect in belly spot and tail (Bst), a mouse Minute. Development 131: 3907–3920.

Olivier B (2004) Serotonin and aggression. Ann NY Acad Sci 1036: 382–392.

Olsén KH, Grahn M, Lohm J, Langefors Å (1998) MHC and kin discrimination in juvenile Arctic charr, Salvelinus alpinus (L.). Anim Behav 56: 319–327.

Olsson A, Ochsner KN (2008) The role of social cognition in emotion. Trends Cogn Sci 12: 65–71.

Oltvai ZN, Barabasi AL (2002) Life’s complexity pyramid. Science 298: 763–764.

O’Neil KM (2000) Solitary wasps: behavior and natural history. Ithaca, NY: Comstock.

Oohata AA (1996) Factors controlling prespore cell differentiation in Dictyostelium discoideum: Minute amounts of differentiation–inducing factor promote prespore cell differentiation. Differentiation 59: 283–288.

Oono R, Anderson CG, Denison RF (2011) Failure to fix nitrogen by non-reproductive symbiotic rhizobia triggers host sanctions that reduce fitness of their reproductive clonemates. Proc R Soc B Biol Sci 278: 2698–2703.

Opsata M, Nes IF, Holo H (2010) Class IIa bacteriocin resistance in Enterococcus faecalis V583: the mannose PTS operon mediates global transcriptional responses. BMC Microbiol 10: 224.

O’Riain MJ, Bennett NC, Brotherton PNM, McIlrath G, Clutton-Brock TH (2000) Reproductive suppression and inbreeding avoidance in wild populations of co-operatively breeding meerkats (Suricata suricatta). Behav Ecol Sociobiol 48: 471–477.

O’Riain MJ, Faulkes C (2008) African mole-rats: eusociality, relatedness and ecological constraints. In: Korb J, Heinze J, eds. Ecology of social evolution. Berlin, Germany: Springer. pp 207–223.

Orona-Tamayo D, Heil M (2013) Stabilizing mutualisms threatened by exploiters: new insights from ant-plant research. Biotropica 45: 654–665.

Orr HA (2005) The genetic theory of adaptation: a brief history. Nat Rev Genet 6: 119–127.

Orr HA (2007) Absolute fitness, relative fitness, and utility. Evolution 61: 2997–3000.

Orr HA (2010) The population genetics of beneficial mutations. Phil Trans R Soc B Biol Sci 365: 1195–1201.

Osorno JL, Drummond H (2003) Is obligate siblicidal aggression food sensitive? Behav Ecol Sociobiol 54: 547–554.

Öst M, Ydenberg R, Kilpi R, Lindström K (2003) Condition and coalition formation by brood-rearing common eider females. Behav Ecol 14: 311–317.

Öst M, Vitikainen E, Waldeck P, Sundström L, Lindström K, Hollmén T, et al. (2005) Eider females form nonkin brood-rearing coalitions. Mol Ecol 14: 3903–3908.

Oster GF, Wilson EO (1978) Caste and ecology in the social insects. Princeton, NJ: Princeton University Press.

Ostrom E (1990) Governing the commons: The evolution of instituitions for collective action. Cambridge, UK: Cambridge University Press.

Ostrowski EA, Katoh M, Shaulsky G, Queller DC, Strassmann JE (2008) Kin discrimination increases with genetic distance in a social amoeba. PLoS Biol 6: e287.

Ostrowski EA, Shaulsky G (2009) Learning to get along despite struggling to get by. Genome Biol 10: 218.

O’Toole GA, Kaplan HB, Kolter R (2000) Biofilm formation as microbial development. Annu Rev Microbiol 54: 49–79.

Otto SP (2004) Two steps forward, one step back: the pleiotropic effects of favoured alleles. Proc R Soc Lond B 271: 705–714.

Otto SP, Orive ME (1995) Evolutionary consequences of mutation and selection within an individual. Genetics 141: 1173–1187.

Otto SP, Hastings IM (1998) Mutation and selection within the individual. Genetics 102/103: 507–524.

Otto SP, Lenormand T (2002) Resolving the paradox of sex and recombination. Nat Rev Genet 3: 252–261.

Overath P, Sturm T, Rammensee HG (2014) Of volatiles and peptides: in search for MHC-dependent olfactory signals in social communication. Cell Mol Life Sci 71: 2429–2442.

Overson R, Gadau J, Clark RM, Pratt SC, Fewell JH (2014) Behavioral transitions with the evolution of cooperative nest founding by harvester ant queens. Behav Ecol Sociobiol 68: 21–30.

Owens IPF (2006) Where is behavioural ecology going? Trends Ecol Evol 21: 356–361.

Ozaki M, Wada-Katsumata A, Fujikawa K, Iwasaki M, Yokohari F, Satoji Y, et al. (2005) Ant nestmate and non-nestmate discrimination by a chemosensory sensillum. Science 309: 311–314.

Paar J, Oldroyd BP, Huettinger E, Kastberger G (2002) Drifting of workers in nest aggregations of the giant honeybee Apis dorsata. Apidologie 33: 553–561.

Pabalan N, Davey KG, Packer L (2000) Escalation of aggressive interactions during staged encounters in Halictus ligatus Say (Hymenoptera: Halictidae), with a comparison of circle tube behaviors with other Halictine species'. J Insect Behav 13: 627–650.

Pacala SW, Roughgarden J (1985) Population experiments with the Anolis lizards of St Maarten and St Eustatius. Ecology 66: 129–141.

Pacheco JM, Traulsen A, Nowak MA (2006) Coevolution of strategy and structure in complex networks with dynamical linking. Phys Rev Lett 97: 258103.

Pacheco JM, Pinheiro FL, Santos FC (2009) Population structure induces a symmetry breaking favoring the emergence of cooperation. PLoS Comput Biol 5: e1000596.

Pachepsky E, Taylor T, Jones S (2002) Mutualism promotes diversity and stability in a simple artificial ecosystem. Artif Life 8: 5–24.

Packard JM (2003) Wolf behavior: reproductive, social and intelligent. In: Mech LD, Boitani L, eds. Wolves: behavior, ecology and conservation. Chicago, IL: University of Chicago Press. pp 35–65.

Packer C, Pusey AE (1982) Cooperation and competition within coalitions of male lions: kin selection or game theory? Nature 296: 740–742.

Packer C, Ruttan L (1988) The evolution of cooperative hunting. Am Nat 132: 159–198.

Packer C, Scheel D, Pusey AE (1990) Why lions form groups: food is not enough. Am Nat 136: 1–19.

Packer C, Gilbert DA, Pusey AE, Obrien SJ (1991) A molecular genetic-analysis of kinship and cooperation in African lions. Nature 351: 562–565.

Packer C, Pusey AE, Eberly LE (2001) Egalitarianism in female African lions. Science 293: 690–693.

Packer L, Owen RE (1994) Relatedness and sex ratio in a primitively eusocial halictine bee. Behav Ecol Sociobiol 34: 1–10.

Packer P, Owen RE (1994) Relatedness and sex ratio in a primitively eusocial halictine bee. Behav Ecol Sociobiol 34: 1–10.

Page RE (1980) The evolution of multiple mating behavior by honey bee queens (Apis mellifera L.). Genetics 96: 263–273.

Page RE (1997) The evolution of insect societies. Endeavour 21: 114–120.

Page RE Jr, Metcalf RA (1982) Multiple mating, sperm utilization, and social evolution. Am Nat 119: 263–281.

Page RE, Erickson EH (1988) Reproduction by worker honey bees (Apis mellifera). Behav Ecol Sociobiol 23: 117–126.

Page RE Jr, Robinson GE (1991) The genetics of division of labour in honey bee colonies. Adv Insect Physiol 23: 117–169.

Page RE, Robinson GE (1994) Reproductive competition in queenless honey bee colonies (Apis mellifera L.). Behav Ecol Sociobiol 35: 99–107.

Page RE, Erber J, Fondrk MK (1998) The effect of genotype on response thresholds to sucrose and foraging behavior of honey bees (Apis mellifera L.). J Comp Physiol A 182: 489–500.

Page RE, Mitchell SD (1998) Self-organization and the evolution of division of labor. Apidologie 29: 171–190.

Pais R, Lohs C, Wu Y, Wang J, Aksoy S (2008) The obligate mutualist Wigglesworthia glossinidia influences reproduction, digestion, and immunity processes of its host, the tsetse fly. Appl Environ Microb 74: 5965–5974.

Pál C, Papp B (2000) Selfish cells threaten multicellular life. Trends Ecol Evol 15: 351–352.

Pal C, Maciá MD, Oliver A, Schachar I, Buckling A (2007) Coevolution with viruses drives the evolution of bacterial mutation rates. Nature 450: 1079–1081.

Palková Z, Váchová L (2006) Life within a community: benefit to yeast long-term survival. FEMS Microbiol Rev 30: 806–824.

Palmer KA, Oldroyd BP (2000) Evolution of multiple mating in the genus Apis. Apidologie 31: 235–248.

Palmer TM, Stanton ML, Young TP, Goheen JR, Pringle RM, Karban R (2008) Breakdown of an ant-plant mutualism follows the loss of large herbivores from an African savanna. Science 319: 192–195.

Palmer TM, Doak DF, Stanton ML, Bronstein JL, Kiers ET, Young TP, et al. (2010) Synergy of multiple partners, including freeloaders, increases host fitness in a multispecies mutualism. Proc Natl Acad Sci USA 107: 17234–17239.

Pamilo P (1991a) Evolution of colony characteristics in social insects. I. Sex allocation. Am Nat 137: 83–107.

Pamilo P (1991b) Evolution of colony characteristics in social insects. II. Number of reproductive individuals. Am Nat 138: 412–433.

Pamilo P, Rosengren R (1984) Evolution of nesting strategies of ants: genetic evidence from different population types of Formica ants. Biol J Linn Soc 21: 331–348.

Panchanathan K, Boyd R (2004) Indirect reciprocity can stabilize cooperation without the second-order free rider problem. Nature 432: 499–502.

Pankiw T, Huang Z, Winston ML, Robinson GE (1998) Queen mandibular gland pheromone influences worker honey bee (Apis mellifera L.) foraging ontogeny and juvenile hormone titers. J Insect Physiol 44: 685–692.

Pantopoulos K, Hentze MW (1995) Rapid responses to oxidative stress mediated by iron regulatory protein. EMBO J 14: 2917–2924.

Parfrey LW, Grant J, Tekle YI, Lasek-Nesselquist E, Morrison HG, Sogin ML, Patterson DJ, Katz LA (2010) Broadly sampled multigene analyses yield a well-resolved eukaryotic tree of life. Syst Biol 59: 518–533.

Park T, Mertz DB, Petrusewicz K (1961) Genetic strains of Tribolium: their primary characteristics. Physiol Zool 34: 62–80.

Park T, Mertz DB, Grodzinski W, Prus T (1965) Cannibalistic predation in populations of flour beetles. Physiol Zool 38: 289–321.

Parker GA, Mock DW, Lamey TC (1989) How selfish should stronger sibs be? Am Nat 133: 846–868.

Parker GA, Partridge L (1998) Sexual conflict and speciation. Phil Trans R Soc Lond Ser B 353: 261–274.

Parker KJ, Lee TM (2003) Female meadow voles (Microtus pennsylvanicus) demonstrate same-sex partner preferences. J Comp Psychol 117: 283–289.

Parkinson K, Buttery NJ, Wolf JB, Thompson CRL (2011) A simple mechanism for complex social behavior. PLoS Biol 9: e1001039.

Parmigiani S, vom Saal FS, eds. (1994) Infanticide and parental care. New York, NY: Harwood Academic.

Parrish JK, Viscido SV, Grünbaum D (2002) Self-organized fish schools: An examination of emergent properties. Biol Bull 202: 296–305.

Partecke J, Schwabl H (2008) Organizational effects of maternal testosterone on reproductive behavior of adult house sparrows. Dev Neurobiol 68: 1538–1548.

Partridge BL (1981) Internal dynamics and the interrelations of fish in schools. J Comp Physiol A 144: 313–325.

Partridge BL (1982) The structure and function of fish schools. Sci Am 246: 90–99.

Partridge L, French V (1996) Thermal evolution of ectotherm body size: why get big in the cold? In: Johnston IA, Bennett AF, eds. Animals and temperature: Phenotypic and evolutionary adaptation. Cambridge, UK: Cambridge University Press. pp 265–292.

Parvinen K, Dieckmann U, Gyllenberg M, Metz JAJ (2003) Evolution of dispersal in metapopulations with local density dependence and demographic stochasticity. J Evol Biol 16: 143–153.

Pascale RT, Millemann M, Gioja L (2000) Surfing the edge of chaos: the laws of nature and the new laws of business. New York, NY: Random House Inc.

Pasquet A, Krafft B (1992) Cooperation and prey capture efficiency in a social spider Anelosimus eximius (Araneae, Theridiidae). Ethology 90: 121–133.

Pasquet A, Trabalon M, Bagnères AG, Leborgne R (1997) Does group closure exist in the social spider Anelosimus eximius? Behavioral and chemical approach. Insect Soc 44: 159–169.

Passarge J, Hol S, Escher M, Huisman J (2006) Competition for nutrients and light: stable coexistence, alternative stable states, or competitive exclusion? Ecol Monogr 76: 57–72.

Passera L, Aron S, Vargo EL, Keller L (2001) Queen control of sex ratio in fire ants. Science 293: 1308–1310.

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.

Paster BJ, Olsen I, Aas JA, Dewhirst FE (2006) The breadth of bacterial diversity in the human periodontal pocket and other oral sites. Periodontol 2000 42: 80–87.

Patel A, Fondrk MK, Kaftanoglu O, Emore C, Hunt G, et al. (2007) The making of a queen: TOR pathway is a key player in diphenic caste development. PLoS ONE 2: e509.

Paterson S, Vogwill T, Buckling A, Benmayor R, Spiers AJ, et al. (2010) Antagonistic coevolution accelerates molecular evolution. Nature 464: 275–278.

Paton JC (1996) The contribution of pneumolysin to the pathogenicity of Streptococcus pneumoniae. Trends Microbiol 4: 103–106.

Patton J (2005) Meat sharing for coalitional support. Evol Hum Behav 26: 137−157.

Pavelka MSM, Fedigan LM, Zohar S (2002) Availability and adaptive value of reproductive and postreproductive Japanese macaque mothers and grandmothers. Anim Behav 64: 407–414.

Paxton RJ, Thorén PA, Tengö J, Estoup A, Pamilo P (1997) Mating structure and nestmate relatedness in a communal bee, Andrena jacobi (Hymenoptera, Andrenidae), using microsatellites. Mol Ecol 5: 511–519.

Paxton RJ, Kukuk PF, Tengö J (1999) Effects of familiarity and nestmate number on social interactions in two communal bees, Andrena scotica and Panurgus calcaratus (Hymenoptera, Andrenidae). Insect Soc 46: 109–118.

Pearce AN, Huang ZY, Breed MD (2001) Juvenile hormone and aggression in honey bees. J Insect Physiol 47: 1243–1247.

Pearcy M, Aron S, Doums C, Keller L (2004) Conditional use of sex and parthenogenesis for worker and queen production in ants. Science 306: 1780–1783.

Pearson J, Jakobsen I (1993) The relative contribution of hyphae and roots to phosphorus uptake by arbuscular mycorrhizal plants, measured by dual labelling with 32P and 33P. New Phytol 124: 489–494.

Pedersen CA (2004) Biological aspects of social bonding and the roots of human violence. Ann NY Acad Sci 1036: 106–127.

Pedersen CA, Prange AJ (1979) Induction of maternal behavior in virgin rats after intracerebroventricular administration of oxytocin. Proc Natl Acad Sci USA 76: 6661–6665.

Pedersen CA, Ascher JA, Monroe YL, Prange AJ (1982) Oxytocin induces maternal behavior in virgin female rats. Science 216: 648–650.

Pedersen CA, Caldwell JD, Johnson MF, Fort SA, Prange Jr AJ (1985) Oxytocin antiserum delays onset of ovarian steroid-induced maternal behavior. Neuropeptides 6: 175–182.

Pedersen CA, Boccia ML (2003) Oxytocin antagonism alters rat dams oral grooming and upright posturing over pups. Physiol Behav 80: 233–241.

Peeters C (1993) Monogyny and polygyny in ponerine ants with or without queens. In: Keller L, ed. Queen number and sociality in insects. Oxford, UK: Oxford University Press. pp 234–261.

Peeters C (1997) Morphologically ‘primitive’ ants: comparative review of social characters, and the importance of queen-worker dimorphism. In: Choe JC, Crespi BJ, eds. The evolution of social behavior in insects and arachnids. Cambridge, UK: Cambridge University Press. pp 372–391.

Peeters C, Monnin T, Malosse C (1999) Cuticular hydrocarbons correlated with reproductive status in a queenless ant. Proc R Soc Lond B 266: 1323–1327.

Peeters C, Liebig J, Hölldobler B (2002) Polygyny, hierarchy and lack of reproductive skew among gamergates in the ant Harpegnathos saltator. Insect Soc 47: 325–332.

Peeters C, Molet M (2010) Colonial reproduction and life histories. In: Lach L, Parr CL, Abbott KL, eds. Ant ecology. Oxford, UK: Oxford University Press. pp 159–176.

Peeters C, Liebig J (2009) Fertility signaling as a general mechanism of regulating reproductive division of labor in ants. In: Gadau J, Fewell J, eds. Organization of insect societies: from genome to sociocomplexity. Cambridge, MA: Harvard University Press. pp 220–242.

Pellmyr O, Huth CJ (1994) Evolutionary stability of mutualism between yuccas and yucca moths. Nature 372: 257–260.

Pellmyr O, Leebens-Mack J (1999) Forty million years of mutualism: evidence for Eocene origin of the yucca–yucca moth association. Proc Natl Acad Sci USA 96: 9178–9183.

Pellmyr O, Leebens-Mack J, Huth CJ (1996) Nonmutualistic yucca moths and their evolutionary consequences. Nature 380: 155–156.

Pen I, Weissing FJ (2000) Towards a unified theory of cooperative breeding: the role of ecology and life history re-examined. Proc R Soc Lond B Biol Sci 267: 2411–2418.

Pen I, Weissing FJ (2001) Optimal floating and queuing strategies: the logic of territory choice. Am Nat 155: 512–526.

Penn DJ (2002) The scent of genetic compatibility: sexual selection and the major histocompatibility complex. Ethology 108: 1–21.

Penn D, Potts W (1998) How do major histocompatibility complex genes influence odor and mating preferences? Adv Immunol 69: 411–436.

Penn DJ, Frommen JG (2010) Kin recognition: an overview of conceptual issues, mechanisms and evolutionary theory. In: Kappeler PM, ed. Animal behaviour: evolution and mechanisms. Heidelberg, Germany: Springer. pp 55–85.

Pentz JT, Travisano M, Ratcliff WC (2014) Clonal development is evolutionarily superior to aggregation in wild-collected Saccharomyces cerevisiae. In: Sayama H, Rieffel J, Risi S, Doursat R, Lipson H, eds. Artificial Life 14: Proceedings of the Fourteenth International Conference on the Synthesis and Simulation of Living Systems. Cambridge, MA: MIT Press. pp 550-554. DOI: http://dx.doi.org/10.7551/978-0-262-32621-6-ch088

Pepper JW, Smuts B (2002) A mechanism for the evolution of altruism among nonkin: positive assortment through environmental feedback. Am Nat 160: 205–213.

Perc M (2006a) Coherence resonance in a spatial prisoner's dilemma game. New J Phys 8: 22.

Perc M (2006b) Chaos promotes cooperation in the spatial prisoner's dilemma game. Europhys Lett 75: 841–846.

Perc M (2007a) Transition from Gaussian to levy distributions of stochastic payoff variations in the spatial prisoner’s dilemma game. Phys Rev E 75: 022101.

Perc M (2007b) Does strong heterogeneity promote cooperation by group interactions? New J Phys 13: 123027.

Perc M, Marhl M (2006) Evolutionary and dynamical coherence resonances in the pair approximated prisoner's dilemma game. New J Phys 8: 142.

Perc M, Szolnoki A (2008) Social diversity and promotion of cooperation in the spatial prisoner’s dilemma game. Phys Rev E77: 011904.

Perc M, Wang Z (2010) Heterogeneous aspirations promote cooperation in the Prisoner’s Dilemma game. PLoS ONE 5: e15117.

Pérez-Brocal V, Latorre A, Moya A (2013) Symbionts and pathogens: what is the difference? In: Dobrindt U, Hacker JH, Svanborg C, eds. Between pathogenicity and commensalism. Berlin, Germany: Springer. pp 215–243.

Pérez-Tomé J, Toro M (1982) Competition of similar and non-similar genotypes. Nature 299: 153–154.

Pernal SF, Currie RW (2000) Pollen quality of fresh and 1-year-old single pollen diets for worker honey bees (Apis mellifera L.). Apidologie 31: 387–409.

Perreault C, Moya C, Boyd R (2012) A Bayesian approach to the evolution of social learning. Evol Hum Behav 33: 449–459.

Perrin N, Goudet J (2001) Inbreeding, kinship and the evolution of natal dispersal. In: Clobert J, Danchin E, Dhondt AA, Nichols JD, eds. Dispersal. Oxford, UK: Oxford University Press. pp 123–142.

Perry JC, Roitberg BD (2005) Games among cannibals: competition to cannibalize and parent-offspring conflict lead to increased sibling cannibalism. J Evol Biol 18: 1523–1533.

Perry SE, Manson JH (2003) Traditions in monkeys. Evol Anthropol 12: 71–81.

Peschel A, Götz F (1996) Analysis of the Staphylococcus epidermidis genes epiF, -E, and -G involved in epidermin immunity. J Bacteriol 178: 531–536.

Peters RH (1983) The ecological implications of body size. Cambridge, UK: Cambridge University Press.

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: 171–174.

Petrovic P, Kalisch R, Singer T, Dolan RJ (2008) Oxytocin attenuates affective evaluations of conditioned faces and amygdala activity. J Neurosci 28: 6607–6615.

Pettis JS, Higo HA, Pankiw T, Winston ML (1997) Queen rearing suppression in the honey bee - evidence for a fecundity signal. Insect Soc 44: 311–322.

Pfaller MA, Mujeeb I, Hollis RJ, Jones RN, Doern GV (2000) Evaluation of the discriminatory powers of the Dienes test and ribotyping as typing methods for Proteus mirabilis. J Clin Microbiol 38: 1077–1080.

Pfeiffer KJ, Crailsheim K (1998) Drifting of honeybees. Insect Soc 45: 151–167.

Pfeiffer T, Schuster S, Bonhoeffer S (2001) Cooperation and competition in the evolution of ATP-producing pathways. Science 292: 504–507.

Pfennig DW (1995) The absence of joint nesting advantage in desert seed harvester ants: evidence from a field experiment. Anim Behav 49: 567–575.

Pfennig DW (1997) Kinship and cannibalism. BioScience 47: 667–775.

Pfennig DW, Sherman PW (1995) Kin recognition. Sci Am 272: 98–103.

Pfennig DW, Collins JP, Ziemba RE (1999) A test of alternative hypotheses for kin recognition in cannibalistic tiger salamanders. Behav Ecol 10: 436–443.

Pfennig DW, Rice AM, Martin RA (2007) Field and experimental evidence for competition's role in phenotypic divergence. Evolution 61: 257–271.

Pfennig KS, Pfennig DW (2005) Character displacement as the “best of a bad situation”: fitness trade-offs resulting from selection to minimize resource and mate competition. Evolution 59: 2200–2208.

Philippi T, Seger J (1989) Hedging one’s evolutionary bets, revisited. Trends Ecol Evol 4: 41–44.

Pianka ER (1970) On r- and K-selection. Am Nat 104: 592–597.

Pianka ER(1974a) Niche overlap and diffuse competition.Proc Natl Acad Sci USA 71: 2141–2145.

Pianka ER (1974b) Evolutionary ecology. New York, NY: Harper and Row.

Picker MD, Hoffman MT, Leverton B (2007) Density of Microhodotermes viator (Hodotermitidae) mounds in southern Africa in relation to rainfall and vegetative productivity gradients. J Zool 271: 37–44.

Piertney SB, Oliver MK (2006) The evolutionary ecology of the major histocompatibility complex. Heredity 96: 7–21.

Pigliucci M (2001) Phenotypic plasticity: beyond nature and nurture. Baltimore, MD: Johns Hopkins University Press.

Pigliucci M, Murren CJ, Schlichting CD (2006) Phenotypic plasticity and evolution by genetic assimilation. J Exp Biol 209: 2362–2367.

Pijpe J (2007) The evolution of lifespan in the butterfly Bicyclus anynana. Enschede, the Netherlands: Print Partners Ipskamp.

Pike N, Foster W (2004) Fortress repair in the social aphid species Pemphigus spyrothecae. Anim Behav 67: 909–914.

Pike N, Foster WA (2008) The ecology of altruism in a clonal insect. In: Korb J, Heinze J, eds. Ecology of social evolution. Berlin, Germany: Springer-Verlag. pp 37–56.

Piotrowski A, Bruder CE, Andersson R, de Ståhl TD, Menzel U, Sandgren J, et al. (2008) Somatic mosaicism for copy number variation in differentiated human tissues. Hum Mutat 29: 1118–1124.

Piper MDW, Selman C, McElwee JJ, Partridge L (2008) Separating cause from effect: how does insulin/IGF signalling control lifespan in worms, flies and mice? J Intern Med 263: 179–191.

Piper WH (1994) Courtship, copulation, nesting behavior and brood parasitism in the Venezuelan stripe-backed wren. Condor 96: 654–671.

Pirk CWW, Neumann P, Ratnieks FLW (2003) Cape honeybees, Apis mellifera capensis, police worker-laid eggs despite the absence of relatedness benefits. Behav Ecol 14: 347–352.

Pirk CWW, Neumann P, Hepburn HR, Moritz RFA, Tautz J (2004) Egg viability and worker policing in honey bees. Proc Natl Acad Sci USA 101: 8649–8651.

Pitcher TJ, Parrish JK (1993) Functions of shoaling behavior in teleosts. In: Pitcher TJ, ed. The behavior of teleost fishes. New York, NY: Chapman & Hall. pp 363–439.

Pitelka LF, Ashmun JW (1985) Physiology and integration of ramets in clonal plants. In: Jackson JBC, Buss LW, Cook RE, eds. Population biology and evolution of clonal organisms. New Haven, CT: Yale University Press. pp 399–435.

Platt TG, Bever JD (2009) Kin competition and the evolution of cooperation. Trends Ecol Evol 24: 370–377.

Platt TG, Fuqua C, Bever JD (2012) Resource and competitive dynamics shape the benefits of public goods cooperation in a plant pathogen. Evolution 66: 1953–1965.

Plettner E, Slessor KN, Winston ML, Robinson GE, Page RE (1993) Mandibular gland components and ovarian development as measures of caste differentiation in the honey bee (Apis mellifera L). J Insect Physiol 39: 235–240.

Poethke HJ, Liebig J (2008) Risk-sensitive foraging and the evolution of cooperative breeding and reproductive skew. BMC Ecol 8: 2.

Poiani A, Jermiin LS (1994) A comparative analysis of some life history traits between cooperatively and non-cooperatively breeding Australian passerines. Evol Ecol 8: 1–8.

Polis GA (1981) The evolution and dynamics of intraspecific predation. Annu Rev Ecol Syst 12: 225–251.

Polizzi JM, Forschler BT (1999) Factors that affect aggression among the worker caste of Reticulitermes spp. subterranean termites (Isoptera: Rhinotermitidae). J Insect Behav 12: 133–146.

Pollard KA, Blumstein DT (2008) Time allocation and the evolution of group size. Anim Behav 76: 1683–1699.

Pollock GB, Rissing SW (1985) Mating season and colony foundation of the seed-harvester ant, Veromessor pergandei. Psyche 92: 125–134.

Pomerening JR (2008) Uncovering mechanisms of bistability in biological systems. Curr Opin Biotechnol 19: 381–388.

Pomeroy PP, Twiss SD, Redman P (2000) Philopatry, site fidelity and local kin associations within grey seal breeding colonies. Ethology 106: 899–919.

Pomeroy PP, Wilmar JW, Amos W, Twiss SD (2001) Reproductive performance links to fine-scale spatial patterns of female grey seal relatedness. Proc R Soc Lond Ser B Biol Sci 268: 711–717.

Poncela J, Gómez-Gardeñes J, Floría LM, Moreno Y (2007) Robustness of cooperation in the evolutionary prisoner’s dilemma on complex systems. New J Phys 9: 184.

Pope TR (2000) Reproductive success increases with degree of kinship in cooperative coalitions of female red howler monkeys (Alouatta seniculus). Behav Ecol Sociobiol 48: 253–267.

Popova NK (2006) From genes to aggressive behavior: the role of serotonergic system. BioEssays 28: 495–503.

Porter RH (1998) Olfaction and human kin recognition. Genetica 104: 259–263.

Porter RH, Wyrick M, Pankey J (1978) Sibling recognition in spiny mice (Aeomys cahirinus). Behav Ecol Sociobiol 3: 61–68.

Porter SS, Simms EL (2014) Selection for cheating across disparate environments in the legume-rhizobium mutualism. Ecol Lett DOI: 10.1111/ele.12318.

Portugal SJ, Hubel TY, Fritz J, Heese S, Trobe D, Voelkl B, et al. (2014) Upwash exploitation and downwash avoidance by flap phasing in ibis formation flight. Nature 505: 399–404.

Post JR, Evans DO (1989) Size-dependent overwinter mortality of young-of-the-year yellow perch (Perca flavescens): laboratory, in situ enclosure, and field experiments. Can J Fish Aquat Sci 46: 1958–1968.

Poston JP (1997) Dominance, access to colonies, and queues for mating opportunities by male boat-tailed grackles. Behav Ecol Sociobiol 41: 89–98.

Potts R (2013) Hominin evolution in settings of strong environmental variability. Quaternary Sci Rev 73: 1–13.

Poulin R, Grutter AS (1996) Cleaning symbioses: proximate and adaptive explanations. BioScience 46: 512–517.

Powell GVN (1974) Experimental analysis of social value of flocking by starlings (Sturnus vulgaris) in relation to predation and foraging. Anim Behav 22: 501–505.

Powell RA, Zimmerman JW, Seaman DE (1997) Ecology and behaviour of North American black bears: home ranges, habitat and social organization. London, UK: Chapman & Hall.

Powers ST, Penn AS, Watson RA (2011) The concurrent evolution of cooperation and the population structures that support it. Evolution 65: 1527–1543.

Pöysä H (1999) Conspecific nest parasitism is associated with inequality in nest predation risk in the common goldeneye (Bucephala clangula). Behav Ecol 10: 533–540.

Pöysä H (2003) Parasitic common goldeneye (Bucephala clangula) females lay preferentially in safe neighbourhoods. Behav Ecol Sociobiol 54: 30–35.

Pöysä H (2006) Public information and conspecific nest parasitism in goldeneyes: targeting safe nests by parasites. Behav Ecol 17: 459–465.

Pöysä H, Ruusila V, Milonoff M, Virtanen J (2001) Ability to assess nest predation risk in secondary hole-nesting birds: an experimental study. Oecologia (Berlin) 126: 201–207.

Pöysä H, Pesonen M (2007) Nest predation and the evolution of conspecific brood parasitism: from risk spreading to risk assessment. Am Nat 169: 94–104.

Pöysä H, Eadie JM, Lyon BE (2014) Conspecific brood parasitism in waterfowl and cues parasites use. Wildfowl Special Issue 4: 192–219.

Pratt SC, Mallon EB, Sumpter DJT, Franks NR (2002) Quorum sensing, recruitment, and collective decision-making during colony emigration by the ant Leptothorax albipennis. Behav Ecol Sociobiol 52: 117–127.

Pratte M (1989) Foundress association in the paper wasp Polistes dominulus Christ. (Hymen., Vesp.). Effects of dominance hierarchy on the division of labour. Behaviour 111: 208–219.

Premnath S, Sinha A, Gadagkar R (1996) Dominance relationship in the establishment of reproductive division of labour in a primitively eusocial wasp (Ropalidia marginata). Behav Ecol Sociobiol 39: 125–132.

Preston SD, de Waal FBM (2002) Empathy: its ultimate and proximate bases. Behav Brain Sci 25: 1–72.

Preston T (1981) The clay pedestal. Seattle, WA: Madrona Publishers.

Prete FR (1990) The conundrum of the honey bees: One impediment to the publication of Darwin‘s theory. J Hist Biol 23: 271–290.

Prezoto F, Santos-Prezoto HH (2005) Occurrence of nest fusion in the swarm-founding wasp Polybia paulista (Hymenoptera: Vespidae). Sociobiology 45: 99–103.

Price M (2007) The controversy of group selection theory. SCQ 1.

Price MV, Waser NM (1982) Population structure, frequency-dependent selection, and the maintenance of sexual reproduction. Evolution 36: 35–43.

Price PW (1991) The web of life: development over 3.8 billion years of trophic relationships. In: Margulis L, Fester R, eds. Symbiosis as a source of evolutionary innovation. Cambridge, MA: MIT Press. pp 262–272.

Price PW, Fernandes GW, Waring GL (1987) Adaptive nature of insect galls. Environ Entomol 16: 15–24.

Pride RE (2005) Optimal group size and seasonal stress in ring-tailed lemurs (Lemur catta). Behav Ecol 16: 550–560.

Priess JR, Thomson JN (1987) Cellular interactions in early C. elegans embryos. Cell 48: 241–250.

Prindle A, Hasty J (2010) Stochastic emergence of groupthink. Science 328: 987.

Pringle A, Bever JD (2008) Analogous effects of arbuscular mycorrhizal fungi in the laboratory and a North Carolina field. New Phytol 180: 162–175.

Prins HHT (1996) Ecology and behaviour of the African buffalo: social inequality and decision making. London, UK: Chapman & Hall.

Proctor CJ, Broom M, Ruxton GD (2001) Modelling antipredator vigilance and flight response in group foragers when warning signals are ambiguous. J Theor Biol 211: 409–417.

Proulx R (2007) Ecological complexity for unifying ecological theory across scales: A field ecologist’s perspective. Ecol Complex 4: 85–92.

Provost E (1987) Role of the queen in the intra-colonial aggressivity and nestmate recognition in Leptothorax lichtensteini ants. In: Eder J, Rembold H, eds. Chemistry and biology of social insects (Proc. 10th Int. Congr. IUSSI, Munich, 1986). Munich, Germany: Verlag J. Peperny. pp 479.

Pruett-Jones SG, Lewis MJ (1990) Sex ratio and habitat limitation promote delayed dispersal in superb fairy wrens. Nature 348: 541–542.

Pruitt JN, Riechert SE (2009) Frequency-dependent success of cheaters during foraging bouts might limit their spread within colonies of a socially polymorphic spider. Evolution 63: 2966–2973.

Pruitt JN, Riechert SE (2011) How within-group behavioural variation and task efficiency enhance fitness in a social group. Proc R Soc B 278: 1209–1215.

Pruitt JN, Oufiero CE, Avilés L, Riechert SE (2012) Iterative evolution of increased behavioral variation characterizes the transition to sociality in spiders and proves advantageous. Am Nat 180: 496–510.

Puga-Gonzalez I, Hildenbrandt H, Hemelrijk CK (2009) Emergent patterns of social affiliation in primates, a model. PLoS Comput Biol 5: e1000630.

Pulliam HR (1973) On the advantages of flocking. J Theor Biol 38: 419–422.

Pulliam HR, Caraco T (1984) Living in groups: Is there an optimal group size? In: Krebs JR, Davies NB, eds. Behavioural ecology: an evolutionary approach. Oxford, UK: Blackwell Scientific. pp 122–147.

Purcell J (2011) Geographic patterns in the distribution of social systems in terrestrial arthropods. Biol Rev Camb Philos Soc 86: 475–491.

Purves D (1980) Neuronal competition. Nature 287: 585–586.

Pusenius J, Viitala J, Marienberg T, Ritvanen S (1998) Matrilineal kin clusters and their effect on reproductive success in the field vole Microtus agrestis. Behav Ecol 9: 85–92.

Pusey AE, Packer C (1987a) The evolution of sex-biased dispersal in lions. Behaviour 101: 275–310.

Pusey AE, Packer C (1987b) Dispersal and philopatry. In: Smuts BB, Cheney DL, Seyfarth RM, Wrangham RW, Struhsaker TT, eds. Primate societies. Chicago, IL: University of Chicago Press. pp 250–266.

Pusey AE, Packer C (1997) The ecology of relationships. In: Krebs JR, Davies NB, eds. Behavioural ecology. An Evolutionary Approach, 4th edn. Oxford, UK: Blackwell Science. pp 254–283.

Putnam RD (2000) Bowling alone: the collapse and revival of American community. New York, NY: Simon & Schuster.

Puurtinen M, Mappes T (2009) Between-group competition and human cooperation. Proc R Soc B Biol Sci 276: 355–360.

Queller DC (1985) Kinship, reciprocity and synergism in the evolution of social behaviour. Nature 318: 366–367.

Queller DC (1989) The evolution of eusociality: reproductive head starts of workers. Proc Natl Acad Sci USA 86: 3224–3226.

Queller DC (1992a) Does population viscosity promote kin selection? Trends Ecol Evol 7: 322–324.

Queller DC (1992b) Quantitative genetics, inclusive fitness, and group selection. Am Nat 139: 540–558.

Queller DC (1992c) A general model for kin selection. Evolution 46: 376–380.

Queller DC (1994a) Genetic relatedness in viscous populations. Evol Ecol 8: 70–73.

Queller DC (1994b) Extended parental care and the origin of eusociality. Proc R Soc Lond Ser B Biol Sci 256: 105–111.

Queller DC (1996) The origin and maintenance of eusociality: the advantage of extended parental care. In: Turillazzi S, West-Eberhard MJ, eds. Natural history and evolution of paper wasps. Oxford, UK: Oxford University Press. pp 218–234.

Queller DC (2000) Relatedness and the fraternal major transitions. Phil Trans R Soc Lond B 355: 1647–1655.

Queller DC (2004) Social evolution—Kinship is relative. Nature 430: 975–976.

Queller DC, Hughes CR, Strassmann JE (1990) Wasps fail to make distinctions. Nature 344: 388.

Queller DC, Strassmann JE (1998) Kin selection and social insects. BioScience 48: 165–175.

Queller DC, Zacchi F, Cervo R, Turillazzi S, Henshaw MT, Santorelli LA, Strassmann JE (2000) Unrelated helpers in a social insect. Nature 405: 784–787.

Queller DC, Strassmann JE (2002) The many selves of social insects. Science 296: 311–313.

Queller DC, Strassmann JE (2003) Eusociality. Curr Biol 13: R861–R863.

Queller DC, Ponte E, Bozzaro S, Strassmann JE (2003) Single-gene greenbeard effects in the social amoeba Dictyostelium discoideum. Science 299: 105–106.

Quinlan AR, Hall IM (2012) Characterizing complex structural variation in germline and somatic genomes. Trends Genet 28: 43–53.

Quinn TP, Busack CA (1985) Chemosensory recognition of siblings in juvenile coho salmon (Oncorhynehus kisuteh). Anim Behav 33: 51–56.

Rabenold KN (1984) Cooperative enhancement of reproductive success in tropical wren societies. Ecology 65: 871–885.

Rabenold KN (1990) Campylorhynchus wrens: the ecology of delayed dispersal and cooperation in the Venezuelan savanna. In: Stacey PB, Koenig WD, eds. Cooperative breeding in birds. Camridge, UK: Cambridge University Press. pp 157–196.

Rabenold PP, Rabenold KN, Piper WH, Haydock J, Zack SW (1990) Shared paternity revealed by genetic analysis in cooperatively breeding tropical wrens. Nature 348: 538–540.

Råberg L, de Roode JC, Bell AS, Stamou P, Gray D, Read AF (2006) The role of immune-mediated apparent competition in genetically diverse malaria infections. Am Nat 168: 41–53.

Rachinsky A (1994) Octopamine and serotonin influence on corpora allata activity in honey bee (Apis mellifera) larvae. J Insect Physiol 40: 549–554.

Rachinsky A, Strambi C, Strambi A, Hartfelder K (1990) Caste and metamorphosis: hemolymph titers of juvenile hormone and ecdysteroids in last instar honeybee larvae. Gen Comp Endocrinol 79: 31–38.

Rachinsky A, Hartfelder K (1990) Corpora allata activity, a prime regulating element for caste-specific juvenile hormone titre in honey bee larvae (Apis mellifera carnica). J Insect Physiol 36: 189–194.

Radke S, De Bruijn ERA (2012) The other side of the coin: oxytocin decreases the adherence to fairness norms. Front Hum Neurosci 6: 193.

Raff RA, Raff EC (1970) Respiratory mechanisms and the metazoan fossil record. Nature 228: 1003–1005.

Raffa KF, Berryman AA (1983) The role of host plant resistance in the colonization behavior and ecology of bark beetles (Coleoptera, Scolytidae). Ecol Monog 53: 27–49.

Raffa KF, Berryman AA (1987) Interacting selective pressures in conifer-bark beetle systems: a basis for reciprocal adaptations? Am Nat 129: 234–262.

Ràfols I, Amagai A, Maeda Y, MacWilliams HK, Sawada Y (2001) Cell type proportioning in Dictyostelium slugs: lack of regulation within a 2.5-fold tolerance range. Differentiation 67: 107–116.

Ragsdale JE (1999) Reproductive skew extended: the effect of resource inheritance on social organization. Evol Ecol Res 1: 859–874.

Raichle ME (2006) The brain’s dark energy. Science 314: 1249–1250.

Raihani N, Thornton A, Bshary R (2012) Punishment and cooperation in nature. Trends Ecol Evol 27: 288–295.

Rainey HJ, Zuberbühler K, Slater PJB (2004) Hornbills can distinguish between primate alarm calls. Proc R Soc Lond B 271: 755–759.

Rainey PB, Travisano M (1998) Adaptive radiation in a heterogeneous environment. Nature 394: 69–72.

Rainey PB, Rainey K (2003) Evolution of cooperation and conflict in experimental bacterial populations. Nature 425: 72–74.

Raj A, van Oudenaarden A (2008) Nature, nurture, or chance: stochastic gene expression and its consequences. Cell 135: 216–226.

Raleigh MJ, Yuwiler A, Brammer GL, McGuire MT, Geller E, Flannery JW (1981) Peripheral correlates of serotonergically-influenced behaviors in vervet monkeys (Cercopithecus aethiops sabaeus). Psychopharmacology 72: 241–246.

Raleigh MJ, Brammer GL, McGuire MT, Yuwiler A (1985) Dominant social status facilitates the behavioral effects of serotonergic agonists. Brain Res 348: 274–282.

Ram RJ, VerBerkmoes NC, Thelen MP, Tyson GW, Baker BJ, Blake RC, et al. (2005) Community proteomics of a natural microbial biofilm. Science 308: 1915–1920.

Ramos D, Victor T, Seidl-de-Moura ML, Daly M (2013) Future discounting by slum-dwelling youth versus university students in Rio de Janeiro. J Res Adolescence 23: 95–102.

Rand DG, Nowak MA (2013) Human cooperation. Trends Cogn Sci 17: 413–425.

Range F, Horn L, Viranyi Z, Huber L (2009) The absence of reward induces inequity aversion in dogs. Proc Natl Acad Sci USA 106: 340–345.

Rangel J, Mattila HR, Seeley TD (2009) No intracolonial nepotism during colony fissioning in honey bees. Proc R Soc B Biol Sci 276: 3895–3900.

Rand DG, Arbesman S, Christakis NA (2011) Dynamic social networks promote cooperation in experiments with humans. Proc Natl Acad Sci USA 108: 19193–19198.

Rand DM, Haney RA, Fry AJ (2004) Cytonuclear coevolution: the genomics of cooperation. Trends Ecol Evol 19: 645–653.

Rangel J, Mattila HR, Seeley TD (2009) No intracolonial nepotism during colony fissioning in honey bees. Proc R Soc B Biol Sci 276: 3895–3900.

Rankin DJ, Bargum K, Kokko H (2007) The tragedy of the commons in evolutionary biology. Trends Ecol Evol 22: 643–651.

Rankin SM, Chambers J, Edwards JP (1997) Juvenile hormone in earwigs: roles in oogenesis, mating, and maternal behaviors. Arch Insect Biochem Physiol 35: 427–442.

Rannala BH, Brown CR (1994) Relatedness and conflict over optimal group size. Trends Ecol Evol 9: 117–119.

Rasa OAE (1994) Altruistic infant care or infanticide: the dwarf mongooses’ dilemma. In: Parmigiani S, vom Saal FS, eds. Infanticide and parental care. London, UK: Harwood Academic Publishers. pp 301–320.

Rasmussen HN, Rasmussen FN (2009) Orchid mycorrhiza: Implications of a mycophagous life style. Oikos 118: 334–345.

Ratcliff WC (2010) Resource hoarding facilitates cheating in the legume-rhizobia symbiosis and bet-hedging in the soil. PhD thesis. Minneapolis, MN: University of Minnesota.

Ratcliff WC, Kadam SV, Denison RF (2008) Polyhydroxybutyrate supports survival and reproduction in starving rhizobia. FEMS Microbiology Ecology 65: 391–399.

Ratcliff WC, Denison RF (2009) Rhizobitoxine producers gain more poly-3-hydroxybutyrate in symbiosis than do competing rhizobia, but reduce plant growth. ISME J 3: 870–872.

Ratcliff WC, Hawthorne P, Travisano M, Denison RF (2009) When stress predicts a shrinking gene pool, trading early reproduction for longevity can increase fitness, even with lower fecundity. PLoS ONE 4: e6055.

Ratikainen II (2012) Foraging in a variable world: Adaptations to stochasticity. PhD thesis. Trondheim, Norway: Norwegian University of Science and Technology.

Ratledge C, Dover LG (2000) Iron metabolism in pathogenic bacteria. Annu Rev Microbiol 54: 881–941.

Ratnieks FLW (1988) Reproductive harmony via mutual policing by workers in eusocial Hymenoptera. Am Nat 132: 217–236.

Ratnieks FLW (1993) Egg-laying, egg-removal, and ovary development by workers in queenright honey bee colonies. Behav Ecol Sociobiol 32: 191–198.

Ratnieks FLW, Visscher PK (1989) Worker policing in the honey bee. Nature 342: 796–797.

Ratnieks FLW, Wenseleers T (2005) Policing insect societies. Science 307: 54–56.

Ratnieks FLW, Foster KR, Wenseleers T (2006) Conflict resolution in insect societies. Annu Rev Entomol 51: 581–608.

Ratnieks FL, Wenseleers T (2008) Altruism in insect societies and beyond: voluntary or enforced? Trends Ecol Evol 23: 45–52.

Ratnieks FL, Helanterä H (2009) The evolution of extreme altruism and inequality in insect societies. Phil Trans R Soc B Biol Sci 364: 3169–3179.

Ratnieks FLW, Foster KR, Wenseleers T (2011) Darwin’s special difficulty: the evolution of ‘neuter insects’ and current theory. Behav Ecol Sociobiol 65: 481–492.

Ravary F, Lecoutey E, Kaminski G, Châline N, Jaisson P (2007) Individual experience alone can generate lasting division of labor in ants. Curr Biol 17: 1308–1312.

Rawls J (1971) A theory of justice, Cambridge, MA: Harvard University Press.

Raymond B, Davis D, Bonsall MB (2007) Competition and reproduction in mixed infections of pathogenic and non-pathogenic Bacillus spp. J Insect Path 96: 151–155.

Raymond B, Ellis RJ, Bonsall MB (2009) Moderation of pathogen-induced mortality: the role of density in Bacillus thuringiensis virulence. Biol Lett 5: 218–220.

Real L, Caraco T (1986) Risk and foraging in stochastic environments. Annu Rev Ecol Syst 17: 371–390.

Reed DR, Lawler MP, Tordoff MG (2008) Reduced body weight is a common effect of gene knockout in mice. BMC Genet. 9: 4.

Reed RD, Papa R, Martin A, Hines HM, Counterman BA, Pardo-Diaz C, et al. (2011) Optix drives the repeated convergent evolution of butterfly wing pattern mimicry. Science 333: 1137–1141.

Rees J, Davis LV, Lenhoff HM (1970) Paths and rates of food distribution in the colonial hydroid Pennaria. Comp Biochem Physiol 34: 309–316.

Reeve H (1991) Polistes. In: Ross KG, Matthews RW, eds. The social biology of wasps. Ithaca, NY: Cornell University Press. pp 99–148.

Reeve HK, Ratnieks FLW (1993) Queen–queen conflicts in polygynous societies: mutual tolerance and reproductive skew. In: Keller L, ed. Queen number and sociality in insects. Oxford, UK: Oxford University Press. pp 45–85.

Reeve HK, Ratnieks FLW (1993) Resolutions of con£icts in polygynous societies: mutual tolerance and reproductive skew. In: Keller L, ed. Queen number and sociality in insects. Oxford, UK: Oxford University Press. pp 45–85.

Reeve HK, Keller L (1997) Reproductive bribing and policing as evolutionary mechanisms for the suppression of within-group selfishness. Am Nat 150: S42–S58.

Reeve HK, Emlen ST, Keller L (1998) Reproductive sharing in animal societies: reproductive incentives or incomplete control by dominant breeders? Behav Ecol 9: 267–278.

Reeve HK, Keller L (2001) Tests of reproductive-skew models in social insects. Annu Rev Entomol 46: 347–385.

Reeve HK, Shen SF (2006) A missing model in reproductive skew theory: the bordered tug-of-war. Proc Natl Acad Sci USA 103: 8430–8434.

Reeve HK, Shen SF (2013) Unity and disunity in the search for a unified reproductive skew theory. Anim Behav 85: 1137–1144.

Rehan SM, Schwarz MP, Richards MH (2011) Fitness consequences of ecological constraints and implications for the evolution of sociality in an incipiently social bee. Biol J Linn Soc 103: 57–67.

Rehan SM, Leys R, Schwarz MP (2012) A mid-Cretaceous origin of sociality in Xylocopine bees with only two origins of true worker castes indicates severe barriers to eusociality. PLoS ONE 7: e34690.

Reimers S, Maylor EA, Stewart N, Chater N (2009) Associations between a one-shot delay discounting measure and age, income, education and real-world impulsive behavior. Pers Indiv Differ 47: 973–978.

Reiss M (1988) Scaling of home range size: body size, metabolic needs and ecology. Trends Ecol Evol 3: 85–86.

Rembold H, Czoppelt C, Rao PJ (1974) Effect of juvenile hormone treatment on caste differentiation in the honeybee, Apis mellifera. J Insect Physiol 20: 1193–1202.

Ren J, Wang WX, Qi F (2007) Randomness enhances cooperation: A resonance-type phenomenon in evolutionary games. Phys Rev E 75: 045101.

Requejo RJ, Camacho J (2011) Evolution of cooperation mediated by limiting resources: connecting resource based models and evolutionary game theory. J Theor Biol 272: 35–41.

Requejo RJ, Camacho J (2012a) Coexistence of cooperators and defectors in well mixed populations mediated by limiting resources. Phys Rev Lett 108: 038701.

Requejo RJ, Camacho J (2012b) Analytical model for well-mixed populations of cooperators and defectors under limiting resources. Phys Rev E 85: 066112.

Requejo RJ, Camacho J (2013) Scarcity may promote cooperation in the population of simple agents. Phys Rev E 87: 022819.

Reyer HU (1980) Flexible helper structure as an ecological adaptation in the pied kingfisher (Ceryle rudis rudis L.). Behav Ecol Sociobiol 6: 219–227.

Reyer HU (1984) Investment and relatedness: a cost/benefit analysis of breeding and helping in the pied kingfisher (Ceryle rudis). Anim Behav 32: 1163–1178.

Reyer HU, Dittami JP, Hall MR (1986) Avian helpers at the nest: are they psychologically castrated? Ethology 71: 216–228.

Reynolds JD, Sheldon BC (2003) Animal behaviour: Wise fathers. Nature 422: 669–670.

Reznick D, Bryant MJ, Bashey F (2002) r- and K-selection revisited: the role of population regulation in life-history evolution. Ecology 83: 1509–1520.

Reznick DN, Bryant MJ, Roff D, Ghalambor CK, Ghalambor DE (2004) Effect of extrinsic mortality on the evolution of senescence in guppies. Nature 431: 1095–1099.

Rheindt FE, Gadau J, Strehl CP, Hölldobler B (2004) Extremely high mating frequency in the Florida harvester ant (Pogonomyrmex badius). Behav Ecol Sociobiol 56: 472–481.

Rhine RJ, Wasser SK, Norton GW (1988) 8-year study of social and ecological correlates of mortality among immature baboons of Mikumi National Park, Tanzania. Am J Primatol 16: 199–212.

Ricard J, Noat G (1986) Catalytic efficiency, kinetic co-operativity of oligometric enzymes and evolution. J Theor Biol 123: 431–451.

Rice WR (1984) Sex-chromosomes and the evolution of sexual dimorphism. Evolution 38: 735–742.

Rice WR (1992) Sexually antagonistic genes: experimental evidence. Science 256: 1436–1439.

Rice WR (1996) Sexually antagonistic male adaptation triggered by experimental arrest of female evolution. Nature 381: 232–234.

Rice WR (1998a) Male fitness increases when females are eliminated from gene pool: implications for the Y chromosome. Proc Natl Acad Sci USA 95: 6217–6221.

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.

Richard DS, Rybczynski R, Wilson TG, Wang Y, Wayne ML, et al. (2005) Insulin signaling is necessary for vitellogenesis in Drosophila melanogaster independent of the roles of juvenile hormone and ecdysteroids: Female sterility of the chico insulin signaling mutation is autonomous to the ovary. J Insect Physiol 51: 455–464.

Richard FJ, Hefetz A, Christides JP, Errard C (2004) Food influence on colonial recognition and chemical signature between nestmates in the fungus-growing ant Acromyrmex subterraneus. Chemoecology 14: 9–16.

Richard FJ, Poulsen M, Hefetz A, Errard C, Nash DR, Boomsma JJ (2007) The origin of chemical profiles of fungal symbionts and their significance for nestmate recognition in Acromyrmex leaf-cutting ants. Behav Ecol Sociobiol 61: 1637–1649.

Richard FJ, Hunt JH (2013) Intracolony chemical communication in social insects. Insect Soc 60: 275–291.

Richards EJ (2011) Natural epigenetic variation in plant species: a view from the field. Curr Opin Plant Biol 14: 204–209.

Richards MH (2001) Nesting biology and social organization of Halictus sexcinctus (Fabricius) in southern Greece. Can J Zool 79: 2210–2220.

Richards MH, Packer L (1995) Annual variation in survival and reproduction of the primitively eusocial sweat bee Halictus ligatus (Hymenoptera: Halictidae). Can J Zool 73: 933–941.

Richards MH, Packer L (1996) The socioecology of body size variation in the primitively eusocial sweat bee, Halictus ligatus (Hymenoptera: Halictidae). Oikos 77: 68–76.

Richerson PJ, Boyd R (2000) Built for speed: Pleistocene climate variation and the origin of human culture. Perspectives Ethol 13: 1–45.

Rideout RM, Rose GA, Burton MPM (2005) Skipped spawning in female iteroparous fishes. Fish Fisheries 6: 50–72.

Ridley AR, van den Heuvel IM (2012) Is there a difference in reproductive performance between cooperative and non-cooperative species? A southern African comparison. Behaviour 149: 821–848.

Ridley J, Sutherland WJ (2002) Kin competition within groups: the offspring depreciation hypothesis. Proc R Soc Lond B Biol Sci 269: 2559–2564.

Riechert SE, Roeloffs R, Echternacht AC (1986) The ecology of the cooperative spider Agelena consociata in Equatorial Africa (Araneae, Agelenidae). J Arachnol 14: 175–191.

Riehl C (2013) Evolutionary routes to non-kin cooperative breeding in birds. Proc R Soc B 280: 20132245.

Riehle MA, Brown MR (1999) Insulin stimulates ecdysteroid production through a conserved signaling cascade in the mosquito Aedes aegypti. Insect Biochem Mol Biol 29: 855–860.

Riehle MA, Fan Y, Cao C, Brown MR (2006) Molecular characterization of insulin-like peptides in the yellow fever mosquito, Aedes aegypti: expression, cellular localization, and phylogeny. Peptides 27: 2547–2560.

Rieucau G, Fernö A, Ioannou CC, Handegard NO (2014) Towards of a firmer explanation of large shoal formation, maintenance and collective reactions in marine fish. Rev Fish Biol Fisheries DOI 10.1007/s11160-014-9367-5.

Riley MA, Wertz JE (2002) Bacteriocins: evolution, ecology, and application. Annu Rev Microbiol 56: 117–137.

Riley MA, Goldstone CM, Wertz JE, Gordon D (2003) A phylogenetic approach to assessing the targets of microbial warfare. J Evol Biol 16: 690–697.

Rinkevich B, Shapira M, Weissman IL, Saito Y (1992) Allogeneic responses between 3 remote populations of the cosmopolitan ascidian Botryllus schlosseri. Zool Sci 9: 989–994.

Rissing SW, Pollock GB (1987) Queen aggression, pleometrotic advantage and brood raiding in the ant Veromessor pergandei (Hymenoptera, Formicidae). Anim Behav 35: 975–981.

Rissing SW, Pollock GB (1988) Pleometrosis and polygyny in ants. In: Jeanne RL, ed. Interindividual behavioral variability in social insects. Boulder, CO: Westview Press. pp 179–222.

Rissing SW, Pollock GB, Higgins MR, Hagen RH, Smith DR (1989) Foraging specialization without relatedness or dominance among co-founding ant queens. Nature 338: 420–422.

Rissing SW, Johnson RA, Martin JW (2000) Colony founding behavior of some desert ants: geographic variation in metrosis. Psyche 103: 95–105.

Rittschof CC, Robinson GE (2014) Genomics: moving behavioural ecology beyond the phenotypic gambit. Anim Behav 92: 263–270.

Ritz DA (1997) Costs and benefits as a function of group size: experiments on a swarming mysid, Paramesopodopsis rufa Fenton. In: Parrish JK, Hamner WM, eds. Animal groups in three dimensions. Cambridge, UK: Cambridge University Press. pp 194–206.

Roberts G (1997) How many birds does it take to put a flock to flight? Anim Behav 54: 1517–1522.

Roberts RL, Williams JR, Wang AK, Carter CS (1998) Cooperative breeding and monogamy in prairie voles: influence of the sire and geographical variation. Anim Behav 55: 1131–1140.

Robinson GE (1987) Regulation of honey bee age polyethism by juvenile hormone. Behav Ecol Sociobiol 20: 329–338.

Robinson GE (1992) Regulation of division of labor in insect societies. Annu Rev Entomol 37: 637–665.

Robinson GE (1999) Integrative animal behaviour and sociogenomics. Trends Ecol Evol 14: 202–205.

Robinson GE, Page RE Jr (1988) Genetic determination of guarding and undertaking in honey-bee colonies. Nature 333: 356–358.

Robinson GE, Strambi C, Strambi A, Feldlaufer MF (1991) Comparison of juvenile hormone and ecdysteroid hemolymph titers in adult worker and queen honey bees (Apis mellifera). J Insect Physiol 37: 929–935.

Robinson GE, Vargo EL (1997) Juvenile hormone in adult eusocial Hymenoptera: gonadotropin and behavioral pacemaker. Arch Insect Biochem Physiol 35: 559–583.

Robinson GE, Ben-Shahar Y (2002) Social behavior and comparative genomics: new genes or new gene regulation? Genes Brain Behav 1: 197–203.

Robinson GE, Grozinger CM, Whitfield CW (2005) Sociogenomics: social life in molecular terms. Nat Rev Genet 6: 257–270.

Robson AJ (1996) A biological basis for expected and non-expected utility. J Econ Theory 68: 397–424.

Rockenbach B, Milinski M (2009) How to treat those of ill repute. Nature 457: 39–40.

Rockman MV, Kruglyak L (2006) Genetics of global gene expression. Nat Rev Genet 7: 862–872.

Rödel HG, Starkloff A, Bautista A, Friedrich AC, Von Holst D (2008) Infanticide and maternal offspring defence in European rabbits under natural breeding conditions. Ethology 114: 22–31.

Rodemann JF, Dubberke ER, Reske KA, Seo DH, Stone CD (2007) Incidence of Clostridium difficile infection in inflammatory bowel disease. Clin Gastroenterol Hepatol 5: 339–344.

Rodman PS (1988) Resources and group sizes of primates. In: Slobodchikoff CN, ed. The ecology of social behavior. San Diego, CA: Academic Press. pp 83–108.

Rodrigues SM, Saslow LR, Garcia N, John OP, Keltner D (2009) Oxytocin receptor genetic variation relates to empathy and stress reactivity in humans. Proc Natl Acad Sci USA 106: 21437–21441.

Rodriguez R, Redman R (2008) More than 400 million years of evolution and some plants still can't make it on their own: plant stress tolerance via fungal symbiosis. J Exp Bot 59: 1109–1114.

Rodríguez- Gironés MA (1996) Siblicide: the evolutionary blackmail. Am Nat 148: 101–122.

Roff DA (1996) The evolution of threshold traits in animals. Q Rev Biol 71: 3–36.

Roff DA (1998) The maintenance of phenotypic and genetic variation in threshold traits by frequency-dependent selection. J Evol Biol 11: 513–529.

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.

Rogers AR (1994) Evolution of time preference by natural selection. Am Econ Rev 84: 460–481.

Rogers C, Reale V, Kim K, Chatwin H, Li C, Evans P, de Bono M (2003) Inhibition of Caenorhabditis elegans social feeding by FMRFamide-related peptide activation of NPR-1. Nat Neurosci 6: 1178–1185.

Rohwer S (1978) Parent cannibalism of offspring and egg raiding as a courtship strategy. Am Nat 112: 429–439.

Roiloa SR, Alpert P, Tharayil N, Hancock G, Bhowmik PC (2007) Greater capacity for division of labour in clones of Fragaria chiloensis from patchier habitats. J Ecol 95: 397–405.

Roitberg BD, Mangel M, Lalonde RG, Roitberg CA, van Alphen JJM, Vet L (1992) Seasonal dynamic shifts in patch exploitation by parasitic wasps. Behav Ecol 3: 156–165.

Roitberg BD, Sircom J, Roitberg CA, van Alphen JJM, Mangel M (1993) Life expectancy and reproduction. Nature 364: 108.

Roitberg BD, Boivin G, Vet LEM (2001) Fitness, parasitoids, and biological control: an opinion. Can Entomol 133: 429–438.

Rolland C, Danchin E, de Fraipont M (1998) The evolution of coloniality in birds in relation to food, habitat, predation, and life-history traits: a comparative analysis. Am Nat 151: 514–529.

Rollinson N, Brooks RJ (2007) Proximate constraints on reproductive output in a northern population of Painted Turtles: An empirical test of the bet-hedging paradigm. Can J Zool 85: 177–184.

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.

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.

Ronay R, Carney DR (2013) Testosterone’s negative relationship with empathic accuracy and perceived leadership ability. Soc Psychol Person Sci 4: 92–99.

Ronce O (2007) How does it feel to be like a rolling stone? Ten questions about dispersal evolution. Annu Rev Ecol Evol Syst 38: 231–253.

Rood JP (1986) Ecology and social evolution in the mongooses. In: Rubenstein DI, Wrangham RW, eds. Ecological aspects of social evolution: birds and mammals. Princeton, NJ: Princeton University Press. pp 131–152.

Rood JP (1978) Dwarf mongoose helpers at the den. J Comp Ethol 48: 277–278.

Rood JP (1990) Group size, survival, reproduction, and routes to breeding in dwarf mongooses. Anim Behav 39: 566–572.

Rosati AG, Stevens JR, Hare B, Hauser MD (2007) The evolutionary origins of human patience: temporal preferences in chimpanzees, bonobos, and human adults. Curr Biol 17: 1663–1668.

Rösch GA (1930) Untersuchungen über die Arbeitsteilung im Bienenstaat. II. Die Tätigkeiten der Arbeitsbienen unter experimentell veränderten Bedingungen. Z Vgl Physiol 12: 1–71.

Röseler PF (1991) Reproductive competition during colony establishment. In: Ross KG, Matthews R, eds. The social biology of wasps. Ithaca, NY: Cornell University Press. pp 309–335.

Röseler PF, Röseler I, van Honk CGJ (1981) Evidence for inhibition of corpora allata activity in workers of Bombus terrestris by a pheromone from the queen's mandibular glands. Experientia 37: 348–351.

Röseler PF, van Honk CGJ (1990) Castes and reproduction in bumblebees. In: Engels W, ed. Social insects: an evolutionary approach to castes and reproduction. Berlin, Germany: Springer Verlag. pp 147–166.

Rosengarten RD, Nicotra ML (2011) Model systems of invertebrate allorecognition. Curr Biol 21: R82–R92.

Rosengaus RB, Traniello JFA (2001) Disease susceptibility and the adaptive nature of colony demography in the dampwood termite Zootermopsis angusticollis. Behav Ecol Sociobiol 50: 546–556.

Rosengren R, Sundström L, Fortelius W (1993) Monogyny and polygyny in Formica ants, the result of alternative dispersal tactics. In: Keller L, ed. Queen number and sociality in insects. Oxford, UK: Oxford University Press. pp 308–333.

Rosenzweig ML (1995) Species diversity in space and time. Cambridge, UK: Cambridge University Press.

Ross HE, Young LJ (2009) Oxytocin and the neural mechanisms regulating social cognition and affiliative behavior. Front Neuroendocrinol 30: 534–547.

Ross HE, Freeman SM, Spiegel LL, Ren XH, Terwilliger EF, Young LJ (2009) Variation in oxytocin receptor density in the nucleus accumbens has differential effects on affiliative behaviors in monogamous and polygamous voles. J Neurosci 29: 1312–1318.

Ross KG, Fletcher DJ (1985) Comparative study of genetic and social structure in two forms of the fire ant Solenopsis invicta (Hymenoptera: Formicidae). Behav Ecol Sociobiol 17: 349–356.

Ross L, Gardner A, Hardy N, West SA (2013) Ecology, not the genetics of sex determination, determines who helps in eusocial populations. Curr Biol 23: 2383–2387.

Rosset H, Chapuisat M (2006) Sex allocation conflict in ants: when the queen rules. Curr Biol 16: 328–331.

Ross-Gillespie A, Gardner A, West SA, Griffin AS (2007) Frequency dependence and cooperation: theory and a test with bacteria. Am Nat 170: 331–342.

Ross-Gillespie A, Kümmerli R (2014) Collective decision-making in microbes. Front Microbiol 5: 54.

Roubik DW (1982) The ecological impact of nectar-robbing bees and pollinating hummingbirds on a tropical shrubs. Ecology 63: 354–360.

Roulin A, Dreiss AN (2012) Sibling competition and cooperation over parental care. In: Royle NJ, Smiseth PT, Kölliker M, eds. The evolution of parental care. Oxford, UK: Oxford University Press. pp 133–149.

Roulston TH, Buczkowski G, Silverman J (2003) Nestmate discrimination in ants: effect of bioassay on aggressive behavior. Insect Soc 50: 151–159.

Rousset F (2004) Genetic structure and selection in subdivided populations. Princeton, NJ: Princeton University Press.

Rousset F, Roze D (2007) Constraints on the origin and maintenance of genetic kin recognition. Evolution 61: 2320–2330.

Rousset F, Lion S (2011) Much ado about nothing: Nowak et al.’s charge against inclusive fitness theory. J Evol Biol 24: 1386–1392.

Roux W (1881) Der Kampf der Theile im Organismus. Leipzig, Germany: W. Engelmann.

Rouzine IM, Rodrigo A, Coffin JM (2001) Transition between stochastic evolution and deterministic evolution in the presence of selection: general theory and application to virology. Microbiol Mol Biol Rev 65: 151–185.

Rowley I (1981) The communal way of life in the Splendid Wren, Malurus splendens. Z Tierpsychol 55: 228–267.

Rowley I, Russell EM (1990) Splendid fairy-wrens: demonstrating the importance of longevity. In: Stacey PB, Koenig WD, eds. Cooperative breeding in birds: Long-term studies of ecology and behavior. Cambridge, UK: Cambridge University Press. pp 1–30.

Royle NJ, Hartley IR, Parker GA (2002) Begging for control: when are offspring solicitation behaviours honest? Trends Ecol Evol 17: 434–440.

Royle NJ, Smiseth PT, Kölliker M, eds. (2012) The evolution of parental care. Oxford, UK: Oxford University Press.

Rubenstein DI (1982) Risk, uncertainty and evolutionary strategies. In: King's College Sociobiology Group, ed. Current problems in sociobiology. Cambridge, UK: Cambridge University Press. pp 91–111.

Rubenstein DI, Wrangham RW, eds. (1986) Ecological aspects of social evolution: birds and mammals. Princeton, NJ: Princeton University Press.

Rubenstein DR (2011) Spatiotemporal environmental variation, risk aversion, and the evolution of cooperative breeding as a bet-hedging strategy. Proc Natl Acad Sci USA 108(Suppl. 2): 10816–10822.

Rubenstein DR, Lovette IJ (2007) Temporal environmental variability drives the evolution of cooperative breeding in birds. Curr Biol 17: 1414–1419.

Rudge T, Burrage K (2008) Effects of intrinsic and extrinsic noise can accelerate juxtacrine pattern formation. Bull Math Biol 70: 971–991.

Ruff JS, Nelson AC, Kubinak JL, Potts WK (2012) MHC signaling during social communication. Adv Exp Med Biol 738: 290–313.

Rumble AC, Van Lange PAM, Parks CD (2010) The benefits of empathy: when empathy may sustain cooperation in social dilemmas. Eur J Soc Psychol 40: 856–866.

Runnegar B (1986) Molecular palaeontology. Palaeontology 29: 1–24.

Ruse M (1999) The Darwinian Revolution: Science Red in Tooth and Claw, 2nd edn. Chicago, IL: University of Chicago 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.

Russell AF (2004) Mammals: comparisons and contrasts. In: Koenig WD, Dickinson JL, eds. Ecology and evolution of cooperative breeding in birds. Cambridge, UK: Cambridge University Press. pp 210–227.

Russell AF, Hatchwell BJ (2001) Experimental evidence for kin-biased helping in a cooperatively breeding vertebrate. Proc R Soc Lond Ser B Biol Sci 268: 2169–2174.

Russell EM (1989) Cooperative breeding: a Gondwanan perspective. Emu 89: 61–62.

Russell EM (2000) Avian life histories: Is extended parental care the southern secret? Emu 100: 377–399.

Russell EM, Rowley I (1993) Philopatry or dispersal: competition for territory vacancies in the splendid fairywren, Malurus splendens. Anim Behav 45: 519–539.

Russell EM, Yom-Tov Y, Geffen E (2004) Extended parental care and delayed dispersal: Northern, tropical, and southern passerines compared. Behav Ecol 15: 831–838.

Russell S, Kelley J, Graves J, Magurran A (2004) Kin structure and shoal composition dynamics in the guppy, Poecilia reticulata. Oikos 106: 520–526.

Russo-Marie F, Roederer M, Sager B, Herzenberg LA, Kaiser D (1993) Beta- galactosidase activity in single differentiating bacterial cells. Proc Natl Acad Sci USA 90: 8194–8198.

Ruther J, Sieben S, Schricker B (2002) Nestmate recognition in social wasps: manipulation of hydrocarbon profiles induces aggression in the European hornet. Naturwissenschaften 89: 111–114.

Ryan F (2002) Darwin’s blind spot. Boston, MA: Houghton Mifflin Harcourt.

Ryan FP (2006) Genomic creativity and natural selection: a modern synthesis. Biol J Linn Soc 88: 655–672.

Ryan MJ (1985) The Tungara frog: a study in sexual selection and communication. Chicago, IL: University of Chicago Press.

Rychlik L, Zwolak R (2006) Interspecific aggression and behavioural dominance among four sympatric species of shrews. Can J Zool 84: 434–448.

Ryding E, Lindström M, Träskman-Bendz L (2008) The role of dopamine and serotonin in suicidal behaviour and aggression. Progr Brain Res 172: 307–315.

Rypstra AL (1990) Prey capture and feeding efficiency of social and solitary spiders: a comparison. Acta Zool Fenn 190: 339–343.

Rypstra AL, Tirey RS (1991) Prey size, prey perishability and group foraging in a social spider. Oecologia 86: 25–30.

Rytkönen S, Soppela M (1995) Vicinity of sparrowhawk nest affects willow tit nest defense. Condor 97: 1074–1078.

Sääksvuori L, Mappes T, Puurtinen M (2011) Costly punishment prevails in intergroup conflict. Proc R Soc B Biol Sci 278: 3428–3436.

Sachs JL (2006) Cooperation within and among species. J Evol Biol 19: 1415–1418.

Sachs JL (2013) Origins, evolution, and breakdown of bacterial symbiosis. In: Levin SA, ed. Encyclopedia of biodiversity, 2nd edn, Vol. 5. Waltham, MA: Academic Press. pp 637–644.

Sachs JL, Mueller UG, Wilcox TP, Bull JJ (2004) The evolution of cooperation. Q Rev Biol 79: 135–160.

Sachs JL, Simms EL (2006) Pathways to mutualism breakdown. Trends Ecol Evol 21: 585–592.

Sachs JL, Wilcox TP (2006) A shift to parasitism in the jellyfish symbiont Symbiodinium microadriaticum Proc R Soc Lond B 273: 425–429.

Sachs JL, Rubenstein DR (2007) The evolution of cooperative breeding; is there cheating? Behav Process 76: 131–137.

Sachs JL, Russell JE, Lii YE, Black KC, Lopez G, Patil SA (2010) Host control over infection and proliferation of a cheater symbiont. J Evol Biol 23: 1919–1927.

Sachs JL, Skophammer RG, and Regus JU (2011a) Evolutionary transitions in bacterial symbiosis. Proc Natl Acad Sci USA 108: 10800–10807.

Sachs JL, Russell JE, Hollowell AC (2011b) Evolutionary instability of symbiotic function in Bradyrhizobium japonicum. PLoS ONE 6: e26370.

Sachs T, Novoplansky A (1997) What does aclonal organization suggest concerning clonal plants? In: de Kroon H, van Groenendael J, eds. The ecology and evolution of clonal plants. Leiden, the Netherlands: Backhuys Publishers. pp 55–77.

Saito K (2012) Social preferences under uncertainty: equality of opportunity vs. equality of outcome. Social Science Working Paper. 1362. California Institute of Technology. http://hss-prod-storage.cloud.caltech.edu.s3.amazonaws.com/hss_working_papers/sswp1362.pdf

Saito Y (1997) Sociality and kin selection in Acari. In: Crespi BJ, Choe JC, eds. The evolution of social behavior in insects and arachnids. Cambridge, UK: Cambridge University Press. pp 443–457.

Sakagami SF (1977) Seasonal change of nest survival and related aspects in an aggregation of Lasioglossum duplex (Dalla Torre), a eusocial halictine bee (Hymenoptera: Halictidae). Res Popul Ecol 19: 69–86.

Sakagami SF, Hayashida K (1968) Bionomics and sociology of the summer matrifilial phase in the social halictine bee, Lasioglossum duplex. J Fac Sci Hokkaido Univ Zool (Ser VI) 16: 413–513.

Sakagami SF, Munakata M (1972) Distribution and bionomics of a transpalearctic eusocial halictine bee, Lasioglossum (Evylaeus) calceatum in northern Japan with notes on its solitary life cycle at high altitude. J Fac Sci Hokkaido Univ Ser VI Zool 18: 411–439.

Saltzman W, Ahmed S, Fahimi A, Wittwer DJ, Wegner FH (2006) Social suppression of female reproductive maturation and infanticidal behavior in cooperatively breeding Mongolian gerbils. Horm Behav 49: 527–537.

Saltzman W, Digby LJ, Abbott DH (2009) Reproductive skew in female common marmosets: what can proximate mechanisms tell us about ultimate causes? Proc R Soc B Biol Sci 276: 389–399.

Salzman AG (1985) Habitat selection in a clonal plant. Science 228: 603–604.

Salzman AG, Parker MA (1985) Neighbors ameliorate local salinity stress for a rhizomatous plant in a heterogeneous environment. Oecologia 65: 273–277.

Sampathkumar B, Napper S, Carrillo CD, Willson P, Taboada E, Nash JH, et al. (2006) Transcriptional and translational expression patterns associated with immobilized growth of Campylobacter jejuni. Microbiology 152: 567–577.

Samuel CT (1987) Factors affecting colony size in the stenogastrine wasp Liostenogaster flavolineata. PhD thesis. Kuala Lumpur, Malaysia: University of Malaya.

Sanchez Y, Taulien J, Borkovich KA, Lindquist S (1992) Hsp104 is required for tolerance to many forms of stress. EMBO J 11: 2357–2364.

Sandoz KM, Mitzimberg SM, Schuster M (2007) Social cheating in Pseudomonas aeruginosa quorum sensing. Proc Natl Acad Sci USA 104: 15876–15881.

Sansom OJ, Mansergh FC, Evans MJ, Wilkins JA, Clarke AR (2007) Deficiency of SPARC suppresses intestinal tumorigenesis in APCMin/+ mice. Gut 56: 1410–1414.

Santangelo N, Bass A (2006) New insights into neuropeptide modulation of aggression: field studies of arginine vasotocin in a territorial tropical damselfish. Proc R Soc B Biol Sci 273: 3085–3092.

Santorelli LA, Thompson CR, Villegas E, Svetz J, Dinh C, Parikh A, et al. (2008) Facultative cheater mutants reveal the genetic complexity of cooperation in social amoebae. Nature 451: 1107–1110.

Santos FC, Pacheco JM (2005) Scale-free networks provide a unifying framework for the emergence of cooperation. Phys Rev Lett 95: 098104.

Santos FC, Pacheco JM, Lenaerts T (2006a) Evolutionary dynamics of social dilemmas in structured heterogeneous populations. Proc Natl Acad Sci USA 103: 3490–3494.

Santos FC, Pacheco JM, Lenaerts T (2006b) Cooperation prevails when individuals adjust their social ties. PLoS Comput Biol 2: e140.

Santos FC, Santos MD, Pacheco JM (2008) Social diversity promotes the emergence of cooperation in public goods games. Nature 454: 213–216.

Santos FC, Pinheiro FL, Lenaerts T, Pacheco JM (2012) The role of diversity in the evolution of cooperation. J Theor Biol 299: 88–96.

Santos PC, Schinemann JA, Gabaro J, de Graca Bichalho M (2004) New evidence on MHC-based disassortative odour preferences in humans: a study with 58 Brazilian students. Hum Immunol 65: S108.

Sapolsky RM (2004) Why Zebras don’t get ulcers. New York, NY: Owl Books.

Saran S, Meima ME, Alvarez-Curto E, Weening KE, Rozen DE, Schaap P (2002) cAMP signaling in Dictyostelium. Complexity of cAMP synthesis, degradation and detection. J Muscle Res Cell Motil 23: 793–802.

Sasagawa H, Sasaki M, Okada I (1989) Hormonal control of the division of labor of adult honey bees (Apis mellifera L.). I. Effect of methoprene on corpora allata and hypopharyngeal gland, and its α-glucosidase activity. Appl Entomol Zool 24: 66–77.

Sasaki K, Satoh T, Obara Y (1996) Cooperative foundation of colonies by unrelated foundresses in the ant Polyrhachis moesta. Insect Soc 43: 217–226.

Sauer K, Camper AK, Ehrlich GD, Costerton JW, Davies DG (2002) Pseudomonas aeruginosa displays multiple phenotypes during development as a biofilm. J Bacteriol 184: 1140–1154.

Saunders B (2008) The equality of lotteries. Philosophy 83: 359–372.

Savage A, Ziegler TE, Snowdon CT (1988) Sociosexual development, pair bond formation, and mechanisms of fertility suppression in female cotton-top tamarins (Saguinus oedipus oedipus). Am J Primatol 14: 345–359.

Schaber JA, Carty NL, McDonald NA, Graham ED, Cheluvappa R, Griswold JA, Hamood AN (2004) Analysis of quorum sensing-deficient clinical isolates of Pseudomonas aeruginosa. J Med Microbiol 53: 841–853.

Schall JD (1999) Weighing the evidence: how the brain makes a decision. Nat Neurosci 2: 108–109.

Schall JD (2001) Neural basis of deciding, choosing and acting. Nat Rev Neurosci 2: 33–42.

Scharloo W (1991) Canalization: genetic and developmental aspects. Annu Rev Ecol Syst 22: 65–93.

Schaub R, Mumme RL, Woolfenden GE (1992) Predation on the eggs and nestlings of Florida scrub jays. Auk 109: 585–593.

Schausberger P (2007) Kin recognition by juvenile predatory mites: prior association or phenotype matching? Behav Ecol Sociobiol 62: 119–125.

Scheel D, Packer C (1991) Group hunting behavior of lions—a search for cooperation. Anim Behav 41: 697–709.

Schembri MA, Hjerrild L, Gjermansen M, Klemm P (2003) Differential expression of the Escherichia coli autoaggregation factor antigen 43. J Bacteriol 185: 2236–2242.

Schenk HJ (1999) Clonal splitting in desert shrubs. Plant Ecol 141: 41–52.

Schildberg-Hörisch H (2010) Is the veil of ignorance only a concept about risk? An experiment. J Public Econ 94: 1062–1066.

Schino G, Aureli F (2010) The relative roles of kinship and reciprocity in explaining primate altruism. Ecol Lett 13: 45–50.

Schluter D (1994) Experimental evidence that competition promotes divergence in adaptive radiation. Science 266: 798–801.

Schluter D (1996) Ecological causes of adaptive radiation. Am Nat 148: S40–S64.

Schluter D (2000) The ecology of adaptive radiation. Oxford, UK: Oxford University Press.

Schluter D (2001) Ecology and the origin of species. Trends Ecol Evol 16: 372–380.

Schluter D (2010) Resource competition and coevolution in sticklebacks. Evo Edu Outreach 3: 54–61.

Schmidt Capella IC, Hartfelder K (1998) Juvenile hormone effect on DNA synthesis and apoptosis in caste-specific differentiation of the larval honey bee (Apis mellifera L.) ovary. J Insect Physiol 44: 385–391.

Schmidt Capella IC, Hartfelder K (2002) Juvenile-hormone-dependent interaction of actin and spectrin is crucial for polymorphic differentiation of the larval honey bee ovary. Cell Tissue Res 307: 265–272.

Schmitt J, Ehrhardt DW (1987) A test of the sib-competition hypothesis for outcrossing advantage in Impatiens capensis. Evolution 41: 579–590.

Schmitt RJ, Holbrook SJ (2003) Mutualism can mediate competition and promote coexistence. Ecol Lett 6: 898–902.

Schmitz A, Perry SF (1999) Stereological determination of tracheal volume and diffusing capacity of the tracheal walls in the stick insect Carausius morosus (Phasmatodea, Lonchodidae). Physiol Biochem Zool 72: 205–218.

Schmitz OJ, Ritchie ME (1991) Optimal diet selection with variable nutrient intake – balancing reproduction with risk of starvation. Theor Popul Biol 39: 100–114.

Schoech SJ, Mumme RL, Moore MC (1991) Reproductive endocrinology and mechanisms of breeding inhibition in cooperatively breeding Florida scrub jays (Aphelocoma c. coerulescens). Condor 93: 354–364.

Schoener TW (1983) Field experiments on interspecific competition. Am Nat 122: 240–285.

Schradin C (2013) Intraspecific variation in social organization by genetic variation, developmental plasticity, social flexibility or entirely extrinsic factors. Phil Trans R Soc B 368: 20120346.

Schulke O, Bhagavatula J, Vigilant L, Ostner J (2010) Social bonds enhance reproductive success in male macaques. Curr Biol 20: 2207–2210.

Schultz D, Wolynes PG, Ben Jacob E, Onuchic JN (2009) Deciding fate in adverse times: sporulation and competence in Bacillus subtilis. Proc Natl Acad Sci USA 106: 21027–21034.

Schuurman T (1980) Hormonal correlates of agonistic behavior in adult male rats. Prog Brain Res 53: 415–420.

Schwagmeyer PL, Parker GA (1987) Queuing for mates in the thirteen-lined ground squirrel. Anim Behav 35: 1015–1025.

Schwagmeyer PL, Mock DW (2008) Parental provisioning and offspring fitness: size matters. Anim Behav 75: 291–298.

Schwander T, Lo N, Beekman M, Oldroyd B, Keller L (2010) Nature versus nurture in social insect caste differentiation. Trends Ecol Evol 25: 275–357.

Schwaninger M (2004) What can cybernetics contribute to the conscious evolution of organizations and society? Syst Res Behav Sci 21: 515–527.

Schwartz MW, Hoeksema JD (1998) Specialization and resource trade: biological markets as a model of mutualisms. Ecology 79: 1029–1038.

Schwarz MP, Silberbauer LX, Hurst PS (1997) Intrinsic and extrinsic factors associated with social evolution in allodapine bees. In: Choe JC, Crespi BJ, eds. Social behavior in insects and arachnids. Cambridge, UK: Cambridge University Press. pp 333–346.

Schwarz MP, Bull NJ, Hogendoorn K (1998) Evolution of sociality in the allodapine bees: a review of sex allocation, ecology, and evolution. Insect Soc 45: 349–368.

Scofield VL, Schlumpberger JM, West LA, Weissman IL (1982a) Protochordate allorecognition is controlled by a MHC-like gene system. Nature 295: 499–502.

Scofield VL, Schlumpberger JM, Weissman IL (1982b) Colony specificity in the colonial tunicate Botryllus and the origins of vertebrate immunity. Am Zool 22: 783–794.

Scott MP (2006a) The role of juvenile hormone in competition and cooperation by burying beetles. J Insect Physiol 52: 1005–1011.

Scott MP (2006b) Resource defense and juvenile hormone: the ‘challenge hypothesis’ extended to insects. Horm Behav 49: 276–281.

Sear R, Mace R (2008) Who keeps children alive? A review of the effects of kin on child survival. Evol Hum Behav 29: 1–18.

Seddon N, Amos W, Adcock G, Johnson P, Kraaijeveld K, Kraaijeveld-Smit FJL, et al. (2005) Mating system, philopatry and patterns of kinship in the cooperatively breeding subdesert mesite Monias benschi. Mol Ecol 14: 3573–3583.

Sedgley CM, Clewell DB, Flannagan SE (2009) Plasmid pAMS1-encoded, bacteriocin-related ‘‘siblicide’’ in Enterococcus faecalis. J Bacteriol 191: 3183–3188.

Seeley TD (1985) Honeybee ecology: A study of adaptation in social life. Princeton, NJ: Princeton University Press.

Seeley TD (1989) The honey bee colony as a superorganism. Am Sci 77: 546–553.

Seeley T (1995) The wisdom of the hive: the social physiology of honey bee colonies. Cambridge, MA: Harvard University Press.

Seeley TD (1997) Honey bee colonies are group-level adaptive units. Am Nat 150(S1): S22–S41.

Seeley TD, Visscher PK (2004) Quorum sensing during nest-site selection by honeybee swarms. Behav Ecol Sociobiol 56: 594–601.

Seeley TD, Tarpy DR (2007) Queen promiscuity lowers disease within honeybee colonies. Proc R Soc B Biol Sci 274: 67–72.

Sefton M, Shupp R, Walker JM (2007) The effect of rewards and sanctions in provision of public goods. Econ Inq 45: 671–690.

Seger J (1983) Partial bivoltinism may cause alternating sex-ratio biases that favour eusociality. Nature 301: 59–62.

Seger J (1993) Opportunities and pitfalls in co-operative reproduction. In: Keller, L, ed. Queen number and sociality in insects. Oxford, UK: Oxford University Press. pp 1–15.

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.

Seghers BH (1974) Schooling behaviour in the guppy (Poecilia reticulata): an evolutionary response to predation. Evolution 28: 486–489.

Segraves K, Althoff D, Pellmyr O (2005) Limiting cheaters in mutualism: Evidence from hybridization between mutualist and cheater yucca moths. Proc Biol Sci 272: 2195–2201.

Seibt U, Wickler W (1988) Interspecific tolerance in social Stegodyphus spiders (Eresidae, Araneae). J Arachnol 16: 35–39.

Seiler M, Schwitzer C, Gamba M, Holderied MW (2013) Interspecific semantic alarm call recognition in the solitary Sahamalaza sportive lemur, Lepilemur sahamalazensis. PLoS ONE 8: e67397.

Selander RK (1964) Speciation in wrens of the genus Campylorhynchus. Univ Calif Publ Zool 74: 1–224.

Semmann D, Krambeck H, Milinski M (2003) Volunteering leads to rock-paper-scissors dynamics in a public goods game. Nature 425: 390–393.

Semsar K, Kandel FLM, Godwin J (2001) Manipulations of the AVT system shift social status and related courtship and aggressive behaviour in the bluehead wrasse. Horm Behav 40: 21–31.

Seo D, Patrick CJ, Kennealy PJ (2008) Role of serotonin and dopamine system interactions in the neurobiology of impulsive aggression and its comorbidity with other clinical disorders. Aggress Violent Beh 13: 383–395.

Seppä P, Queller DC, Strassmann JE (2002) Reproduction in foundress associations of the social wasp, Polistes carolina: conventions, competition, and skew. Behav Ecol 13: 531–542.

Sernland E, Olsson A, Holmgren NMA (2003) Does information sharing promote group foraging? Proc R Soc Lond Ser B 270: 1137–1141.

Sethi R, Somanathan E (2001) Preference evolution and reciprocity. J Econ Theor 97: 273–297.

Seyfarth RM, Cheney DL (1983) Grooming, alliances and reciprocal altruism in vervet monkeys. Nature 308: 541–543.

Seyfarth R, Cheney D (1990) The assessment by vervet monkeys of their own and another species alarm calls. Anim Behav 40: 754–764.

Seyfarth RM, Cheney DL (2012) The evolutionary origins of friendship. Annu Rev Psychol 63: 4.1–4.25.

Shadlen MN, Newsome WT (2001) Neural basis of a perceptual decision in the parietal cortex (area LIP) of the rhesus monkey. J Neurophys 86: 1916–1936.

Shafir E, Tversky A (1992) Thinking through uncertainty: nonconsequential reasoning and choice. Cogn Psychol 24: 449–474.

Shakarad M, Gadagkar R (1995) Colony founding in the primitively eusocial wasp Ropalidia marginata (Hymenoptera: Vespidae). Ecol Entomol 20: 273–282.

Shao XL, He SY, Zhuang XY, Fan Y, Li YH, Yao YG (2014) mRNA expression and DNA methylation in three key genes involved in caste differentiation in female honeybees (Apis mellifera). Zool Res 35: 92–98.

Shapiro JA (1988) Bacteria as multicellular organisms. Sci Am 258: 82–89.

Shapiro JA (1998) Thinking about bacterial populations as multicellular organisms. Annu Rev Microbiol 52: 81–104.

Shapiro JA, Dworkin M (1997) Bacteria as multicellular organisms. Oxford, UK: Oxford University Press.

Shapiro LE, Dewsbury DA (1990) Differences in affiliative behavior, pair bonding, and vaginal cytology in two species of vole (Microtus ochrogaster and M. montanus). J Comp Psychol 104: 268–274.

Sharp SP, Clutton-Brock TH (2011) Reluctant challengers: why do subordinate female meerkats rarely displace their dominant mothers? Behav Ecol 22: 1337–1343.

Shaulsky G, Loomis WF (1993) Cell type regulation in response to expression of ricin A in Dictyostelium. Dev Biol 160: 85–98.

Shaulsky G, Loomis WF (1995) Mitochondrial DNA replication but no nuclear DNA replication during development of Dictyostelium. Proc Natl Acad Sci USA 92: 5660–5663.

Shaulsky G, Kessin RH (2007) The cold war of the social amoebae. Curr Biol 17: R684–R692.

Shaw E (1978) Schooling fishes. Am Sci 66: 166–175.

Shear W, Kukalová-Peck J (1990) The ecology of Paleozoic terrestrial arthropods: the fossil evidence. Can J Zool 68: 1807–1834.

Sheldon BC (2002) Relating paternity to paternal care. Phil Trans R Soc Lond Ser B Biol Sci 357: 341–350.

Shelley EL, Tanaka MY, Ratnathicam AR, Blumstein DT (2004) Can Lanchester's laws help explain interspecific dominance in birds? Condor 106: 395–400.

Shellman-Reeve JS (1997) The spectrum of eusociality in termites. In: Choe JC, Crespi BJ, eds. The evolution of social behavior in insects and arachnids. Cambridge, UK: Cambridge University Press. pp 52–93.

Shen SF (2009) The evolution of non-kin cooperation in joint-nesting Taiwan yuhinas, yuhina brunneiceps. Doctoral dissertation, Cornell University. eCommons@Cornell. http://hdl.handle.net/1813/12833.

Shen SF, Vehrencamp SL, Johnstone RA, Chen HC, Chan SF, et al. (2012) Unfavourable environment limits social conflict in Yuhina brunneiceps. Nat Commun 3: 885.

Shenoy M, Borges R (2008) A novel mutualism between an ant-plant and its resident pollinator. Naturwissenschaften 95: 61–65.

Sher G (1980) What makes a lottery fair? Noûs 14: 203–216.

Sherborne AL, Thom MD, Paterson S, Jury F, Ollier WE, Stockley P, et al. (2007) The genetic basis of inbreeding avoidance in house mice. Curr Biol 17: 2061–2066.

Sherman PW (1977) Nepotism and the evolution of alarm calls. Science 197: 1246–1253.

Sherman PW (1980) The limits of ground squirrel nepotism. In: Barlow GW, Silverberg J, eds. Sociobiology: beyond nature/nurture? Boulder, CO: Westview Press. pp 505–544.

Sherman PW (1981) Kinship, demography, and Belding’ s ground squirrel nepotism. Behav Ecol Sociobiol 8: 251–259.

Sherman PW, Holmes WG (1985) Kin recognition: issues and evidence. Fortschr Zool 31: 437–460.

Sherman PW, Seeley TD, Reeve HK (1988) Parasites, pathogens and polyandry in social hymenoptera. Am Nat 131: 602–610.

Sherman PW, Lacey EA, Reeve HK, Keller L (1995) The eusociality continuum. Behav Ecol 6: 102–108.

Sherman PW, Reeve HK, Pfennig DW (1997) Recognition systems. In: Krebs JR, Davies NB, eds. Behavioural ecology, 4th edn. Oxford, UK: Blackwell Scientific. pp 69–96.

Shibao H (1999) Lack of kin discrimination in the eusocial aphid Pseudoregma bambucicola (Homoptera: Aphididae). J Ethol 17: 17–24.

Shimkets LJ (1999) Intercellular signaling during fruiting-body development of Myxococcus xanthus. Annu Rev Microbiol 53: 525–549.

Shirtcliff EA, Vitacco MJ, Graf AR, Gostisha AJ, Merz JL, Zahn-Waxler C (2009) Neurobiology of empathy and callousness: implications for the development of antisocial behavior. Behav Sci Law 27: 137–171.

Shreeves G, Field J (2002) Group size and direct fitness in social queues. Am Nat 159: 81–95.

Shreeves G, Cant MA, Bolton A, Field J (2003) Insurance-based advantages for subordinate co-foundresses in a temperate paper wasp. Proc R Soc Lond B 270: 1617–1622.

Shutters ST (2012) Punishment leads to cooperative behavior in structured societies. Evol Comput 20: 301–319.

Shykoff JA, Schmid-Hempel P (1991) Genetic relatedness and eusociality: parasite-mediated selection on the genetic composition of groups. Behav Ecol Sociobiol 28: 371–376.

Sibly RM (1983) Optimal group size is unstable. Anim Behav 31: 947–948.

Sibly RM, Barker D, Denham MC, Hone J, Page M (2005) On the regulation of populations of mammals, birds, fish, and insects. Science 309: 607–610.

Siegfried WR, Underhill LG (1975) Flocking as an anti-predator strategy in doves. Anim Behav 23: 504–508.

Sigmund K (2002) The economics of fairness. Sci Am 286: 82–87.

Sigmund K (2010) The calculus of selfishness. Princeton, NJ: Princeton University Press.

Sigmund K, Hauert C, Nowak MA (2001) Reward and punishment. Proc Natl Acad Sci USA 98: 10757–10762.

Sigmund K, De Silva H, Traulsen A, Hauert C (2010) Social learning promotes institutions for governing the commons. Nature 466: 861–863.

Sikes B, Cottenie K, Klironomos JN (2009) Plant and fungal identity determines pathogen protection of plant roots by arbuscular mycorrhizas. J Ecol 97: 1274–1280.

Silberberg A, Crescimbene L, Addessi E, Anderson JR, Visalberghi E (2009) Does inequity aversion depend on a frustration effect? A test with capuchin monkeys (Cebus apella). Anim Cogn 12: 505–509.

Silk JB (2002) Kin selection in primate groups. Int J Primatol 23: 849–875.

Silk JB (2007) The adaptive value of sociality in mammalian groups. Phil Trans R Soc B 362: 539–559.

Silk JB, Clark-Wheatley CB, Rodman PS, Samuels A (1981a) Differential reproductive success and facultative adjustments of sex ratios among captive female bonnet macaques (Macaca radiata). Anim Behav 29: 1106–1120.

Silk JB, Samuels A, Rodman PS (1981b) The influence of kinship, rank and sex on affiliation and aggression between adult female and immature bonnet macaques (Macaca radiata). Behaviour 78: 111–137.

Silk JB, Alberts SC, Altmann J (2003) Social bonds of female baboons enhance infant survival. Science 302: 1231–1234.

Silk JB, Beehner JC, Bergman TJ, Crockford C, Engh AL, Moscovice LR, et al. (2010) Strong and consistent social bonds enhance the longevity of female baboons. Curr Biol 20: 1359–1361.

Silk JB, House BR (2011) Evolutionary foundations of human prosocial sentiments. Proc Natl Acad Sci USA 108(Suppl. 2): 10910–10917.

Silva J, Nichols J, Theunissen TW, Guo G, van Oosten AL, Barrandon O, et al. (2009) Nanog is the gateway to the pluripotent ground state. Cell 138: 722–737.

Silverman J, Liang D (2001) Colony disassociation following diet partitioning in a unicolonial ant. Naturwissenschaften 88: 73–77.

Simms EL, Taylor DL (2002) Partner choice in nitrogen-fixation mutualisms of legumes and rhizobia. Integr Comp Biol 42: 369–380.

Simms EL, Taylor DL, Povich J, Shefferson RP, Sachs JL, Urbina M, Tausczik Y (2006) An empirical test of partner choice mechanisms in a wild legume–rhizobium interaction. Proc R Soc B Biol Sci 273: 77–81.

Simons AM (2009) Fluctuating natural selection accounts for the evolution of diversification bet hedging. Proc R Soc B Biol Sci 276: 1987–1992.

Simpson C (2012) The evolutionary history of division of labour. Proc Biol Sci 279: 116–121.

Simpson P (1979) Parameters of cell competition in the compartments of the wing disc of Drosophila. Dev Biol 69: 182–193.

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, Lively CM (1996) The rock-paper-scissors game and the evolution of alternative male strategies. Nature 380: 240–243.

Sinervo B, Calsbeek R (2006) The developmental, physiological, neural and genetic causes and consequences of frequency-dependent selection in the wild. Annu Rev Ecol Evol Syst 37: 581–610.

Sinervo B, Chaine A, Clobert J, Calsbeek R, Hazard L, Lancaster L, McAdam AG, Alonzo S, Corrigan G, Hochberg ME (2006) Self-recognition, color signals and cycles of greenbeard mutualism and altruism. Proc Natl Acad Sci USA 103: 7372–7377.

Sinervo B, Clobert J, Miles DB, McAdam A, Lancaster LT (2008) The role of pleiotropy vs signaller–receiver gene epistasis in life history trade-offs: dissecting the genomic architecture of organismal design in social systems. Heredity 101: 197–211.

Singer T, Lamm C (2009) The social neuroscience of empathy. Ann NY Acad Sci 1156: 81–96.

Singer TL, Espelie KE (1992) Social wasps use nest paper hydrocarbons for nestmate recognition. Anim Behav 44: 63–68.

Singh PB, Brown RE, Roser B (1987) MHC antigens in urine as olfactory recognition cues. Nature 327: 161–164.

Singh PK, Parsek MR, Greenberg EP, Welsh MJ (2002) A component of innate immunity prevents bacterial biofilm development. Nature 417: 552–555.

Sirviö A, Gadau J, Rueppell O, Lamatsch D, Boomsma JJ, et al. (2006) High recombination frequency creates genotypic diversity in colonies of the leaf-cutting ant Acromyrmex echinatior. J Evol Biol 19: 1475–1485.

Sirviö A, Johnston JS, Wenseleers T, Pamilo P (2011) A high recombination rate in eusocial Hymenoptera: evidence from the common wasp Vespula vulgaris. BMC Genet 12: 95.

Sjöblom T, Jones S, Wood LD, Parsons DW, Lin J, et al. (2006) The consensus coding sequences of human breast and colorectal cancers. Science 314: 268–274.

Skaar EP (2010) The battle for iron between bacterial pathogens and their vertebrate hosts. PLoS Pathog 6: e1000949.

Skår J, Coveney PV (2003) Developmental and physiological circuits: dissecting complexity: a report on a talk given by Dr Leroy Hood. Phil Trans R Soc Lond A 361: 1313–1317.

Skubic E, Taborsky M, McNamara JM, Houston AI (2004) When to parasitize? A dynamic optimization model of reproductive strategies in a cooperative breeder. J Theor Biol 227: 487–501.

Skyrms B (2004) The stag hunt and the evolution of social structure. Cambridge, UK: Cambridge University Press.

Skyrms B, Pemantle R (2000) A dynamic model of social network formation. Proc Natl Acad Sci USA 97: 9340–9346.

Slade AJ, Hutchings MJ (1987) Clonal integration and plasticity in foraging behaviour in Glechoma hederacea. J Ecol 75: 1023–1036.

Slatkin M (1974) Hedging one’s evolutionary bets. Nature 250: 704–705.

Slatkin M (1981) Estimating levels of gene flow in natural populations. Genetics 99: 323–335.

Sledge M, Boscaro TS (2001) Cuticular hydrocarbons and reproductive status in the social wasp Polistes dominulus. Behav Ecol Sociobiol 49: 401–409.

Sledge MF, Dani FR, Cervo R, Dapporto L, Turillazzi S (2001) Recognition of social parasites as nest-mates: adoption of colony-specific host cuticular odours by the paper wasp parasite Polistes sulcifer. Proc R Soc B 268: 2253–2260.

Sledge MF, Trinca I, Massolo A, Boscaro F, Turillazzi S (2004) Variation in cuticular hydrocarbon signatures, hormonal correlates and establishment of reproductive dominance in a polistine wasp. J Insect Physiol 50: 73–83.

Slev PR, Nelson AC, Potts WK (2006) Sensory neurons with MHC-like peptide binding properties: disease consequences. Current Opin Immunol 18: 608–616.

Slobodchikoff CN (1984) Resources and the evolution of social behavior. In: Price PW, Slobodchikoff CN, Gaud WS, eds. A new ecology: novel approaches to interactive systems. New York, NY: Wiley. pp 227–251.

Slobodchikoff CN, ed. (1988) The ecology of social behavior. San Diego, CA: Academic Press.

Slobodchikoff CN, Schulz WC (1988) Cooperation, aggression, and the evolution of social behavior. In: Slobodchikoff CN, ed. The ecology of social behavior. San Diego, CA: Academic Press. pp 13–32.

Slobodchikoff CN, Shields WM (1988) Ecological trade-offs and social behavior. In: Slobodchikoff CN, ed. The ecology of social behavior. San Diego, CA: Academic Press. pp 3–10.

Slos S, Stoks R (2008) Predation risk induces stress proteins and reduces antioxidant defense. Funct Ecol 22: 637–642.

Smaldino PE, Schank JC, McElreath R (2013a) Increased costs of cooperation help cooperators in the long run. Am Nat 181: 451–463.

Smaldino PE, Newson L, Schank JC, Richerson PJ (2013b) Simulating the evolution of the human family: cooperative breeding increases in harsh environments. PloS ONE 8: e80753.

Smale L, Holekamp KE, White PA (1999) Siblicide revisited in the spotted hyaena: does it conform to obligate or facultative models? Anim Behav 58: 545–551.

Smallegange IM, Tregenza T (2008) Local competition between foraging relatives: growth and survival of bruchid beetle larvae. J Insect Behav 21: 375–386.

Smallwood PD (1996) An introduction to risk sensitivity: the use of Jensen’s inequality to clarify evolutionary arguments of adaptation and constraint. Am Zool 36: 392–401.

Smead R (2014) The role of social interaction in the evolution of learning. Br J Philos Sci doi: 10.1093/bjps/axt047.

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: 1182–1190.

Smiseth PT, Moore AJ (2004) Behavioral dynamics between caring males and females in a beetle with facultative biparental care. Behav Ecol 15: 621–628.

Smiseth PT, Kölliker M, Royle NJ (2012) What is parental care? In: Royle NJ, Smiseth PT, Kölliker M, eds. The evolution of parental care. Oxford, UK: Oxford University Press. pp 1–17.

Smith AA, Hölldobler B, Liebig J (2009) Cuticular hydrocarbons reliably identify cheaters and allow enforcement of altruism in a social insect. Curr Biol 19: 78–81.

Smith AR, Wcislo WT, O’Donnell S (2003) Assured fitness returns favor sociality in a mass provisioning sweat bee, Megalopta genalis (Hymenoptera: Halictinae). Behav Ecol Sociobiol 54: 14–21.

Smith AR, Wcislo WT, O’Donnell S (2007) Survival and productivity benefits to social nesting in the sweat bee Megalopta genalis (Hymenoptera: Halictidae). Behav Ecol Sociobiol 61: 1111–1120.

Smith AS, Agmo A, Birnie AK, French JA (2009) Manipulation of the oxytocin system alters social behavior and attraction in pair-bonding primates, Callithrix penicillata. Horm Behav 57: 255–262.

Smith AS, Wang Z (2012) Salubrious effects of oxytocin on social stress-induced deficits. Horm Behav 61: 320–330.

Smith C, Reay P (1991) Cannibalism in teleost fish. Rev Fish Biol Fisheries 1: 41–64.

Smith CR, Toth AL, Suarez AV, Robinson GE (2008) Genetic and genomic analyses of the division of labour in insect societies. Nat Rev Genet 9: 735–748.

Smith CR, Mutti NS, Jasper WC, Naidu A, Smith CD, Gadau J (2012) Patterns of DNA methylation in development, division of labor and hybridization in an ant with genetic caste determination. PLoS ONE 7: e42433.

Smith DC (1987) Adult recruitment in chorus frogs: effects of size and date at metamorphosis. Ecology 68: 344–350.

Smith DC, Douglas AE (1987) The biology of symbiosis. London, UK: Edward Arnold Ltd.

Smith FA, Smith SE, Callow JA (1996) Mutualism and parasitism: Diversity in function and structure in the arbuscular (VA) mycorrhizal symbiosis. In: Callow JA, ed. Advances in botanical research. Birmingham, UK: Academic Press. pp 1–43.

smith j, Van Dyken JD, Zee PC (2010) A generalization of Hamilton’s rule for the evolution of microbial cooperation. Science 328: 1700–1703.

Smith JE (2014) Hamilton’s legacy: kinship, cooperation and social tolerance in mammalian groups. Anim Behav 92: 291–304.

Smith JE, Memenis SK, Holekamp KE (2007) Rank-related partner choice in the fission–fusion society of spotted hyenas (Crocuta crocuta). Behav Ecol Sociobiol 61: 753–765.

Smith JE, Van Horn RC, Powning KS, Cole AR, Graham KE, Memenis SK, et al. (2010) Evolutionary forces favoring intragroup coalitions among spotted hyenas and other animals. Behav Ecol 21: 284–303.

Smith JE, Chung LK, Blumstein DT (2013) Ontogeny and symmetry of social partner choice among free-living yellow-bellied marmots. Anim Behav 85: 715–725.

Smith SE, Smith FA, Jakobsen I (2004) Functional diversity in arbuscular mycorrhizal (AM) symbioses: the contribution of the mycorrhizal P uptake pathway is not correlated with mycorrhizal responses in growth or total P uptake. New Phytol 162: 511–524.

Smith SE, Smith FA (2012) Fresh perspectives on the roles of arbuscular mycorrhizal fungi in plant nutrition and growth. Mycologia 104: 1–13.

Smith SM, Beattie AJ, Kent DS, Stow AJ (2009) Ploidy of the eusocial beetle Austroplatypus incompertus (Schedl)(Coleoptera, Curculionidae) and implications for the evolution of eusociality. Insect Soc 56: 285–288.

Smukalla S, Caldara M, Pochet N, Beauvais A, Guadagnini S, Yan C, et al. (2008) FLO1 is a variable green beard gene that drives biofilm-like cooperation in budding yeast. Cell 135: 726–737.

Snijder B, Pelkmans L (2011)Origins of regulated cell-to-cell variability. Nat Rev Mol Cell Biol 12: 119–126.

Snyder LE (1993) Non-random behavioural interactions among genetic subgroups in a polygynous ant. Anim Behav 46: 431–439.

So PM, Dugeon D (1989) Life-history responses of larviparous Boettcherisca formosensis (Diptera: Sarcophagidae) to larval competition for food, including comparisons with oviparous Hemipyrellia ligurriens (Calliphoridae). Ecol Entomol 14: 349–356.

Soares MC, Bshary R, Mendonca R, Grutter AS, Oliveira RF (2012) Arginine vasotocin regulation of interspecific cooperative behaviour in a cleaner fish. PLoS ONE 7: e39583.

Soares MC, Cardoso SC, Grutter AS, Oliveira RF, Bshary R (2014) Cortisol mediates cleaner wrasse switch from cooperation to cheating and tactical deception. Horm Behav 66: 346–350.

Sober E (2001) The two faces of fitness. In: Singh R, Paul D, Crimbas C, Beatty J, eds. Thinking about Evolution: Historical, Philosophical and Political Perspectives. 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.

Sobolewski A, Weeks G (1988) The requirement for DIF for prestalk and stalk cell formation in Dictyostelium discoideum: a comparison of in vivo and in vitro differentiation conditions. Dev Biol 127: 296–303.

Söderbom F, Loomis WF (1998) Cell-cell signaling during Dictyostelium development. Trends Microbiol 6: 402–406.

Søgaard-Andersen L, Yang Z (2008) Programmed cell death: role for MazF and MrpC in Myxococcus multicellular development. Curr Biol 18: R337–339.

Sokolowski MB (2001) Drosophila: genetics meets behaviour. Nat Rev Genet 2: 879–890.

Solís CR, Hughes CR, Klingler CJ, Strassmann JE, Queller DC (1998) Lack of kin discrimination during wasp colony fission. Behav Ecol 9: 172–176.

Solomon NG (2003) A reexamination of factors influencing philopatry in rodents. J Mammal 84: 1182–1197.

Solís CR, Hughes CR, Klingler CJ, Strassmann JE, Queller DC (1998) Lack of kin discrimination during wasp colony fission. Behav Ecol 9: 172–176.

Solomon NG, French JA (1997) Cooperative breeding in mammals. Cambridge, UK: Cambridge University Press.

Solomon NG, Brant CL, Callahan PA, Steinly Jr BA (2001) Mechanisms of reproductive suppression in female pine voles (Microtus pinetorum). Reproduction 122: 297–304.

Sommeijer MJ, Van Veen JW (1990) The polygyny of Myrmica rubra: selective oophagy and trophallaxis as mechanisms of reproductive dominance. Entomol Exp Appl 56: 229–239.

Sommer K, Hölldobler B, Rembold H (1993) Behavioral and physiological aspects of reproductive control in a Diacamma species from Malaysia (Formicidae, Ponerinae). Ethology 94: 162–170.

Sommer K, Hölldobler B (1995) Colony founding by queen association and determinants of reduction in queen number in the ant Lasius niger. Anim Behav 50: 287–294.

Sørensen JG, Kristensen TN, Loeschcke V (2003) The evolutionary and ecological role of heat shock proteins. Ecol Lett 6: 1025–1037.

Soro A, Field J, Bridge C, Cardinal SC, Paxton RJ (2010) Genetic differentiation across the social transition in a socially polymorphic sweat bee, Halictus rubicundus. Mol Ecol 19: 3351–3363.

Soroker V, Vienne C, Hefetz A (1994) The postpharyngeal gland as a “gestalt” organ for nestmate recognition in the ant Cataglyphis niger. Naturwissenschaften 81: 510–513.

Sorvari J, Theodora P, Turillazzi S, Hakkarainen H, Sundström L (2008) Food resources, chemical signaling and nestmate recognition in the ant Formica aquilonia. Behav Ecol 19: 441–447.

Soucy SL (2002) Nesting biology and socially polymorphic behavior of the sweat bee Halictus rubicundus (Hymenoptera: Halictidae). Ann Entomol Soc Amer 95: 57–65.

Soucy SL, Danforth BN (2002) Phylogeography of the socially polymorphic sweat bee Halictus rubicundus (Hymenoptera: Halictidae). Evolution 56: 330–341.

Soucy SL, Giray T, Roubik DW (2003) Solitary and group nesting in the orchid bee Euglossa hyacinthina (Hymenoptera, Apidae). Insect Soc 50: 248–255.

Sparkman AM, Schwartz TS, Madden JA, Boyken SE, Ford NB, Serb JM, Bronikowski AM (2012) Rates of molecular evolution vary in vertebrates for insulin-like growth factor-1 (IGF-1), a pleiotropic locus that regulates life history traits. Gen Comp Endocrinol 178: 164–173.

Spehr M, Kelliher KR, Li XH, Boehm T, Leinders-Zufall T, Zufall F (2006) Essential role of the main olfactory system in social recognition of major histocompatibility complex peptide ligands. J Neurosci 26: 1961–1970.

Spencer H (1864) The principles of biology. Vol. 1. London, UK: Williams and Norgate.

Spencer SL, Gaudet S, Albeck JG, Burke JM, Sorger PK (2009) Non-genetic origins of cell-to-cell variability in TRAIL-induced apoptosis. Nature 459: 428–432.

Spencer SL, Sorger PK (2011) Measuring and modeling apoptosis in single cells. Cell 144: 926–939.

Spieler M (2003) Risk of predation affects aggregation size: a study with tadpoles of Phrynomantis microps (Anura: Microhylidae). Anim Behav 65: 179–184.

Spinks AC, Jarvis JUM, Bennett NC (2000) Comparative patterns of philopatry and dispersal in two common mole-rat populations: implications for the evolution of mole-rat sociality. J Anim Ecol 69: 224–234.

Spiro MH (1974) On the tax incidence of the Pennsylvania Lottery. Natl Tax J 27: 57–61.

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: 1081–1090.

Spong G, Creel S (2004) Effects of kinship on territorial conflicts among groups of lions, Panthera leo. Behav Ecol Sociobiol 55: 325–331.

Spottiswoode CN, Stevens M (2010) Visual modeling shows that avian host parents use multiple visual cues in rejecting parasitic eggs. Proc Natl Acad Sci USA 107: 8672–8676.

Sramkova A, Schulz C, Twele R, Francke W, Ayasse M (2008) Fertility signals in the bumblebee Bombus terrestris (Hymenoptera: Apidae). Naturwissenschaften 95: 515–522.

Stacey PB, Ligon JD (1987) Territory quality and dispersal options in the acorn woodpecker, and a challenge to the habitat-saturation model of cooperative breeding. Am Nat 130: 654–676.

Stacey PB, Koenig WD, eds. (1990) Cooperative breeding in birds. Cambridge, UK: Cambridge University Press.

Stacey PB, Ligon JD (1987) Territory quality and dispersal options in the acorn woodpecker, and a challenge to the habitat-saturation model of cooperative breeding. Am Nat 130: 654–676.

Stacey PB, Ligon JD (1991) The benefits-of-philopatry hypothesis for the evolution of cooperative breeding – variation in territory quality and group-size effects. Am Nat 137: 831–846.

Stacy AR, Diggle SP, Whiteley M (2012) Rules of engagement: defining bacterial communication. Curr Opin Microbiol 15: 155–161.

Stadler ZK, Vijai J, Thom P, Kirchhoff T, Hansen NA, Kauff ND, et al. (2010) Genome-wide association studies of cancer predisposition. Hematol Oncol Clin N Am 24: 973–996.

Stander PE (1991) Foraging dynamics of lions in semi-arid environment. Can J Zool 70: 8–21.

Stanley NR, Britton RA, Grossman AD, Lazazzera B (2003) Identification of catabolite repression as a physiological regulator of biofilm formation by Bacillus subtilis using DNA microarrays. J Bacteriol 185: 1951–1957.

Stanley NR, Lazazzera BA (2004) Environmental signals and regulatory pathways that influence biofilm formation. Mol Microbiol 52: 917–924.

Stanley SM (1998) Macroevolution: pattern and process. Baltimore, MD: The Johns Hopkins University Press.

Starmer C, Sugden R (1989) Probability and juxtaposition effects: an experimental investigation of the common ratio effect. J Risk Uncertainty 2: 159–178.

Starr CK (1984) Sperm competition, kinship, and sociality in the aculeate Hymenoptera. In: Smith RL, ed. Sperm competition and the evolution of animal mating systems. Orlando, FL: Academic Press. pp 427–464.

Starr CK (1985) Enabling mechanisms in the origin of sociality in the hymenoptera: the sting’s the thing. Ann Entomol Soc Am 78: 836–840.

Starr MP (1975) A generalized scheme for classifying organismic associations. Symp Soc Exp Biol 29: 1–20.

Statman M (2002) Lottery players/stock traders. Financ Anal J 58: 14–21.

Stearns CC (1992) The evolution of life histories. Oxford, UK: Oxford University Press.

Stearns FW (2010) One hundred years of pleiotropy: a retrospective. Genetics 186: 767–773.

Stecher B, Hardt WD (2008) The role of microbiota in infectious disease. Trends Microbiol 16: 107–114.

Stein RB, Gossen ER, Jones KE (2005) Neuronal variability: noise or part of the signal? Nat Rev Neurosci 6: 389–397.

Steinberg CEW (2012) Stress ecology: environmental stress as ecological driving force and key player in evolution. Dordrecht, The Netherlands: Springer.

Steinmoen H, Knutsen E, Håvarstein LS (2002) Induction of natural competence in Streptococcus pneumoniae triggers lysis and DNA release from a subfraction of the cell population. Proc Natl Acad Sci USA 99: 7681–7686.

Steinmoen H, Teigen A, Håvarstein LS (2003) Competence-induced cells of Streptococcus pneumoniae lyse competence-deficient cells of the same strain during cocultivation. J Bacteriol 185: 7176–7183.

Stephens DW (1981) The logic of risk-sensitive foraging preferences. Anim Behav 29: 628–629.

Stephens DW (1989) Variance and the value of information. Am Nat 134: 128–140.

Stephens PA, Russell AF, Young AJ, Sutherland WJ, Clutton-Brock TH (2005) Dispersal, eviction, and conflict in meerkats (Suricata suricatta): an evolutionarily stable strategy model. Am Nat 165: 120–135.

Stern DL, Foster WA (1996) The evolution of soldiers in aphids. Biol Rev 71: 27–79.

Sternberg PW, Horvitz HR (1986) Pattern formation during vulval development in C. elegans. Cell 44: 761–772.

Stevens JR, Hauser MD (2004) Why be nice? Psychological constraints on the evolution of cooperation. Trends Cogn Sci 8: 60–65.

Stevens JR, Gilby IC (2004) A conceptual framework for nonkin food sharing: timing and currency of benefits. Anim Behav 67: 603–614.

Stevens JR, Hallinan EV, Hauser MD (2005) The ecology and evolution of patience in two New World monkeys. Biol Lett 1: 223–226.

Stevens JR, Cushman FA, Hauser MD (2005) Evolving the psychological mechanisms for cooperation. Annu Rev Ecol Evol Syst 36: 499–518.

Stevens L, Mertz DB (1985) Genetic stability of cannibalism in Tribolium confusum. Behav Genet 15: 549–559.

Stevens L, Goodnight CJ, Kalisz S (1995) Multilevel selection in natural populations of Impatiens capensis. Am Nat 145: 513–526.

Stevenson PA, Dyakonova V, Rillich J, Schildberger K (2005) Octopamine and experience-dependent modulation of aggression in crickets. J Neurosci 25: 1431–1441.

Stewart AJ, Plotkin JB (2013) From extortion to generosity, evolution in the Iterated Prisoner’s Dilemma. Proc Natl Acad Sci USA 110: 15348–15353.

Stinson CH (1979) On the selective advantage of fratricide in raptors. Evolution 33: 1219–1225.

Stockhammer KA (1967) Some notes on the biology of the blue sweat bee, Lasioglossum coeruleum. J Kansas Entomol Soc 40: 177–189.

Stockholm D, Benchaouir R, Picot J, Rameau P, Neildez TM, et al. (2007) The origin of phenotypic heterogeneity in a clonal cell population in vitro. PLoS ONE 2: e394.

Stockholm D, Edom-Vovard F, Coutant S, Sanatine P, Yamagata Y, et al. (2010) Bistable cell fate specification as a result of stochastic fluctuations and collective spatial cell behaviour. PLoS ONE 5: e14441.

Stockley P, Bro-Jørgensen J (2011) Female competition and its evolutionary consequences in mammals. Biol Rev 86: 341–366.

Stoddard MC, Stevens M (2011) Avian vision and the evolution of egg color mimicry in the common cuckoo. Evolution 65: 2004–2013.

Stolle E, Wilfert L, Schmid-Hempel R, Schmid-Hempel P, Kube M, Reinhardt R, Moritz RFA (2011) A second generation genetic map of the bumblebee Bombus terrestris (Linnaeus, 1758) reveals slow genome and chromosome evolution in the Apidae. BMC Genomics 12: 48.

Stoner DS, Rinkevich B, Weissman IL (1999) Heritable germ and somatic cell lineage competitions in chimeric colonial protochordates. Proc Natl Acad Sci USA 96: 9148–9153.

Stoodley P, Sauer K, Davies DG, Costerton JW (2002) Biofilms as complex differentiated communities. Annu Rev Microbiol 56: 187–209.

Stookey JM, Gonyou HW (1998) Recognition in swine: recognition through familiarity or genetic relatedness. Appl Anim Behav Sci 55: 291–305.

Stopka P, Macdonald DW (1999) The market effect in the wood mouse, Apodemus sylvaticus: selling information on reproductive status. Ethology 105: 969–982.

Stopka P, Johnson DDP (2012)Host-parasite dynamics lead to mixed cooperative games. Folia Zool 61: 235–240.

Storz G, Hengge-Aronis R, eds. (2000) Bacterial stress responses. Washington, DC: ASM Press.

Stover JP, Kendall BE, Fox GA (2012) Demographic heterogeneity impacts density-dependent population dynamics. Theor Ecol 5: 297–309.

Strassmann JE (1985) Relatedness of workers to brood in the social wasp, Polistes exclamans (Hymenoptera: Vespidae). Z Tierpsychol 69: 141–148.

Strassmann JE (1989) Altruism and relatedness at colony foundation in social insects.Trends Ecol Evol 4: 371–374.

Strassmann JE (2001) The rarity of multiple mating by females in the social Hymenoptera. Insect Soc 48: 1–13.

Strassmann JE, Meyer DC (1983) Gerontocracy in the social wasp, Polistes exlamans. Anim Behav 31: 431–438.

Strassmann JE, Queller DC, Hughes CR (1988) Predation and the evolution of sociality in the paper wasp, Polistes bellicosus. Ecology 69: 1497–1505.

Strassmann JE, Queller DC (1989) Ecological determinants of social evolution. In: Breed MD, Page RE Jr, eds. The genetics of social evolution. Boulder, CO: Westview Press. pp 81–101.

Strassmann JE, Klingler CJ, Arévalo E, Zacchi F, Husain A, Williams J, et al. (1997) Absence of within-colony kin discrimination in behavioural interactions of swarm-founding wasps. Proc R Soc Lond B 264: 1565–1570.

Strassmann JE, Zhu Y, Queller DC (2000a) Altruism and social cheating in the social amoeba Dictyostelium discoideum. Nature 408: 965–967.

Strassmann JE, Seppä P, Queller DC (2000b) Absence of within-colony kin discrimination: foundresses of the social wasp, Polistes carolina, do not prefer their own larvae. Naturwissenschaften 87: 266–269.

Strassmann JE, Queller DC (2007) Insect societies as divided organisms: The complexities of purpose and cross-purpose. Proc Natl Acad Sci USA 314: 645–647.

Strassmann JE, Page Jr RE, Robinson GE, Seeley TD (2011a) Kin selection and eusociality. Nature 471: E5–E6.

Strassmann JE, Gilbert OM, Queller DC (2011b) Kin discrimination and cooperation in microbes. Annu Rev Microbiol 65: 349–367.

Strassmann JE, Queller DC (2011) Evolution of cooperation and control of cheating in a social microbe. Proc Natl Acad Sci USA 108(Suppl 2): 10855–10862.

Strickland D (1991) Juvenile dispersal in gray jays—dominant brood member expels siblings from natal territory. Can J Zool 69: 2935–2945.

Strier KB (2008) The effects of kin on primate life histories. Annu Rev Anthropol 37: 21–36.

Stuart RJ (1988) Collective cues as a basis for nestmate recognition in polygynous leptothoracine ants. Proc Natl Acad Sci USA 85: 4572–4575.

Stubblefield JW, Charnov EL (1986) Some conceptual issues in the origin of eusociality. Heredity 57: 181–187.

Stuefer JF, During HJ, de Kroon H (1994) High benefits of clonal integration in two stoloniferous species, in response to heterogeneous light environments. J Ecol 82: 511–518.

Sturgis SJ, Gordon DM (2012) Nestmate recognition in ants (Hymenoptera: Formicidae): a review. Myrmecol News 16: 101–110.

Suarez AV, Holway DA, Liang D, Tsutsui ND, Case TJ (2002) Spatiotemporal patterns of intraspecific aggression in the invasive Argentine ant. Anim Behav 64: 697–708.

Subramanian S, Kumar S (2004) Gene expression intensity shapes evolutionary rates of the proteins encoded by the vertebrate genome. Genetics 168: 373–381.

Suci PA, Mittelman MW, Yu FP, Geesey GG (1994) Investigation of ciprofloxacin penetration into Pseudomonas aeruginosa biofilms. Antimicrob Agents Chemother 38: 2125–2133.

Sugawara M, Okazaki S, Nukui N, Ezura H, Mitsui H, Minamisawa K (2006) Rhizobitoxine modulates plant–microbe interactions by ethylene inhibition. Biotechnol Adv 24: 382–388.

Suits DB (1977) Gambling taxes: Regressivity and revenue potential. Natl Tax J 30: 19–35.

Summers K, McKeon S, Sellars J, Keusenkothen M, Morris J, Gloeckner D, et al. (2003) Parasitic exploitation as an engine of diversity. Biol Rev 78: 639–675.

Summers K, Crespi B (2005) Cadherins in maternal-foetal interactions: red queen with a green beard? Proc R Soc Lond B 272: 643–649.

Sumner S, Casiraghi M, Foster W, Field J (2002) High reproductive skew in tropical hover wasps. Proc R Soc Lond B 269: 179–86.

Sumner S, Hughes WOH, Pedersen JS, Boomsma JJ (2004) Ant parasite queens revert to mating singly. Nature 428: 35–36.

Sumner S, Lucas E, Barker J, Isaac N (2007) Radio-tagging technology reveals extreme nest-drifting behavior in a eusocial insect. Curr Biol 17: 140–145.

Sumner S, Kelstrup H, Fanelli D (2010) Reproductive constraints, direct fitness and indirect fitness benefits explain helping behaviour in the primitively eusocial wasp, Polistes canadensis. Proc R Soc B 277: 1721–1728.

Sumpter D, Beekman M (2003) From nonlinearity to optimality: pheromone trail foraging by ants. Anim Behav 66: 273–280.

Sun P, Smith AS, Lei K, Liu Y, Wang Z (2014) Breaking bonds in male prairie vole: Long-term effects on emotional and social behavior, physiology, and neurochemistry. Behav Brain Res 265: 22–31.

Suntharalingam P, Cvitkovitch DG (2005) Quorum sensing in streptococcal biofilm formation. Trends Microbiol 13: 3–6.

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, 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: 1665–1675.

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: 406–413.

Suzuki S, Akiyama E (2005) Reputation and the evolution of cooperation in sizable groups. Proc R Soc Lond B 272: 1373–1377.

Svanbäck R, Bolnick DI (2005) Intraspecific competition affects the strength of individual specialization: an optimal diet theory model. Evol Ecol Res 7: 993–1012.

Svanbäck R, Bolnick DI (2007) Intraspecific competition promotes resource use diversity within a natural population. Proc R Soc B Biol Sci 274: 839–844.

Svensson O, Magnhagen C, Forsgren E, Kvarnemo C (1998) Parental behaviour in relation to the occurrence of sneaking in the common goby. Anim Behav 56: 175–179.

Swarup S, Harbison ST, Hahn LE, Morozova TV, Yamamoto A, Mackay TFC, Anholt RRH (2012) Extensive epistasis for olfactory behavior, sleep and waking activity in Drosophila melanogaster. Genet Res 94: 9–20.

Swarup S, Huang W, Mackay TFC, Anholt RRH (2013) Analysis of natural variation reveals neurogenetic networks for Drosophila olfactory behavior. Proc Natl Acad Sci USA 110: 1017–1022.

Swenson W, Wilson DS, Elias R (2000) Artificial ecosystem selection. Proc Natl Acad Sci USA 97: 9110–9114.

Szabó G, Hauert C (2002) Evolutionary prisoner’s dilemma games with voluntary participation. Phys Rev E 66: 062903.

Szabó G, Vukov J (2004) Cooperation for volunteering and partially random partnerships. Phys Rev E 69: 036107.

Szabó G, Vukov J, Szolnoki A (2005) Phase diagrams for an evolutionary prisoner’s dilemma game on two-dimensional lattices. Phys Rev E 72: 047107.

Szabó G, Fáth G (2007) Evolutionary games on graphs. Phys Rep 446: 97–216.

Szathmáry E, Maynard Smith J (1995) The major evolutionary transitions. Nature 374: 227–231.

Székely T, Thomas GH, Cuthill IC (2006) Sexual conflict, ecology, and breeding systems in shorebirds. BioScience 56: 801–808.

Szolnoki A, Perc M (2008) Coevolution of teaching activity promotes cooperation. New J Phys 10: 043036.

Szolnoki A, Perc M, Szabó G (2008) Diversity of reproduction rate supports cooperation in the prisoner's dilemma game on complex networks. Eur Phys J B 61: 505–509.

Szolnoki A, Perc M (2009) Emergence of multilevel selection in the prisoner's dilemma game on coevolving random networks. New J Phys 11: 093033.

Szolnoki A, Perc M (2010) Reward and cooperation in the spatial public goods game. Eur Phys Lett 92: 38003.

Szolnoki A, Antonioni A, Tomassini M, Perc M (2014) Binary birth-death dynamics and the expansion of cooperation by means of self-organized growth. Europhys Lett 105: 48001.

Sztajer H, Szafranski SP, Tomasch J, Reck M, Nimtz M, Rohde M, Wagner-Döbler I (2014) Cross-feeding and interkingdom communication in dual-species biofilms of Streptococcus mutans and Candida albicans. ISME J 8: 2256–2271.

Taborsky B, Arnold C, Junker J, Tschopp A (2012) The early social environment affects social competence in a cooperative breeder. Anim Behav 83: 1067–1074.

Taborsky M (1985) Breeder–helper conflict in a cichlid fish with broodcare helpers—an experimental analysis. Behaviour 95: 45–75.

Takase H, Nitanai H, Hoshino K, Otani T (2000) Impact of siderophore production on Pseudomonas aeruginosa infections in immunocompromised mice. Infect Immun 68: 1834–1839.

Takayama S, Isogai A (2005) Self-incompatibility in plants. Annu Rev Plant Biol 56: 467–489.

Takayanagi Y, Yoshida M, Bielsky IF, Ross HE, Kawamata M, Onaka T, et al. (2005) Pervasive social deficits, but normal parturition, in oxytocin receptor-deficient mice. Proc Natl Acad Sci USA 102: 16096–16101.

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: 483–495.

Tamori Y, Deng WM (2013) Tissue repair through cell competition and compensatory cellular hypertrophy in postmitotic epithelia. Dev Cell 25: 350–363.

Tan JHW, Bolle F (2007) Team competition and the public goods game. Econ Lett 96: 133–139.

Tan K, Yang M, Wang Z, Radloff SE, Pirk CW (2012) The pheromones of laying workers in two honeybee sister species: Apis cerana and Apis mellifera. J Comp Physiol A 198: 319–323.

Tanaka Y (1996) Sexual selection enhances population extinction in a changing environment. J Theor Biol 180: 197–206.

Tang CL, Wang WX, Wu X, Wang BH (2006) Effects of average degree on cooperation in networked evolutionary game. Eur Phys J B 53: 411–415.

Tang HY, Smith-Caldas MS, Driscoll MV, Salhadar S, Shingleton AW (2011) FOXO regulates organ-specific phenotypic plasticity in Drosophila. PLoS Genet 7: e1002373.

Tang-Martinez Z (2001) The mechanisms of kin discrimination and the evolution of kin recognition in vertebrates: a critical re-evaluation. Behav Process 53: 21–40.

Tanimoto J (2007) Promotion of cooperation by payoff noise in a 2× 2 game. Phys Rev E 76: 041130.

Tanner CJ, Jackson AL (2012) Social structure emerges via the interaction between local ecology and individual behaviour. J Anim Ecol 81: 260–267.

Taper ML, Case TJ (1985) Quantitative genetic models for the coevolution of character displacement. Ecology 66: 355–371.

Taper ML, Case TJ (1992) Models of character displacement and the theoretical robustness of taxon cycles. Evolution 46: 317–333.

Tardiff K (1992) The current state of psychiatry in the treatment of violent patients. Arch Gen Psychiat 49: 493–499.

Tarnita CE, Antal T, Ohtsuki H, Nowak MA (2009a) Evolutionary dynamics in set structured populations. Proc Natl Acad Sci USA 106: 8601–8604.

Tarnita CE, Ohtsuki H, Antal T, Fu F, Nowak MA (2009b) Strategy selection in structured populations. J Theor Biol 259: 570–581.

Tarpy DR (2003) Genetic diversity within honeybee colonies prevents severe infections and promotes colony growth. Proc R Soc Lond Ser B Biol Sci 270: 99–103.

Tarpy DR, Gilley DC, Seeley TD (2004) Levels of selection in a social insect: a review of conflict and cooperation during honey bee (Apis mellifera) queen replacement. Behav Ecol Sociobiol 55: 513–523.

Tarpy DR, Seeley TD (2006) Lower disease infections in honeybee (Apis mellifera) colonies headed by polyandrous vs. monandrous queens. Naturwissenschaften 93: 195–199.

Tatar M, Kopelman A, Epstein D, Tu MP, Yin CM, Garofalo RS (2001) A mutant Drosophila insulin receptor homolog that extends life-span and impairs neuroendocrine function. Science 292: 107–110.

Tatsuta T, Langer T (2008) Quality control of mitochondria: protection against neurodegeneration and ageing. EMBO J 27: 306–314.

Taylor DL, Bruns TD (1997) Independent, specialized invasions of ectomycorrhizal mutualism by two nonphotosynthetic orchids. Proc Natl Acad Sci USA 94: 4510–4515.

Taylor EB, Rutter J (2011) Mitochondrial quality control by the ubiquitin-proteasome system. Biochem Soc Trans 39: 1509–1513.

Taylor MJ, Bandi C, Hoerauf A (2005) Wolbachia bacterial endosymbionts of filarial nematodes. Adv Parasitol 60: 245–284.

Taylor PD (1992a) Altruism in viscous populations—an inclusive fitness model. Evol Ecol 6: 352–356.

Taylor PD (1992b) Inclusive fitness in a homogeneous environment. Proc R Soc Lond B 249: 299–302.

Taylor PD, Day T, Wild G (2007) Evolution of cooperation in a finite homogeneous graph. Nature 447: 469–472.

Taylor PK (2014) Involvement of regulatory non-coding RNA in motility, biofilm formation, and adaptive antibiotic resistance in Pseudomonas aeruginosa. Thesis. Vancouver, Canada: University of British Columbia.

Taylor SE, Klein LC, Lewis BP, Gruenewald TL, Gurung RA, Updegraff JA (2000) Biobehavioral responses to stress in females: tend-and-befriend, not fight-or-flight. Psychol Rev 107: 411–429.

Tebbich S, Teschke I (2014) Coping with uncertainty: Woodpecker finches (Cactospiza pallida) from an unpredictable habitat are more flexible than birds from a stable habitat. PLoS ONE 9: e91718.

Temeles EJ (1994) The role of neighbors in territorial systems – when are they dear enemies. Anim Behav 47: 339–350.

Terborgh JW (1983) Five New World primates: a study in comparative ecology. Princeton, NJ: Princeton University Press.

Terborgh J, Janson CH (1986) The socioecology of primate groups. Annu Rev Ecol Syst 17: 111–135.

Thébault E, Fontaine C (2010) Stability of ecological communities and the architecture of mutualistic and trophic networks. Science 329: 853–856.

The Honeybee Genome Sequencing Consortium (2006) Insights into social insects from the genome of the honeybee Apis mellifera. Nature 443: 931–949.

Theise ND, Harris R (2006) Postmodern biology: (Adult)( stem) cells are plastic, stochastic, complex, and uncertain. Handbook of Experimental Pharmacology 174: 389–408.

Theraulaz G, Bonabeau E, Deneubourg JL (1998) The origin of nest complexity in social in-sects. Complexity 3: 15–25.

Theraulaz G, Gautrais J, Camazine S, Deneubourg JL (2003) The formation of spatial patterns in social insects: from simple behaviours to complex structures. Phil Trans R Soc London Ser A 361: 1263–1282.

Thiery JP, Chopin D (1999) Epithelial cell plasticity in development and tumor progression. Cancer Metastasis Rev 18: 31–42.

Thoday JM (1953) Components of fitness. Symp Soc Exp Biol 7: 97–113.

Thoits PA (2011) Mechanisms linking social ties and support to physical and mental health. J Health Soc Behav 52: 145–161.

Thomaidou D, Mione MC, Cavanagh JFR, Parnavelas JG (1997) Apoptosis and its relation to the cell cycle in the developing cerebral cortex. J Neurosci 17: 1075–1085.

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 L (1975) Symbiosis as an immunologic problem: the immune system and infectious diseases. In: Neter E, Milgrom F, eds. 4th International convocation on immunology, Buffalo. Basel, Switzerland: Karger. pp 2–11.

Thomas LK, Manica A (2003) Filial cannibalism in an assassin bug. Anim Behav 66: 205–210.

Thomas VC, Hiromasa Y, Harms N, Thurlow L, Tomich J, Hancock LE (2009) A fratricidal mechanism is responsible for eDNA release and contributes to biofilm development of Enterococcus faecalis. Mol Microbiol 72: 1022–1036.

Thompson CRL, Kay RR (2000a) The role of DIF-1 signaling in Dictyostelium development. Mol Cell 6: 1509–1514.

Thompson CRL, Kay RR (2000b) Cell-fate choice in Dictyostelium: intrinsic biases modulate sensitivity to DIF signaling. Dev Biol 227: 56–64.

Thompson DJ (1990) The effects of survival and weather on lifetime egg production in a model damselfly. Ecol Entomol 15: 455–462.

Thompson DJ, Hassall C, Lowe CD, Watts PC (2011) Field estimates of reproductive success in a model insect: behavioural surrogates are poor predictors of fitness. Ecol Lett 14: 905–913.

Thompson GJ, Hurd PL, Crespi BJ (2013) Genes underlying altruism. Biol Lett 9: 20130395.

Thompson JN (2005) The geographic mosaic of coevolution. Chicago, IL: University of Chicago Press.

Thompson JN, Cunningham BM (2002) Geographical structure and dynamics of coevolutionary selection. Nature 417: 735–738.

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: 254–260.

Thorne BL (1997) Evolution of eusociality in termites. Annu Rev Ecol Syst 28: 27–54.

Thorne BL, Breisch NL, Muscedere ML (2003) Evolution of eusociality and the soldier caste in termites: influence of accelerated inheritance. Proc Natl Acad Sci USA 100: 12808–12813.

Thornhill R (1976) Sexual selection and nuptial feeding behavior in Bittacus apicalus (Insecta: Mecoptera). Am Nat 110: 529–548.

Thornhill R (1984) Alternative female choice tactics in the scorpionfly Hylobittacus apicalus (Mecoptera) and their implications. Am Zool 24: 367–383.

Thorpe W (1974) Animal nature and human nature. London, UK: Methuen.

Thumm G, Götz F (1997) Studies on prolysostaphin processing and characterization of the lysostaphin immunity factor (Lif) of Staphylococcus simulans biovar staphylolyticus. Mol Microbiol 23: 1251–1265.

Tibbetts EA, Reeve HK (2003) Benefits of foundress associations in the paper wasp Polistes dominulus: increased productivity and survival but not assurance of fitness returns. Behav Ecol 14: 510–514.

Tibbetts EA, Dale J (2007) Individual recognition: it is good to be different. Trends Ecol Evol 22: 529–537.

Tibbetts EA, Izzo AS (2009) Endocrine mediated phenotypic plasticity: condition-dependent effects of juvenile hormone on dominance and fertility of wasp queens. Horm Behav 56: 527–531.

Tibbetts EA, Huang ZY (2010) The challenge hypothesis in an insect: juvenile hormone increases during reproductive conflict following queen loss in Polistes wasps. Am Nat 176: 123–130.

Tibbetts EA, Izzo A, Huang ZY (2011) Behavioral and physiological factors associated with juvenile hormone in Polistes wasp foundresses. Behav Ecol Sociobiol 65: 1123–1131.

Tibbetts EA, Crocker KC (2014) The challenge hypothesis across taxa: social modulation of hormone titres in vertebrates and insects. Anim Behav 92: 281–290.

Tielbörger K, Kadmon R (2000) Temporal environmental variation tips the balance between facilitation and interference in desert plants. Ecology 81: 1544–1553.

Tilman D (1982) Resource competition and community structure. Princeton, NJ: Princeton University Press.

Timmer M, Cordero MI, Sevelinges Y, Sandi C (2011) Evidence for a role of oxytocin receptors in the long-term establishment of dominance hierarchies. Neuropsychopharmacology 36: 2349–2356.

Tindo M, D’Agostino P, Francescato E, Dejean A, Turillazzi S (1997a) Associative colony foundation in the tropical wasp Belonogaster juncea juncea (Hymenoptera: Vespidae). Insect Soc 44: 365–377.

Tindo M, Turillazzi S, Dejean A (1997b) Behavioral role differentiation in the primitively eusocial wasp Belonogaster juncea juncea (Hymenoptera: Vespidae). J Insect Behav 10: 571–580.

Tindo M, Kenne M, Dejean A (2008) Advantages of multiple foundress colonies in Belonogaster juncea juncea L.: greater survival and increased productivity. Ecol Entomol 33: 293–297.

Tinkle DW, Dunham AE, Congdon JD (1993) Life history and demographic variation in the lizard Sceloporus graciosus: a long-term study. Ecology 74: 2413–2429.

Todrank J, Heth G, Johnston RE (1998) Kin recognition in golden hamsters: evidence for kinship odours. Anim Behav 55: 377–386.

Tomassini M, Pestelacci E, Luthi L (2007) Social dilemmas and cooperation in complex networks. Int J Mod Phys C 18: 1173–1185.

Tomassini M, Pestelacci E (2010) Coordination games on dynamical networks. Games 1: 242–246.

Tonegawa S (1976) Reiteration frequency of immunoglobulin light chain genes: further evidence for somatic generation of antibody diversity. Proc Natl Acad Sci USA 73: 203–207.

Tonegawa S (1983) Somatic generation of antibody diversity. Nature 302: 575–581.

Tonsor SJ (1989) Relatedness and intraspecific competition in Plantago lanceolata. Am Nat 134: 897–906.

Torreblanca M, Meseguer I, Ventosa A (1994) Production of halocin is a practically universal feature of archael halophilic rods. Lett Appl Microbiol 19: 201–205.

Torres CW, Brandt M, Tsutsui ND (2007) The role of cuticular hydrocarbons as chemical cues for nestmate recognition in the invasive Argentine ant (Linepithema humile). Insect Soc 54: 363–373.

Toth AL, Varala K, Newman TC, Miguez FE, Hutchison SK, Willoughby DA, et al. (2007) Wasp gene expression supports an evolutionary link between maternal behavior and eusociality. Science 318: 441–444.

Tóth E, Queller DC, Imperatriz-Fonseca VL, Strassmann JE (2002) Genetic and behavioral conflict over male production between workers and queens in the stingless bee Paratrigona subnuda. Behav Ecol Sociobiol 53: 1–8.

Tóth E, Queller DC, Dollin A, Strassmann JE (2004) Conflict over male parentage in stingless bees. Insect Soc 51: 1–11.

Toyama M (2003) Relationship between reproductive resource allocation and resource capacity in the matriphagous spider, Chiracanthium japonicum (Araneae: Clubionidae). J Ethol 21: 1–7.

Toyoizumi H (2009) Sample path analysis of contribution and reward in cooperative groups. J Theor Biol 256: 311–314.

Toyoizumi H, Field J (2014) Dynamics of social queues. J Theor Biol 346: 16–22.

Toyooka Y, Shimosato D, Murakami K, Takahashi K, Niwa H (2008) Identification and characterization of subpopulations in undifferentiated ES cell culture. Development 135: 909–918.

Traniello JFA (1989) Foraging strategies of ants. Annu Rev Entomol 34: 191–210.

Tranquillo RT (1990) Models of chemical gradient sensing by cells. In: Alt W, Hoffman G, eds.Lecture Notes in biomathematics. Vol. 89: Biological motion. New York, NY: Springer-Verlag. pp 415–441.

Trappetti C, Potter AJ, Paton AW, Orrioni MR, Paton JC (2011) LuxS mediates iron-dependent biofilm formation, competence, and fratricide in Streptococcus pneumoniae. Infect Immun 79: 4550–4558.

Traulsen A (2010) Mathematics of kin-and group-selection: formally equivalent? Evolution 64: 316–323.

Traulsen A, Nowak MA (2006) Evolution of cooperation by multilevel selection. Proc Natl Acad Sci USA 103: 10952–10955.

Trauth MH, Maslin MA, Deino A, Strecker MR (2005) Late Cenozoic moisture history of East Africa. Science 309: 2051–2053.

Travis JMJ, Dytham C (1998) The evolution of dispersal in a metapopulation: a spatially explicit, individual-based model. Proc R Soc Lond B Biol Sci 265: 17–23.

Travisano M, Velicer GJ (2004) Strategies of microbial cheater control. Trends Microbiol 12: 72–78.

Trefilov A, Berard J, Krawczak M, Schmidtke J (2000) Natal dispersal in rhesus macaques is related to serotonin transporter gene promoter variation. Behav Genet 30: 295–301.

Tregenza T, Wedell N (2000) Genetic compatibility, mate choice and patterns of parentage: invited review. Mol Ecol 9: 1013–1027.

Treherne JE, Foster WA (1982) Group-size and anti-predator strategies in a marine insect. Anim Behav 30: 536–542.

Trench RK (1997) Diversity of symbiotic dinoflagellates and the evolution of microbial-invertebrate symbioses. Proceedings of the 8th International Coral Reef Symposium 2: 1275–1286.

Trillmich F, Wolf JBW (2008) Parent-offspring and sibling conflict in Galápagos fur seals and sea lions. Behav Ecol Sociobiol 62: 363–375.

Trivers RL (1971) The evolution of reciprocal altruism. Q Rev Biol 46: 35–57.

Trivers RL (1972) Parental investment and sexual selection. In: Campbell B, ed. Sexual selection and the descent of man. Chicago, IL: Aldine. pp 136–179.

Trivers RL (1974) Parent-offspring conflict. Am Zool 14: 249–264.

Trivers RL (1985) Social evolution. Menlo Park, CA: Benjamin/Cummings Publishing Co.

Trivers R (2004) Mutual benefits at all levels of life. Science 304: 964–965.

Trivers RL, Hare H (1976) Haplodiploidy and the evolution of social insects. Science 191: 249–263.

Trosko JE, Ruch RJ (1998) Cell-cell communication in carcinogenesis. Front Biosci 3: D208–236.

True JR, Mercer JM, Laurie CC (1996) Differences in crossover frequency and distribution among three sibling species of Drosophila. Genetics 142: 507–523.

Trumbo ST (2002) Hormonal regulation of parental care in insects. In: Pfaff D, ed. Hormones, brain and behavior. Vol. 3. San Diego, CA: AcademicPress. pp 115–139.

Tschinkel WR (1992a) Brood raiding and the population-dynamics of founding and incipient colonies of the fire ant, Solenopsis invicta. Ecol Entomol 17: 179–188.

Tschinkel WR (1992b) Brood raiding in the fire ant, Solenopsis invicta (Hymenoptera: Formicidae): laboratory and field observations. Ann Entomol Soc Am 85: 638–646.

Tschinkel WR (2006) The fire ants. Cambridge, MA: Harvard University Press.

Tschinkel WR, Howard DF (1983) Colony founding by pleometrosis in the fire ant, Solenopsis invicta. Behav Ecol Sociobiol 12: 103–113.

Tsuchida K, Saigo T, Nagata N, Tsujita S, Takeuchi K, Miyano S (2003) Queen-worker conflicts over male production and sex allocation in a primitively eusocial wasp. Evolution 57: 2365–2373.

Tsutsui ND (2004) Scents of self: the expression component of self/nonself recognition systems. Ann Zool Fenn 41: 713–727.

Tu MP, Yin CM, Tatar M (2002) Impaired ovarian ecdysone synthesis of Drosophila melanogaster insulin receptor mutants. Aging Cell 1: 158–160.

Tu MP, Yin CM, Tatar M (2005) Mutations in insulin signaling pathway alter juvenile hormone synthesis in Drosophila melanogaster. Gen Comp Endocrinol 142: 347–356.

Tuomi J, Agrell J, Mappes T (1997) On the evolutionary stability of female infanticide. Behav Ecol Sociobiol 40: 227–233.

Turkington R (1996) Intergenotypic interactions in plant mixtures. Euphytica 92: 105–119.

Turner PE, Chao L (1999) Prisoner's dilemma in an RNA virus. Nature 398: 441–443.

Twig G, Hyde B, Shirihai OS (2008) Mitochondrial fusion, fission and autophagy as a quality control axis: the bioenergetic view. Biochim Biophys Acta 1777: 1092–1097.

Twomey E, Morales V, Summers K (2008) The effect of kinship on intraspecific competition in larvae of the poison frog Ameerega bassleri (Anura, Dendrobatidae). Phyllomedusa 7: 121–126.

Tyler WA (1995) The adaptive significance of colonial nesting in a coral-reef fish. Anim Behav 49: 949–966.

Uecker H, Hermisson J (2011) On the fixation process of a beneficial mutation in a variable environment. Genetics 188: 915–930.

Uetz GW (1996) Risk sensitivity and the paradox of colonial web-building in spiders. Am Zool 36: 459–470.

Uetz GW, Hieber CS (1997) Colonial web-building spiders: balancing the costs and benefits of group-living. In: Choe JC, Crespi BJ, eds. The evolution of social behavior in insects and arachnids. Cambridge, UK: Cambridge University Press. pp 458–475.

Uitdehaag J (2011) Bet hedging based cooperation can limit kin selection and form a basis for mutualism. J Theor Biol 280: 76–87.

Underwood RM, Lewis MJ, Hare JF (2004) Reduced worker relatedness does not affect cooperation in honey bee colonies. Can J Zool 82: 1542–1545.

Unger LS (1991) Altruism as a motivation to volunteer. J Econ Psychol 12: 71–100.

Ursino CM, De Mársico MA, Sued M, Farall AS, Reboreda J (2011) Brood parasitism disproportionately increases nest provisioning and helper recruitment in a cooperatively breeding bird. Behav Ecol Sociobiol 65: 2279–2286.

Utne-Palm AC, Hart PJB (2000) The effects of familiarity on competitive interactions between threespined sticklebacks. Oikos 91: 225–232.

Uvnäs-Moberg K (1998) Oxytocin may mediate the benefits of positive social interaction and emotions. Psychoneuroendocrinology, 23: 819–835.

Uyenoyama M (1979) Evolution of altruism under group selection in large and small populations in fluctuating environments. Theor Popul Biol 15: 58–85.

Uyenoyama M, Feldman MW (1980) Theories of kin and group selection: a population genetics perspective. Theor Popul Biol 17: 380–414.

Uyenoyama MK, Feldman MW (1992) Altruism: some theoretical ambiguities. In: Keller EF, Lloyd EA, eds. Keywords in evolutionary biology. Cambridge, MA: Harvard University Press. pp 34–40.

Vabø R, Skaret G (2008) Emerging school structures and collective dynamics in spawning herring: a simulation study. Ecol Model 214: 125–140.

Valdar W, Solberg LC, Gauguier D, Cookson WO, Rawlins JNP, Mott R, Flint J (2006) Genetic and environmental effects on complex traits in mice. Genetics 174: 959–984.

Valsecchi E, Hale P, Corkeron P, Amos W (2002) Social structure in migrating humpback whales (Megaptera novaeangliae). Mol Ecol 11: 507–518.

Vamosi SM (2005) On the role of enemies in divergence and diversification of prey: a review and synthesis. Can J Zool 83: 894–910.

Vamosi SM, Schluter D (2002) Impacts of trout predation on fitness of sympatric sticklebacks and their hybrids. Proc R Soc Lond Ser B Biol Sci 269: 923–930.

van Anders SM, Goldey K, Kuo P (2011) The steroid/peptide theory of social bonds: integrating testosterone and peptide responses for classifying social behavioral contexts. Psychoneuroendocrinology 36: 1265–1275.

van Baalen M, Rand DA (1998) The unit of selection in viscous populations and the evolution of altruism. J Theor Biol 193: 631–648.

van Baalen M, Jansen VAA (2001) Dangerous liaisons: the ecology of private interest and common good. Oikos 95: 211–224.

Vance CP (2001) Symbiotic nitrogen fixation and phosphorus acquisition. Plant nutrition in a world of declining renewable resources. Plant Physiol 127: 390–397.

Vance RE, Isberg RR, Portnoy DA (2009) Patterns of pathogenesis: discrimination of pathogenic and nonpathogenic microbes by the innate immune system. Cell Host Microbe 6: 10–21.

van Damme P, Appelbaum S, Hecht T (1989) Sibling cannibalism in Koi carp, Cypvinus carpio L., larvae and juveniles reared under controlled conditions. J Fish Biol 34: 855–863.

van den Berghe EP, Gross MR (1989) Natural selection resulting from female breeding competition in a pacific salmon (Coho: Oncorhynchus kisutch). Evolution 43: 125–140.

van de Peppel J, Holstege FC (2005) Multifunctional genes. Mol Syst Biol 1: 2005.0003.

van de Pol M, Pen I, Heg D, Weissing FJ (2007) Variation in habitat choice and delayed reproduction: adaptive queuing strategies or individual quality differences? Am Nat 170: 530–541.

Vander Meer RK, Morel L (1998) Nestmate recognition in ants. In: Vander Meer RK, Breed MD, Espelie KE, Winston ML, eds. Pheromone communication in social insects. Boulder, CO: Westview Press. pp 79–103.

Vander Meer RK, Breed MD, Winston ML, Espelie KE, eds. (1998) Pheromone communication in social insects. Boulder, CO: Westview Press.

Vander Meer RK, Alonso LE (2002) Queen primer pheromone affects conspecific fire ant (Solenopsis invicta) aggression. Behav Ecol Sociobiol 51: 122–130.

Vander Meer RK, Preston CA, Hefetz A (2008) Queen regulates biogenic amine level and nestmate recognition in workers of the fire ant, Solenopsis invicta. Naturwissenschaften 95: 1155–1158.

van der Ploeg JR (2005) Regulation of bacteriocin production in Streptococcus mutans by the quorum-sensing system required for development of genetic competence. J Bacteriol 187: 3980–3989.

Vanderpool CK, Gottesman S (2004) Involvement of a novel transcriptional activator and small RNA in post-transcriptional regulation of the glucose phosphoenolpyruvate phosphotransferase system. Mol Microbiol 54: 1076–1089.

van Dijk RE (2009) Sexual conflict over parental care in penduline tits. PhD thesis. Bath, UK: University of Bath.

van Duijn M, Keijzer F, Franken D (2006) Principles of minimal cognition: Casting cognition as sensorimotor coordination. Adaptive Behavior 14: 157–170.

van Doorn GS, Riebli T, Taborsky M (2014) Coaction versus reciprocity in continuous-time models of cooperation. J Theor Biol 356: 1–10.

Van Dyken JD (2010) The components of kin competition. Evolution 64: 2840–2854.

Van Dyken JD, Linksvayer TA, Wade MJ (2011) Kin selection–mutation balance: a model for the origin, maintenance, and consequences of social cheating. Am Nat 177: 288–300.

Van Dyken JD, Wade MJ (2012) Origins of altruism diversity II: Runaway coevolution of altruistic strategies via “reciprocal niche construction”. Evolution 66: 2498–2513.

van Erp AMM, Miczek KA (2000) Aggressive behavior, increased accumbal dopamine, and decreased cortical serotonin in rats. J Neurosci 20: 9320–9325.

Van Hierden YM, Korte SM, Ruesink EW, Van Reenen CG, Engel B, Korte-Bouws GAH, et al. (2002) Adrenocortical reactivity and central serotonin and dopamine turnover in young chicks from a high and low feather-pecking line of laying hens. Physiol Behav 75: 653–659.

van Honk CGJ, Velthuis HHW, Röseler PF, Malotaux ME (1980) The mandibular glands of Bombus terrestris queens as a source of queen pheromones. Entomol Exp Appl 28: 191–198.

van Honk CGJ, Hogeweg P (1981) The ontogeny of the social structure in a captive Bombus terrestris colony. Behav Ecol Sociobiol 9: 111–119.

van Honk CGJ, Röseler PF, Velthuis HHW, Hoogeveen JC (1981) Factors influencing the egg laying of workers in a captive Bombus terrestris colony. Behav Ecol Sociobiol 9: 9–14.

van Honk J, Schutter DJ, Bos PA, Kruijt AW, Lentjes EG, Baron-Cohen S (2011) Testosterone administration impairs cognitive empathy in women depending on second-to-fourth digit ratio. Proc Natl Acad Sci USA 108: 3448–3452.

Van Horn R, Engh A, Scribner K, Funk S, Holekamp K (2004) Behavioural structuring of relatedness in the spotted hyena (Crocuta crocuta) suggests direct fitness benefits of clan-level cooperation. Mol Ecol 13: 449–458.

Van Lange PAM (2008) Does empathy trigger only altruistic motivation? How about selflessness and justice? Emotion 8: 766–774.

van Ooyen A (2001) Competition in the development of nerve connections: a review of models. Network 12: R1–R47.

van Ooyen A, Willshaw DJ (1999) Competition for neurotrophic factor in the development of nerve connections. Proc Biol Sci 266: 883–892.

Van Oystaeyen A, Oliveira RC, Holman L, van Zweden JS, Romero C, Oi CA, et al. (2014) Conserved class of queen pheromones stops social insect workers from reproducing. Science 343: 287–290.

van Schaik CP (1983) Why are diurnal primates living in groups? Behaviour 87: 120–144.

van Schaik CP, van Hooff JARAM (1983) On the ultimate causes of primate social systems. Behaviour 85: 91–117.

van Schaik C, van Noordwijk MA, de Boer RJ, den Tunkelaar I (1983a) The effect of group size on time budgets and social behaviour in wild long-tailed macaques (Macaca fascicularis). Behav Ecol Sociobiol 13: 173–181.

van Schaik C, van Noordwijk M, Warsono B, Sutriono E (1983b) Party size and early detection of predators in sumatran forest primates. Primates 24: 211–221.

Van Segbroeck S, Santos FC, Lenaerts T, Pacheco JM (2009) Reacting differently to adverse ties promotes cooperation in social networks. Phys Rev Lett 102: 058105.

Van Valen L (1973) A new evolutionary law. Evol Theory 1: 1–30.

van Veelen M (2009) Group selection, kin selection, altruism and cooperation: when inclusive fitness is right and when it can be wrong. J Theor Biol 259: 589–600.

van Veelen M, García J, Sabelis MW, Egas M (2010a) Call for a return to rigour in models. Nature 467: 661.

van Veelen M, García J, Avilés L (2010b) It takes grouping and cooperation to get sociality. J Theor Biol 264: 1240–1253.

Van Vuren D, Armitage KB (1994) Survival of dispersing and philopatric yellow-bellied marmots—what is the cost of dispersal. Oikos 69: 179–181.

van Wilgenburg E, Dang S, Forti AL, Koumoudouros TJ, Ly A, Elgar MA (2007) An absence of aggression between non-nestmates in the bull ant Myrmecia nigriceps. Naturwissenschaften 94: 787–790.

van Wilgenburg E, Symonds MRE, Elgar MA (2011) Evolution of cuticular hydrocarbon diversity in ants. J Evol Biol 24: 1188–1198.

van Zweden JS, Dreier S, d’Ettorre P (2009) Disentangling environmental and heritable nestmate recognition cues in a carpenter ant. J Insect Physiol 55: 159–164.

van Zweden JS, Brask JB, Christensen JH, Boomsma JJ, Linksvayer TA, d’Ettorre P (2010) Blending of heritable recognition cues among ant nestmates creates distinct colony gestalt odours but prevents within?colony nepotism. J Evol Biol 23: 1498–1508.

van Zweden JS, d’Ettorre P (2010) Nestmate recognition in social insects and the role of hydrocarbons. In: Blomquist GJ, Bagnères AG, eds. Insect hydrocarbons: biology, biochemistry and chemical ecology. Cambridge, UK: Cambridge University Press. pp 222–243.

Vargo EL (1992) Mutual pheromonal inhibition among queens in polygyne colonies of the fire ant Solenopsis invicta. Behav Ecol Sociobiol 31: 205–210.

Vargo EL (1999) Reproductive development and ontogeny or queen pheromone production in the fire ant Solenopsis invicta. Physiol Entomol 24: 370–376.

Varma A, Chincholkar S, eds. (2007) Microbial Siderophores, vol. 12 of Soil Biology. Berlin, Germany: Springer.

Vasil ML, Ochsner UA (1999) The response of Pseudomonas aeruginosa to iron: genetics, biochemistry and virulence. Mol Microbiol 34: 399–413.

Vaughn RL, Würsig B, Packard J (2010) Dolphin prey herding: Prey ball mobility relative to dolphin group and prey ball sizes, multispecies associates, and feeding duration. Mar Mammal Sci 26: 213–225.

Veening JW, Smits WK, Kuipers OP (2008) Bistability, epigenetics, and bet-hedging in bacteria. Annu Rev Microbiol 62: 193–210.

Vehrencamp SL (1979) The roles of individual, kin and group selection in the evolution of sociality. In: Marler P, Vandenbergh J, eds. Social behavior and communication. New York, NY: Plenum. pp 351–394.

Vehrencamp SL (1983) A model for the evolution of despotic versus egalitarian societies. Anim Behav 31: 667–682.

Vehrencamp SL (1984) Exploitation in co-operative societies: models of fitness biasing in co-operative breeders. In: Barnard CJ, ed. Producers and scroungers: strategies of exploitation and parasitism. London, UK: Croom Helm. pp 229–266.

Vehrencamp SL, Bowen BS, Koford RR (1986) Breeding roles and pairing patterns within communal groups of groove-billed anis. Anim Behav 34: 347–366.

Vehrencamp SL, Koford RR, Bowen BS (1988) The effect of breeding-unit size on fitness components in groove-billed anis. In: Clutton-Brock TH, ed. Reproductive success: studies of individual variation in contrasting breeding systems. Chicago, IL: University of Chicago Press. pp 291–304.

Velicer GJ (2003) Social strife in the microbial world. Trends Microbiol 11: 330–337.

Velicer GJ, Kroos L, Lenski RE (1998) Loss of social behaviors by Myxococcus xanthus during evolution in an unstructured habitat. Proc Natl Acad Sci USA 95: 12376–12380.

Velicer GJ, Kroos L, Lenski RE (2000) Developmental cheating in the social bacterium Myxococcus xanthus. Nature 404: 598–601.

Velicer GJ, Lenski RE, Kroos L (2002) Rescue of social motility lost during evolution of Myxococcus xanthus in an asocial environment. J Bacteriol 184: 2719–2727.

Velicer GJ, Stredwick KL (2002) Experimental social evolution with Myxococcus xanthus. Antonie Leeuwenhoek 81: 155–164.

Velicer GJ, Raddatz G, Keller H, Deiss S, Lanz C, Dinkelacker I, Schuster SC (2006) Comprehensive mutation identification in an evolved bacterial cooperator and its cheating ancestor. Proc Natl Acad Sci USA 103: 8107–8112.

Velicer G, Vos M (2009) Sociobiology of the myxobacteria. Annu Rev Microbiol 63: 599–623.

Velthuis HHW (1970) Ovarian development in Apis mellifera worker bees. Entomol Exp Appl 13: 377–394.

Veltman CJ (1989) Flock, pair and group living lifestyles without cooperative breeding by Australian magpies Gymnorhina tibicen. Ibis 131: 601–608.

Venditti C, Meade A, Pagel M (2010) Phylogenies reveal new interpretation of the Red Queen. Nature 463: 349–352.

Vendruscolo M, Zurdo J, MacPhee CE, Dobson CM (2003) Protein folding and misfolding: a paradigm of self–assembly and regulation in complex biological systems. Phil Trans R Soc Lond A 361: 1205–1222.

Veenema AH (2012) Toward understanding how early-life social experiences alter oxytocin-and vasopressin-regulated social behaviors. Horm Behav 61: 304–312.

Verberk WCEP, Bilton DT (2011) Can oxygen set thermal limits in an insect and drive gigantism? PLoS ONE 6: e22610.

Vergoz V, Schreurs HA, Mercer AR (2007) Queen pheromone blocks aversive learning in young worker bees. Science 317: 384–386.

Verhoeven KJF, Jansen JJ, van Dijk PJ, Biere A (2010a) Stress-induced DNA methylation changes and their heritability in asexual dandelions. New Phytol 185: 1108–1118.

Verhoeven KJF, van Gurp TP (2012) Transgenerational effects of stress exposure on offspring phenotypes in apomictic dandelion. PLoS ONE 7: e38605.

Vermeulen L, Sprick MR, Kemper K, Stassi G, Medema JP (2008) Cancer stem cells--old concepts, new insights. Cell Death Differ 15: 947–958.

Vigilant L, Hofreiter M, Siedel H, Boesch C (2001) Paternity and relatedness in wild chimpanzee communities. Proc Natl Acad Sci USA 98: 12890–12895.

Villinger J, Waldman B (2008) Self-referent MHC type matching in frog tadpoles. Proc R Soc B Biol Sci 275: 1225–1230.

Vineis P (2003) Cancer as an evolutionary process at the cell level: an epidemiological perspective. Carcinogenesis 24: 1–6.

Violle C, Pu Z, Jiang L (2010) Experimental demonstration of the importance of competition under disturbance. Proc Natl Acad Sci USA 107: 12925–12929.

Viricel A, Strand A, Rosel P, Ridoux V, Garcia P (2008) Insights on common dolphin (Delphinus delphis) social organization from genetic analysis of a mass-stranded pod. Behav Ecol Sociobiol 63: 173–185.

Virkkunen M, Rawlings R, Tokola R, Poland RE, Guidotti A, Nemeroff C, et al. (1994) CSF biochemistries, glucose metabolism, and diurnal activity rhythms in alcoholic, violent offenders, fire setters, and healthy participants. Arch Gen Psychiatry 51: 20–27.

Visca P, Leoni L, Wilson MJ, Lamont IL (2002) Iron transport and regulation, cell signalling and genomics: lessons from Escherichia coli and Pseudomonas. Mol Microbiol 45: 1177–1190.

Visscher PK (1989) A quantitative study of worker reproduction in honey bee colonies. Behav Ecol Sociobiol 25: 247–254.

Vitousek PM (1984) Litterfall, nutrient cycling, and nutrient limitation in tropical forests. Ecology 65: 285–298.

Vivarelli S, Wagstaff L, Piddini E (2012) Cell wars: regulation of cell survival and proliferation by cell competition. Essays Biochem 53: 69–82.

Vlamakis H, Aguilar C, Losick R, Kolter R (2008) Control of cell fate by the formation of an architecturally complex bacterial community. Gene Dev 22: 945–953.

Voigt CC, Streich WJ (2003) Queuing for harem access in colonies of the greater sac-winged bat. Anim Behav 65: 149–156.

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 Dawans B, Fischbacher U, Kirschbaum C, Fehr E, Heinrichs M (2012) The social dimension of stress reactivity: acute stress increases prosocial behavior in humans. Psychol Sci 23: 651–660.

von Foerster H (1960) On self-organising systems and their environments. In: Yovitts MC, Cameron S, eds. Self-organising systems. London, UK: Pergamon Press. pp 31–50.

von Neumann J, Morgenstern O (1944) Theory of games and economic behavior. Princeton, NJ: Princeton University Press.

Voogd S (1956) The influence of a queen on the ovary development in worker bees. Experientia 12: 199–201.

Voors MJ, Nillesen EE, Verwimp P, Bulte EH, Lensink R, Van Soest DP (2012) Violent conflict and behavior: a field experiment in Burundi. Am Econ Rev 102: 941–964.

Vos M, Velicer GJ (2008) Isolation by distance in the spore-forming soil bacterium Myxococcus xanthus. Curr Biol 18: 386–391.

Vos M, Velicer GJ (2009) Social conflict in centimeter- and global-scale populations of the bacterium Myxococcus xanthus. Curr Biol 19: 1763–1767.

Voth DE, Ballard JD (2005) Clostridium difficile toxins: mechanism of action and role in disease. Clin Microbiol Rev 18: 247–263.

Vršanský P, Aristov D (2014) Termites (Isoptera) from the Jurassic/Cretaceous boundary: Evidence for the longevity of their earliest genera. Eur J Entomol 111: 137–141.

Vucetich J, Peterson R, Waite T (2004) Raven scavenging favours group foraging in wolves. Anim Behav 67: 1117–1126.

Vucic-Pestic O, Birkhofer K, Rall BC, Scheu S, Brose U (2010) Habitat structure and prey aggregation determine the functional response in a soil predator–prey interaction. Pedobiologia 53: 307–312.

Vukov J, Szabó G, Szolnoki A (2006) Cooperation in the noisy case: Prisoner’s dilemma game on two types of regular random graphs. Phys Rev E 73: 067103.

Vulic M, Kolter R (2001) Evolutionary cheating in Escherichia coli stationary phase cultures. Genetics 158: 519–526.

Wachter B, Höner OP, East ML, Golla W, Hofer H (2002) Low aggression levels and unbiased sex ratios in a prey-rich environment: no evidence of siblicide in Ngorongoro spotted hyenas (Crocuta crocuta). Behav Ecol Sociobiol 52: 348–356.

Waddington CH (1942) Canalization of development and the inheritance of acquired characters. Nature 150: 563–565.

Waddington CH (1953) Genetic assimilation of an acquired character. Evolution 7: 118–126.

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.

Wade MJ (1976) Group selections among laboratory populations of Tribolium. Proc Natl Acad Sci USA 73: 4604–4607.

Wade MJ (1977) An experimental study of group selection. Evolution 31: 134–153.

Wade MJ (1978) Kin selection: a classical approach and a general solution. Proc Natl Acad Sci USA 75: 6154–6158.

Wade MJ (1979) The evolution of social interactions by family selection. Am Nat 113: 399–417.

Wade MJ (1980a) Kin selection: its components. Science 210: 665–667.

Wade MJ (1980b) An experimental study of kin selection. Evolution 34: 844–855.

Wade MJ (1982) Group selection: migration and the differentiation of small populations. Evolution 36: 949–961.

Wade MJ (1985) Soft selection, hard selection, kin selection, and group selection. Am Nat 125: 61–73.

Wade MJ, Wilson DS, Goodnight C, Taylor D, Bar-Yam Y, et al. (2010) Multilevel and kin selection in a connected world. Nature 463: E8–E9. discussion E9–E10.

Wagner D, Tissot M, Cuevas W, Gordon DM (2000) Harvester ants utilize cuticular hydrocarbons in nestmate recognition. J Chem Ecol 26: 2245–2257.

Wagner D, Tissot M, Gordon D (2001) Task-related environment alters the cuticular hydrocarbon composition of harvester ants. J Chem Ecol 27: 1805–1819.

Wagner PL, Livny J, Neely MN, Acheson DW, Friedman DI, Waldor MK (2002) Bacteriophage control of Shiga toxin 1 production and release by Escherichia coli. Mol Microbiol 44: 957–970.

Wahaj SA, Van Horn RC, Van Horn TL, Dreyer R, Hilgris R, Schwarz J, et al. (2004) Kin discrimination in the spotted hyena (Crocuta crocuta): nepotism among siblings. Behav Ecol Sociobiol 56: 237–247.

Wahl LM (2002a) Evolving the division of labor: generalists, specialists and task allocation. J Theor Biol 219: 371–388.

Wahl LM (2002b) The division of labor: genotypic versus phenotypic specialization. Am Nat 160: 135–145.

Wahl LM, Nowak MA (1999) The continuous prisoner’s dilemma: II. Linear reactive strategies with noise. J Theor Biol 200: 323–338.

Waite AJ (2013) Environmental stress can favor cooperation by triggering an adaptive race between cooperators and cheaters. PhD thesis. Seattle, WA: University of Washington.

Waite AJ, Shou W (2012) Adaptation to a new environment allows cooperators to purge cheaters stochastically. Proc Natl Acad Sci USA 109: 19079–19086.

Waldman B (1985) Olfactory basis of kin recognition in toad tadpoles. J Comp Physiol A 156: 565–577.

Waldman B (1987) Mechanisms of kin recognition. J Theor Biol 128: 159–185.

Waldman B (1988) The ecology of kin recognition. Annu Rev Ecol Syst 19: 543–571.

Waldman B, Frumhoff PC, Sherman PW (1988) Problems of kin recognition. Trends Ecol Evol 3: 8–13.

Waldman JD (2007) Thinking systems need systems thinking. Syst Res Behav Sci 24: 271–284.

Walker LR, Chapin FS 3rd (1987) Physiological controls over seedling growth in primary succession on an Alaskan floodplain. Ecology 67: 1508–1523.

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: 5483–5488.

Wallace DC, Fan W (2010) Energetics, epigenetics, mitochondrial genetics. Mitochondrion 10: 12–31.

Walters JR (1990) Red-cockaded woodpeckers: A ‘primitive’ cooperative breeder. In: Stacey PB, Koenig WD, eds. Cooperative breeding in birds: long-term studies of ecology and behavior. Cambridge, UK: Cambridge University Press. pp 67–102.

Walters JR, Seyfarth RM (1987) Conflict and cooperation. In: Smuts BB, Cheney DL, Seyfarth RM, Wrangham RW, Struhsaker TT, eds. Primate societies. Chicago, IL: The University of Chicago Press. pp 306–17.

Walters JR, Doerr PD, Carter JH III (1988) The cooperative breeding system of the red-cockaded woodpecker. Ethology 78: 275–305.

Walters JR, Doerr PD, Carter JH (1992) Delayed dispersal and reproduction as a life-history tactic in cooperative breeders—fitness calculations from red-cockaded woodpeckers. Am Nat 139: 623–643.

Walters JR, Copeyon CK, Carter IJH (1992) Test of the ecological basis of cooperative breeding in red-cockaded woodpeckers. Auk 109: 90–97.

Walum H, Lichtenstein P, Neiderhiser JM, Reiss D, Ganiban JM, Spotts EL, et al. (2012) Variation in the oxytocin receptor gene is associated with pair-bonding and social behavior. Biol Psychiatry 71: 419–426.

Wang J, Suri S, Watts DJ (2012) Cooperation and assortativity with dynamic partner updating. Proc Natl Acad Sci USA 109: 14363–14368.

Wang P, Lei J-P, Li M-H, Yu F-H (2012) Spatial heterogeneity in light supply affects intraspecific competition of a stoloniferous clonal plant. PLoS ONE 7: e39105.

Wang S, Yehya N, Schadt EE, Wang H, Drake TA, et al. (2006) Genetic and genomic analysis of a fat mass trait with complex inheritance reveals marked sex specificity. PLoS Genet. 2: e15.

Wang S, Szalay MS, Zhang C, Csermely P (2008) Learning and innovative elements of strategy adoption rules expand cooperative network topologies. PLoS ONE 3: e1917.

Wang W, Fang H, Groom L, Cheng A, Zhang W, et al. (2008) Superoxide flashes in single mitochondria. Cell 134: 279–290.

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: 1868–1877.

Wang Y, Azevedo SV, Hartfelder K, Amdam GV (2013) Insulin-like peptides (AmILP1 and AmILP2) differentially affect female caste development in the honey bee (Apis mellifera L.). J Exp Biol 216: 4347–4357.

Wang Y, Odemer R, Rosenkranz P, Moussian B (2014) Putative orthologues of genetically identified Drosophila melanogaster chitin producing and organising genes in Apis mellifera. Apidologie 45: 733–747.

Wang Z, Liao BY, Zhang J (2010) Genomic patterns of pleiotropy and the evolution of complexity. Proc Natl Acad Sci USA 107: 18034–18039.

Wang Z, Li Y, During HJ, Li L (2011) Do clonal plants show greater division of labour morphologically and physiologically at higher patch contrasts? PLoS ONE 6: e25401.

Ward AJW, Sumpter DJT, Couzin ID, Hart PJB, Krause J (2008) Quorum decision-making facilitates information transfer in fish schools. Proc Natl Acad Sci USA 105: 6948–6953.

Ward PI (1986) Prey availability increases less quickly than nest size in the social spider Stegodyphus mimosarum. Behaviour 97: 213–225.

Ware JL, Grimaldi DA, Engel MS (2010) The effects of fossil placement and calibration on divergence times and rates: an example from the termites (Insecta: Isoptera). Arthropod Struct Dev 39: 204–219.

Wascher CAF, Bugnyar T (2013) Behavioral responses to inequity in reward distribution and working effort in crows and ravens. PLoS ONE 8: e56885.

Waser PM (1981) Sociality or territorial defence? The influence of resource renewal. Behav Ecol Sociobiol 8: 231–237.

Waser PM (1988) Resources, philopatry, and social interactions among mammals. In: Slobodchikoff CN, ed. The ecology of social behavior. San Diego, CA: Academic Press. pp 109–130.

Waser PM (1996) Patterns and consequences of dispersal in gregarious carnivores. In: Gittleman JL, ed. Carnivore behavior, ecology and evolution. Ithaca, NY: Cornell University Press. pp 267–295.

Wasser SK, Barash DP (1983) Reproductive suppression among female mammals: implications for biomedicine and sexual selection theory. Q Rev Biol 58: 513–538.

Wasser SK, Starling AK (1988) Proximate and ultimate causes of reproductive suppression among female yellow baboons at Mikumi National Park, Tanzania. Am J Primatol 16: 97–121.

Waters CM, Bassler BL (2005) Quorum sensing: Cell-to-cell communication in bacteria. Annu Rev Cell Dev Biol 21: 319–346.

Watnick P, Kolter R (2000) Biofilm, city of microbes. J Bacteriol 182: 2675–2679.

Watson RT, Razafindramanana S, Rafanomezantsoa S (1999) Breeding biology, extra-pair birds, productivity, siblicide and conservation of the Madagascar fish eagle. Ostrich 70: 105–111.

Waxman D, Peck JR (1998) Pleiotropy and the preservation of perfection. Science 279: 1210–1213.

Wcislo WT, Danforth BN (1997) Secondarily solitary: the evolutionary loss of social behavior. Trends Ecol Evol 12: 468–474.

Wcislo WT, Tierney SM (2009) The evolution of communal behavior in bees and wasps: an alternative to eusociality. In: Gadau J, Fewel JH, eds. Organization of insect societies: From genome to sociocomplexity. Cambridge, MA: Harvard University Press. pp 148–169.

Weaver ICG, Cervoni N, Champagne FA, D’Alessio AC, Sharma S, Seckl JR, et al. (2004) Epigenetic programming by maternal behavior. Nat Neurosci 7: 847–854.

Webb JS, Givskov M, Kjelleberg S (2003a) Bacterial biofilms: prokaryotic adventures in multicellularity. Curr Opin Microbiol 6: 578–585.

Webb JS, Thompson LS, James S, Charlton T, Tolker-Nielsen T, Koch B, Givskov M, Kjelleberg S (2003b) Cell death in Pseudomonas aeruginosa biofilm development. J Bacteriol 185: 4585–4592.

Wedekind C, Furi S (1997) Body odour preferences in men and women: do they aim for specific MHC combinations or simply heterozygosity? Proc R Soc London Ser B-Biol Sci 264: 1471–1479.

Wedekind C, Seebeck T, Bettens F, Paepke AJ (1995) MHCdependent mate preferences in humans. Proc Roy Soc Lond Ser B-Biol Sci 260: 245–249.

Wedekind C, Milinski M (2000) Cooperation through image scoring in humans. Science 288: 850–852.

Wedekind C, Penn D (2000) MHC genes, body odours, and odour preferences. Nephrol Dial Transplant 15: 1269–1271.

Wegener J, Bienefeld K (2009) Methods for breeding of honey bee using worker-derived drones. Züchtungskunde 81: 265–278.

Wei H, Håvarstein LS (2012) Fratricide is essential for efficient gene transfer between pneumococci in biofilms. Appl Environ Microbiol 78: 5897–5905.

Weidt A, Hofmann SE, König B (2008) Not only mate choice matters: fitness consequences of social partner choice in female house mice. Anim Behav 75: 801–808.

Weihs D (1973) Hydromechanics of fish schooling. Nature 241: 290–291.

Weil T, Hoffmann K, Kroiss J, Strohm E, Korb J (2009) Scent of a queen—cuticular hydrocarbons specific for female reproductives in lower termites. Naturwissenschaften 96: 315–319.

Weimerskirch H, Martin J, Clerquin Y, Alexandre P, Jiraskova S (2001) Energy saving in flight formation. Nature 413: 697–698.

Weinberg R (1998) One renegade cell: the quest for the origins of cancer. London, UK: Phoenix.

Weiner J (1990) Asymmetric competition in plant populations. Trends Ecol Evol 5: 360–364.

Weinstock M (2008) The long-term behavioural consequences of prenatal stress. Neurosci Biobehav Rev 32: 1073–1086.

Weismann A (1889) Essays upon heredity. London, UK: Oxford University Press.

Weismann A (1893) The all-sufficiency of natural selection. Contemp Rev 64: 309–338, 596–610.

Weiss KM (2006) Commentary: evolution of action in cells and organisms. Int J Epidemiol 35: 1159–1160.

Weiss KM, Buchanan AV (2009) The cooperative genome: organisms as social contracts. Int J Dev Biol 53: 753–763.

Weiss KM, Buchanan AV, Lambert BW (2011) The red queen and her king: cooperation at all levels of life. Am J Phys Anthropol 146(S53): 3–18.

Welden CW, Slauson WL (1986) The intensity of competition versus its importance: an overlooked distinction and some implications. Q Rev Biol 61: 23–44.

Wen ZT, Suntharaligham P, Cvitkovitch DG, Burne RA (2005) Trigger factor in Streptococcus mutans is involved in stress tolerance, competence development, and biofilm formation. Infect Immun 73: 219–225.

Weng GZ, Bhalla US, Iyengar R (1999) Complexity in biological signaling systems. Science 284: 92–96.

Wenseleers T, Hart AG, Ratnieks FLW (2004a) When resistance is useless: Policing and the evolution of reproductive acquiescence in insect societies. Am Nat 164: E154–E167.

Wenseleers T, Helanterä H, Hart A, Ratnieks FLW (2004b) Worker reproduction and policing in insect societies: an ESS analysis. J Evol Biol 17: 1035–1047.

Wenseleers T, Tofilski A, Ratnieks FLW (2005) Queen and worker policing in the tree wasp Dolichovespula sylvestris. Behav Ecol Sociobiol 58: 80–86.

Wenseleers T, Ratnieks FLW (2006) Comparative analysis of worker reproduction and policing in eusocial Hymenoptera supports relatedness theory. Am Nat 168: E163–E179.

Wenseleers T, Ratnieks FL (2006) Enforced altruism in insect societies. Nature 444: 50.

Wenzel JW, Pickering J (1991) Cooperative foraging, productivity, and the central limit theorem. Proc Natl Acad Sci USA 88: 36–38.

Werfel J, Bar-Yam Y (2004) The evolution of reproductive restraint through social communication. Proc Natl Acad Sci USA 101: 11019–11024.

West MJ (1967) Foundress associations in polistine wasps: Dominance hierarchies and the evolution of social behavior. Science 157: 1584–1585.

West SA, Murray MG, Machado CA, Griffin AS, Herre EA (2001) Testing Hamilton’s rule with competition between relatives. Nature 409: 510–513.

West SA, Pen I, Griffin AS (2002a) Conflict and cooperation—cooperation and competition between relatives. Science 296: 72–75.

West SA, Kiers ET, Simms EL, Denison RF (2002b) Sanctions and mutualism stability: why do rhizobia fix nitrogen? Proc R Soc Lond B 269: 685–694.

West SA, Kiers ET, Pen I, Denison RF (2002c) Sanctions and mutualism stability: when should less beneficial mutualists be tolerated? J Evol Biol 15: 830–837.

West SA, Buckling A (2003) Cooperation, virulence and siderophore production in bacterial parasites. Proc R Soc Lond B Biol Sci 270: 37–44.

West SA, Griffin AS, Gardner A, Diggle SP (2006) Social evolution theory for microorganisms. Nat Rev Microbiol 4: 597–607.

West SA, Griffin AS, Gardner A (2007a) Evolutionary explanations for cooperation. Curr Biol 17: R661–R672.

West SA, Diggle SP, Buckling A, Gardner A, Griffin AS (2007b) The social lives of microbes. Annu Rev Ecol Evol Syst 38: 53–77.

West SA, Griffin AS, Gardner A (2007c) Social semantics: altruism, cooperation, mutualism, strong reciprocity and group selection. J Evol Biol 20: 415–432.

West SA, Gardner A (2010) Altruism, spite, and greenbeards. Science 327: 1341–1344.

West-Eberhard ΜJ (1975) The evolution of social behavior by kin selection. Q Rev Biol 50: 1–33.

West-Eberhard ΜJ (1978) Polygyny and the evolution of social behaviour in wasps. J Kans Entomol Soc 51: 832–856.

West Eberhard MJ (1981) Intragroup selection and the evolution of insect societies. In: Alexander RD, Tinkle DW, eds. Natural selection and social behavior. New York, NY:  Chiron Press. pp 3–17.

West-Eberhard MJ (1987) The epigenetical origins of insect sociality. In: Eder J, Rembold H, eds. Chemistry and biology of social insects. Munich, Germany: Verlag J. Peperny. pp 369–372.

West-Eberhard MJ (2003) Developmental plasticity and evolution. Oxford, UK: Oxford University Press.

Westneat DF, Fox CW (2010) Evolutionary behavioral ecology. Oxford, UK: Oxford University Press.

Wey TW, Blumstein DT (2012) Social attributes and associated performance measures in marmots: bigger male bullies and weakly affiliating females have higher annual reproductive success. Behav Ecol Sociobiol 66: 1075–1085.

Weyl EG, Frederickson ME, Yu DW, Pierce NE (2010) Economic contract theory tests models of mutualism. Proc Natl Acad Sci USA 107: 15712–15716.

Wheeler DE (1986) Developmental and physiological determinants of caste in social Hymenoptera: Evolutionary implications. Am Nat 128: 13–34.

Wheeler DE, Buck N, Evans JD (2006) Expression of insulin pathway genes during the period of caste determination in the honey bee, Apis mellifera. Insect Mol Biol 15: 597–602.

Wheeler WM (1911) The ant colony as an organism. J Morphol 22: 307–325.

Wheeler WM (1928) The social insects. London, UK: Kegan Paul, Trench, Trubner & Co.

Whitacre JM, Bender A (2010) Degeneracy: a design principle for achieving robustness and evolvability. J Theor Biol 263: 143–153.

Whitchurch CB, Tolker-Nielsen T, Ragas PC, Mattick JS (2002) Extracellular DNA required for bacterial biofilm formation. Science 295: 1487.

Whitcomb WH, Bhatkar A, Nickerson JC (1973) Predators of Solenopsis invicta queens prior to successful colony establishment. Environ Entomol 2: 1101–1103.

White DJ, Gersick AS, Freed-Brown G, Snyder-Mackler N (2010) The ontogeny of social skills: experimental increases in social complexity enhance reproductive success in adult cowbirds. Anim Behav 79: 385–390.

White J, Leclaire S, Kriloff M, Mulard H, Hatch SA, Danchin E (2010) Sustained increase in food supplies reduces broodmate aggression in black-legged kittiwakes. Anim Behav 79: 1095–1100.

Whitehouse MAE, Lubin Y (2005) The functions of societies and the evolution of group living: spider societies as a test case. Biol Rev Camb Philos Soc 80: 347–361.

Whitfield J (2002) Social insects: the police state. Nature 416: 782–784.

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: 287–292.

Widdig A, Nürnberg P, Krawczak M, Streich WJ, Bercovitch FB (2001) Paternal relatedness and age proximity regulate social relationships among adult female rhesus macaques. Proc Natl Acad Sci USA 98: 13769–13773.

Widdig A, Nürnberg P, Krawczak M, Streich WJ, Bercovitch FB (2002) Affiliation and aggression among adult female rhesus macaques: a genetic analysis of paternal cohorts. Behaviour 139: 371–391.

Wiernasz DC, Hines J, Parker DG, Cole BJ (2008) Mating for variety increases foraging activity in the harvester ant, Pogonomyrmex occidentalis. Mol Ecol 17: 1137–1144.

Wijesinghe DK, Handel SN (1994) Advantages of clonal growth in heterogeneous habitats: an experiment with Potentilla simplex. J Ecol 82: 495–502.

Wiklund CG, Andersson M (1994) Natural selection of colony size in a passerine bird. J Anim Ecol 63: 765–774.

Wiklund C, Karlsson B, Leimar O (2001) Sexual conflict and cooperation in butterfly reproduction: a comparative study of polyandry and female fitness. Proc R Soc Lond Ser B Biol Sci 268: 1661–1667.

Wilbur HM, Tinkle DW, Collins JP (1974) Environmental certainty, trophic level, and resource availability in life history evolution. Am Nat 108: 805–817.

Wild G, Traulsen A (2007) The different limits of weak selection and the evolutionary dynamics of finite populations. J Theor Biol 247: 382–390.

Wild G, Gardner A, West SA (2010) Wild, Gardner & West reply. Nature 463: E9–E10.

Wilder CN, Allada G, Schuster M (2009) Instantaneous within-patient diversity of Pseudomonas aeruginosa quorum-sensing populations from cystic fibrosis lung infections. Infect Immun 77: 5631–5639.

Wiley RH, Rabenold KN (1984) The evolution of cooperative breeding by delayed reciprocity and queuing for favorable social positions. Evolution 38: 609–621.

Wilfert L, Gadau J, Schmid-Hempel P (2007) Variation in genomic recombination rates among animal taxa and the case of social insects. Heredity 98: 189–197.

Wilkinson DJ (2009) Stochastic modelling for quantitative description of heterogeneous biological systems. Nat Rev Genet 10: 122–133.

Wilkinson DM, Sherratt TN (2001) Horizontally acquired mutualisms, an unsolved problem in ecology? Oikos 92: 377–384.

Wilkinson GS (1984) Reciprocal food sharing in the vampire bat. Nature 308: 181–184.

Wilkinson GS (1985) The social organization of the common vampire bat: II. Mating system, genetic structure, and relatedness. Behav Ecol Sociobiol 17: 123–134.

Wilkinson GS (1988) Reciprocal altruism in bats and other mammals. Ethol Sociobiol 9: 85–100.

Wilkinson GS (1990) Food sharing in vampire bats. Sci Am 262: 76–82.

Wilkinson HH, Spoerke JM, Parker AM (1996) Divergence in symbiotic compatibility in a legume-Bradyrhizobium mutualism. Evolution 50: 1470–1477.

Williams DA (2004) Female control of reproductive skew in cooperatively breeding brown jays (Cyanocorax morio). Behav Ecol Sociobiol 55: 370–380.

Williams GC (1966a) Natural selection, the costs of reproduction, and a refinement of Lack’s principle. Am Nat 100: 687–690.

Williams GC (1966b) Adaptation and natural selection. Princeton, NJ: Princeton University Press.

Williams GC (1975) Sex and evolution. Princeton, NJ: Princeton University Press.

Williams GC, Williams DC (1957) Natural selection of individually harmful social adaptations among sibs with special reference to social insects. Evolution 11: 32–39.

Williams JG (2006) Transcriptional regulation of Dictyostelium pattern formation. EMBO Rep 7: 694–698.

Williams JR, Insel TR, Harbaugh CR, Carter CS (1994) Oxytocin administered centrally facilitates formation of a partner preference in female prairie voles (Microtus ochrogaster). J Neuroendocrinol 6: 247–250.

Williams PD, Day T (2003) Antagonistic pleiotropy, mortality source interactions, and the evolutionary theory of senescence. Evolution 57: 1478–1488.

Williamson DK, Dunbar RIM (1999) Energetics, time budgets and group size. In: Lee, PC, ed. Comparative primate socioecology. Cambridge, UK: Cambridge University Press. pp 320–338.

Willis LD Jr, Hendrick AC (1992) Life history, growth, survivorship, and production of Hydropysche slossonae in Mill Creek, Virginia. J N Am Benthol Soc 11: 290–303.

Willson MF, Thomas PA, Hoppes WG, Katusic-Malmborg PL, Goldman DA, Bothwell JL (1987) Sibling competition in plants: an experimental study. Am Nat 129: 304–311.

Wilson DS (1975) A theory of group selection. Proc Natl Acad Sci USA 72: 143–146.

Wilson DS (1983) The group selection controversy: history and current status. Annu Rev Ecol Syst 14: 159–187.

Wilson DS (1990) Weak altruism, strong group selection. Oikos 59: 135–140.

Wilson DS (2001) Cooperation and altruism. In: Fox CW, Roff DA, Fairbairn DJ, eds. Evolutionary ecology: concepts and case studies. Oxford, UK: Oxford University Press. pp 222–231.

Wilson DS (2015) Does altruism exist? Culture, genes, and the welfare of others. New Haven, CT: Yale University Press.

Wilson DS, Sober E (1989) Reviving the superorganism. J Theor Biol 136: 337–356.

Wilson DS, Dugatkin LA (1992) Altruism: contemporary debates. In: Keller EF, Lloyd EA, eds. Keywords in evolutionary biology. Cambridge, MA: Harvard University Press. pp 29–33.

Wilson DS, Pollock GB, Dugatkin LA (1992) Can altruism evolve in purely viscous populations? Evol Ecol 6: 331–341.

Wilson DS, Wilson EO (2007a) Rethinking the theoretical foundation of sociobiology. Q Rev Biol 82: 327–348.

Wilson DS, Wilson EO (2007b) Evolution: survival of the selfless. New Sci 196: 42–46.

Wilson DS, Wilson EO (2008) Evolution ‘‘for the good of the group.’’ Am Sci 96: 380–389.

Wilson EO (1971) The insect societies. Cambridge, MA: Harvard University Press.

Wilson EO (1975) Sociobiology: a new synthesis. Cambridge, UK: Cambridge University Press.

Wilson E (1978) What is sociobiology? Society 15: 10–14.

Wilson EO (1985) The sociogenesis of insect colonies. Science 228: 1489–1495.

Wilson EO (1987) Causes of ecological success: the case of the ants. J Anim Ecol 56: 1–9.

Wilson EO (2005) Kin selection as the key to altruism: its rise and fall. Soc Res 72: 159–166.

Wilson EO (2008) One giant leap: how insects achieved altruism and colonial life. BioScience 58: 17–25.

Wilson EO (2012) The social conquest of earth. W. W. Norton.

Wilson EO, Hölldobler B (2005) Eusociality: origin and consequences. Proc Natl Acad Sci USA 102: 13367–13371.

Wilson EO, Nowak MA (2014) Natural selection drives the evolution of ant life cycles. Proc Natl Acad Sci USA 111: 12585–12590.

Wilson JB (1987) Group selection in plant populations. Theor Appl Genet 74: 493–502.

Wilson JS, von Dohlen CD, Forister ML, Pitts JP (2013) Family-level divergences in the stinging wasps (Hymenoptera: Aculeata), with correlations to angiosperm diversification. Evol Biol 40: 101–107.

Wilson M, Daly M (1997) Life expectancy, economic inequality, homicide, and reproductive timing in Chicago neighbourhoods. Brit Med J 314: 1271–1274.

Wingfield JC, Hegner RE, Lewis DM (1991) Circulating levels of luteinizing hormone and steroid hormones in relation to social status in the cooperatively breeding white-browed sparrow weaver, Plocepasser mahali. J Zool 225: 43–58.

Wingfield JC, Jacobs JD, Tramontin AD, Perfito N, Meddle S, Maney DL, Soma K (2000) Toward an ecological basis of hormone–behavior interactions in reproduction of birds. In: Wallen K, Schneider J, eds. Reproduction in context. Cambridge, MA: MIT Press. pp 85–128.

Wingfield JC, Sapolsky RM (2003) Reproduction and resistance to stress: When and how. J Neuroendocrinol 15: 711–724.

Wingreen NS, Levin SA (2006) Cooperation among microorganisms. PLoS Biol 4: e299.

Winslow JT, Insel TR (1991) Social status in pairs of male squirrel monkeys determines the behavioral response to central oxytocin administration. J Neurosci 11: 2032–2038.

Winslow JT, Hastings N, Carter CS, Harbaugh CR, Insel TR (1993) A role for central vasopressin in pair bonding in monogamous prairie voles. Nature 365: 545–548.

Winston ML, Fergusson LA (1985) The effect of worker loss on temporal caste structure in colonies of the honeybee (Apis mellifera L.). Can J Zool 63: 777–780.

Winston ML, Slessor KN (1998) Honey bee primer pheromones and colony organization: gaps in our knowledge. Apidologie 29: 81–95.

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.

Wireman JW, Dworkin M (1977) Developmentally induced autolysis during fruiting body formation by Myxococcus xanthus. J Bacteriol 129: 798–802.

Wirth R, Muscholl A, Wanner G (1996) The role of pheromones in bacterial interactions. Trends Microbiol 4: 96–103.

Wise DH (2006) Cannibalism, food limitation, intraspecific competition and the regulation of spider populations. Annu Rev Entomol 51: 441–465.

Wittenberger JF (1980) Group size and polygamy in social mammals. Am Nat 115: 197–222.

Wolf JB, Brodie ED 3rd, Cheverud JM, Moore AJ, Wade M (1998) Evolutionary consequences of indirect genetic effects. Trends Ecol Evol 13: 64–69.

Wolf M, Kurvers RHJM, Ward AJW, Krause S, Krause J (2013) Accurate decisions in an uncertain world: collective cognition increases true positives while decreasing false positives. Proc R Soc B 280: 20122777.

Wolfe MS (1985) The current status and prospects of multiline cultivars and variety mixtures for disease resistance. Annu Rev Phytopathol 23: 251–273.

Wolff JO (1993) Why are female small mammals territorial. Oikos 68: 364–370.

Wolff JO, Cicirello DM (1989) Field evidence for sexual selection and resource competition infanticide in white-footed mice. Anim Behav 38: 637–642.

Wolschin F, Mutti NS, Amdam GV (2011) Insulin receptor substrate influences female caste development in honeybees. Biol Lett 7: 112–115.

Wong JWY, Kölliker M (2012) The effect of female condition on maternal care in the European earwig. Ethology 118: 450–459.

Wong JWY, Kölliker M (2013) The more the merrier? Condition-dependent brood mixing in earwigs. Anim Behav 86: 845–850.

Wong JWY, Meunier J, Kölliker M (2013) The evolution of parental care in insects: the roles of ecology, life history and the social environment. Ecol Entomol 38: 123–137.

Wong JWY, Lucas C, Kölliker M (2014) Cues of maternal condition influence offspring selfishness. PLoS ONE 9: e87214.

Wong M, Balshine S (2011) The evolution of cooperative breeding in the African cichlid fish, Neolamprologus pulcher. Biol Rev 86: 511–530.

Wong PT, Feng H, Teo WL (1995) Interaction of the dopaminergic and serotonergic systems in the rat striatum: Effects of selective antagonists and uptake inhibitors. Neurosci Res 23: 115−119.

Wood RM, Rilling JK, Sanfey AG, Bhagwagar Z, Rogers RD (2006) Effects of tryptophan depletion on the performance of an iterated prisoner’s dilemma game in healthy adults. Neuropsychopharmacology 31: 1075–1084.

Woodroffe R, Ginsberg JR, Macdonald DW & the IUCN/SSC Canid Specialist Group, eds. (1997) The African wild dog. Status survey & conservation action plan. Gland, Switzerland: IUCN.

Woods HA, Hill RI (2004) Temperature-dependent oxygen limitation in insect eggs. J Exp Biol 207: 2267–2276.

Woolfenden GE (1975) Florida scrub jay helpers at the nest. Auk 92: 1–15.

Woolfenden GE, Fitzpatrick JW (1978) The inheritance of territory in group-breeding birds. BioScience 28: 104–108.

Woolfenden GE, Fitzpatrick JW (1984) The Florida scrub jay: demography of a cooperative-breeding bird. Princeton, NJ: Princeton University Press.

Worden L, Levin SA (2007) Evolutionary escape from the prisoner's dilemma. J Theor Biol 245: 411–422.

Wossler TC (2002) Pheromone mimicry by Apis mellifera capensis social parasites leads to reproductive anarchy in host Apis mellifera scutellata colonies. Apidologie 33: 139–164.

Wossler TC, Crewe RM (1999) Honeybee queen tergal gland secretion affects ovarian development in caged workers. Apidologie 30: 311–320.

Woyciechowski M (1990) Do honey bee, apis mellifera L., workers favour sibling eggs and larvae in Queen rearing? Anim Behav 39: 1220–1222.

Woyciechowski M, Lomnicki A (1987) Multiple mating of queens and the sterility of workers among eusocial Hymenoptera. J Theor Biol 128: 317–327.

Woyciechowski M, Kozlowski J (1998) Division of labor by division of risk according to worker life expectancy in the honey bee (Apis mellifera L.). Apidologie 29: 191–205.

Wrangham RW (1980) An ecological model of female-bonded primate groups. Behaviour 75: 262–300.

Wrangham RW (1983) Ultimate factors determining social structure. In: Hinde RA, ed. Primate social relationships: an integrated approach. Sunderland, MA: Sinauer. pp 255–262.

Wrangham RW, Gittleman JL, Chapman CA (1993) Constraints on group size in primates and carnivores: population density and day-range as assays of exploitation competition. Behav Ecol Sociobiol 32: 199–209.

Wright GA (2009) Bee pheromones: signal or agent of manipulation? Curr Biol 19: R547–R548.

Wright JJ, Bourke PD (2013) On the dynamics of cortical development: synchrony and synaptic self-organization. Front Comput Neurosci 7: 4.

Wright ND, Bahrami B, Johnson E, Di Malta G, Rees G, Frith CD, Dolan RJ (2012) Testosterone disrupts human collaboration by increasing egocentric choices. Proc R Soc B Biol Sci DOI: 10.1098/rspb.2011.2523.

Wright S (1931) Evolution in Mendelian populations. Genetics 16: 97–159.

Wright S (1968) Evolution and the genetics of populations. Vol. 1. Chicago, IL: University of Chicago Press.

Wright S (1969) Evolution and the genetics of populations, Vol. 2. Chicago, IL: University of Chicago Press.

Wu B, Zhou D, Fu F, Luo Q, Wang L, et al. (2010) Evolution of cooperation on stochastic dynamical networks. PLoS ONE 5: e11187.

Wu G (1994) An empirical test of ordinal independence. J Risk Uncertainty 9: 39–60.

Wu JJ, Zhang BY, Zhou ZZ, He QQ, Zheng XD, Cressman R, Tao Y (2009) Costly punishment does not always increase cooperation. Proc Natl Acad Sci USA 106: 17448–17451.

Wu JZ, Axelrod R (1995) How to cope with noise in the iterated Prisoner’s Dilemma. J Conflict Resolut 39: 183–189.

Wu M, Su RQ, Li X, Ellis T, Lai YC, Wang X (2013) Engineering of regulated stochastic cell fate determination. Proc Natl Acad Sci USA 110: 10610–10615.

Wu N, Li Z, Su Y (2012) The association between oxytocin receptor gene polymorphism (OXTR) and trait empathy. J Affect Disord 138: 468–472.

Wu Q, Brown MR (2006) Signaling and function of insulin-like peptides in insects. Annu Rev Entomol 51: 1–24.

Wu ZX, Holme P (2009) Effects of strategy-migration direction and noise in the evolutionary spatial prisoner’s dilemma. Phys Rev E 80: 026108.

Wyatt TD (2003) Pheromones and animal behaviour: communication by smell and taste. Cambridge, UK: Cambridge University Press.

Xavier JB (2011) Social interaction in synthetic and natural microbial communities. Mol Syst Biol 7: 483.

Xavier JB, Foster KR (2007) Cooperation and conflict in microbial biofilms. Proc Natl Acad Sci USA 104: 876–881.

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: 656–661.

Xiong W, Ferrell JE Jr (2003) A positive-feedback-based bistable ‘memory module’ that governs a cell fate decision. Nature 426: 460–465.

Yaari G, Solomon S (2010) Cooperation evolution in random multiplicative environments. Eur Phys J B 73: 625–632.

Yakubu Y (2012) Altruism: analysis of a paradox. Thesis. Hamilton, Canada: McMaster University.

Yakubu Y (2013) The altruism paradox: a consequence of mistaken genetic modeling. Biol Theor 8: 103–113.

Yamada Y, Minamisawa H, Fukuzawa M, Kawata T, Oohata AA (2010) Prespore cell inducing factor, psi factor, controls both prestalk and prespore gene expression in Dictyostelium development. Dev Growth Differ 52: 377–383.

Yamaguchi S, Kimura H, Tada M, Nakatsuji N, Tada T (2005) Nanog expression in mouse germ cell development. Gene Expr Patterns 5: 639–646.

Yamamoto A, Anholt RRH, Mackay TFC (2009) Epistatic interactions attenuate mutations that affect startle behaviour in Drosophila melanogaster. Genet Res 91: 373–382.

Yamamoto S, Takimoto A (2012) Empathy and fairness: psychological mechanisms for eliciting and maintaining prosociality and cooperation in primates. Soc Just Res 25: 233–255.

Yamamura N, Higashi M, Behera N, Wakano JY (2004) Evolution of mutualism through spatial effects. J Theor Biol 226: 421–428.

Yamasaki H, Mesnil M, Omori Y, Mironov N, Krutovskikh V (1995) Intercellular communication and carcinogenesis. Mutat Res 333: 181–188.

Yamazaki K, Boyse EA, Mike V, Thaler HT, Mathieson BJ, Abbott J, et al. (1976) Control of mating preferences in mice by genes in the major histocompatibility complex. J Exp Med 144: 1324–1335.

Yamazaki K, Yamagughi M, Baranoski L, Bard J, Boyse EA, Thomas L (1979) Recognition among mice: evidence from the use of a Y-maze differentially scented by congenic mice of different major histocompatibility types. J Exp Med 150: 755–760.

Yamazaki K, Beauchamp GK, Singer A, Bard J, Boyse EA (1999) Odortypes: Their origin and composition. Proc Natl Acad Sci USA 96: 1522–1525.

Yamazaki K, Beauchamp GK, Curran M, Bard J, Boyse EA (2000) Parent–progeny recognition as a function of MHC odortype identity. Proc Natl Acad Sci USA 97: 10500–10502.

Yanega D (1988) Social plasticity and early-diapausing females in a primitively social bee. Proc Natl Acad Sci USA 85: 4374–4377.

Yanega D (1989) Castes determination and di¡erential diapause within the first brood of Halictus rubicundus. Behav Ecol Sociobiol 24: 97–107.

Yates CA, Erban R, Escudero C, Couzin ID, Buhl J, Kevrekidis IG, et al. (2009) Inherent noise can facilitate coherence in collective swarm motion. Proc Natl Acad Sci USA 106: 5464–5469.

Ydenberg RC (2007) Provisioning. In: Stephens DW, Brown JS, Ydenberg RC, eds. Foraging: behavior and ecology. Chicago, IL: Chicago University Press. pp 273–303.

Yee JR, Cavigelli SA, Delgado B, McClintock MK (2008) Reciprocal affiliation among adolescent rats during a mild group stressor predicts mammary tumors and lifespan. Psychosom Med 70: 1050–1059.

Yildiz FH, Schoolnik GK (1999) Vibrio cholerae O1 El Tor: identification of a gene cluster required for the rugose colony type, exopolysaccharide production, chlorine resistance, and biofilm formation. Proc Natl Acad Sci USA 96: 4028–4033.

Yin Y, Stahl BC, DeWolf WC, Morgentaler A (1998) p53-mediated germ cell quality control in spermatogenesis. Dev Biol 204: 165–171.

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.

Yip EC, Powers KS, Avilés L (2008) Cooperative capture of large prey solves scaling challenge faced by spider societies. Proc Natl Acad Sci USA 105: 11818–11822.

Yip EC, Clarke S, Rayor LS (2009) Aliens among us: nestmate recognition in the social huntsman spider, Delena cancerides. Insect Soc 56: 223–231.

Yorimitsu T, Klionsky DJ (2007) Eating the endoplasmic reticulum: quality control by autophagy. Trends Cell Biol 17: 279–285.

Young AJ, Carlson AA, Monfort SL, Russell AF, Bennett NC, Clutton-Brock TH (2006) Stress and the suppression of subordinate reproduction in cooperatively breeding meerkats. Proc Natl Acad Sci USA 103: 12005–12010.

Young AJ, Clutton-Brock TH (2006) Infanticide by subordinates influences reproductive sharing in cooperatively breeding meerkats. Biol Lett 2: 385–387.

Young AJ, Monfort SL (2009) Stress and the costs of extra-territorial movement in a social carnivore. Biol Lett 5: 439–441.

Young JPW (1981) Sib competition can favor sex in two ways. J Theor Biol 88: 755–756.

Young KA, Gobrogge KL, Liu Y, Wang Z (2011) The neurobiology of pair bonding: insights from a socially monogamous rodent. Front Neuroendocrinol 32: 53–69.

Young LJ (1999) Oxytocin and vasopressin receptors and species-typical social behaviors. Horm Behav 36: 212–221.

Young LJ, Huot B, Nilsen R, Wang Z, Insel TR (1996) Species differences in central oxytocin receptor gene expression: Comparative analysis of promoter sequences. J Neuroendocrinol 8: 777–783.

Young LJ, Nilsen R, Waymire KG, MacGregor GR, Insel TR (1999) Increased affiliative response to vasopressin in mice expressing the V1a receptor from a monogamous vole. Nature 400: 766–768.

Young LJ, Lim MM, Gingrich B, Insel TR (2001) Cellular mechanisms of social attachment. Horm Behav 40: 133–138.

Young LJ, Wang Z (2004) The neurobiology of pair bonding. Nat Neurosci 7: 1048–1054.

Yu DW (2001) Parasites of mutualisms. Biol J Linn Soc 72: 529–546.

Yu YT, Yuan X, Velicer GJ (2010) Adaptive evolution of an sRNA that controls Myxococcus development. Science 328: 993.

Yuan HW, Shen SF, Hung HY (2006) Sexual dimorphism, dispersal patterns, and breeding biology of the Taiwan Yuhina: a joint-nesting passerine. Wilson J Ornithol 118: 558–562.

Yund PO, Feldgarden M (1992) Rapid proliferation of historecognition alleles in populations of a colonial ascidian. J Exp Zool 263: 442–452.

Zachos J, Pagani M, Sloan L, Thomas E, Billups K (2001) Trends, rhythms, and aberrations in global climate 65 Ma to present. Science 292: 686–693.

Zack S (1990) Coupling delayed breeding with short-distance dispersal in cooperatively breeding birds. Ethology 86: 265–286.

Zack S, Stutchbury BJ (1992) Delayed breeding in avian social systems: the role of territory quality and “floater” tactics. Behaviour 123: 194–219.

Zahavi A (1990) Arabian babblers: the quest for social status in a cooperative breeder. In: Stacey PB, Koenig WD, eds. Cooperative breeding in birds: long-term studies of ecology and behaviour. Cambridge, UK: Cambridge University Press. pp 105–130.

Zahavi A (1995) Altruism as a handicap. The limits of kin selection and reciprocity. J Avian Biol 26: 1–3.

Zak PJ, Kurzban R, Matzner WT (2005) Oxytocin is associated with human trustworthiness. Horm Behav 48: 522–527.

Zak PJ, Stanton AA, Ahmadi S (2007) Oxytocin increases generosity in humans. PLoS ONE 2: e1128.

Zambrano MM, Siegele DA, Almiron M, Tormo A, Kolter R (1993) Microbial competition: Escherichia coli mutants that take over stationary phase cultures. Science 259: 1757–1760.

Zanette LRS, Field J (2008) Genetic relatedness in early associations of Polistes dominulus: from related to unrelated helpers. Mol Ecol 17: 2590–2597.

Zanette LRS, Miller SDL, Faria CMA, Almond EJ, Huggins TJ, Jordan WC, Bourke AFG (2012) Reproductive conflict in bumblebees and the evolution of worker policing. Evolution 66: 3765–3777.

Zayed A, Packer L (2005) Complementary sex determination substantially increases extinction proneness of haplodiploid populations. Proc Natl Acad Sci USA 102: 10742–10746.

Zelano B, Edwards VE (2002) An MHC component to kin recognition and mate choice in birds: predictions, progress, and prospects. Am Nat 160(Suppl.): S225–S237.

Zhang B, Wolynes PG (2014) Stem cell differentiation as a many-body problem. Proc Natl Acad Sci USA 111: 10185–10190.

Zhang J, He X (2005) Significant impact of protein dispensability on the instantaneous rate of protein evolution. Mol Biol Evol 22: 1147–1155.

Zhang JL, Zhang CY, Chu TG, Perc M (2011) Resolution of the stochastic strategy spatial prisoner’s dilemma by means of particle swarm optimization. PLoS ONE 6: e21787.

Zhang L, Li WH (2004) Mammalian housekeeping genes evolve more slowly than tissue-specific genes. Mol Biol Evol 21: 236–239.

Zhang QG, Buckling A, Ellis RJ, Godfray HCJ (2009) Coevolution between cooperators and cheats in a microbial system. Evolution 63: 2248–2256.

Zhang YQ, Li YZ, Wang B, Wu ZH, Zhang CY, Gong X, et al. (2005) Characteristics and living patterns of marine myxobacterial isolates. Appl Environ Microb 71: 3331–3336.

Zhang Z (2003) Mutualism or cooperation among competitors promote coexistence and competitive ability. Ecol Model 164: 271–282.

Zheng M, Doan B, Schneider TD, Storz G (1999) OxyR and SoxRS regulation of fur. J Bacteriol 181: 4639–4643.

Zheng M, Kashimori Y, Hoshino O, Fujita K, Kambara T (2005) Behavior pattern (innate action) of individuals in fish schools generating efficient collective evasion from predation. J Theor Biol 235: 153–167.

Zheng Q, Ross J (1991) Comparison of deterministic and stochastic kinetics for nonlinear systems. J Chem Phys 94: 3644–3648.

Ziemek S (2006) Economic analysis of volunteers’ motivations—A cross-country study. J Socio Econ 35: 532–555.

Zimmering S, Sandler L, Nicoletti B (1970) Mechanisms of meiotic drive. Annu Rev Genet 4: 409–436.

Zinck L, Chaline N, Jaisson P (2009) Absence of nepotism in worker-queen care in polygynous colonies of the ant Ectatomma tuberculatum. J Insect Behav 22: 196–204.

Zink AG (2003) Quantifying the costs and benefits of parental care in female treehoppers. Behav Ecol 14: 687–693.

Zöttl M, Heg D, Chervet N, Taborsky M (2013a) Kinship reduces alloparental care in cooperative cichlids where helpers pay-to-stay. Nat Commun 4: 1341.

Zöttl M, Frommen JG, Taborsky M (2013b) Group size adjustment to ecological demand in a cooperative breeder. Proc R Soc B 280: 20122772.

Zube C, Kleineidam C, Kirschner S, Neef J, Rössler W (2008) Organization of the olfactory pathway and odor processing in the antennal lobe of the ant Camponotus floridanus. J Comp Neurol 506: 425–441.

Source(s) of Funding


no funding

Competing Interests


no competing interests

Reviews
0 reviews posted so far

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)