Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Solving the prisoner’s dilemma trap in Hamilton’s model of temporarily formed random groups

José F. Fontanari fontanari@ifsc.usp.br Mauro Santos mauro.santos@uab.es Instituto de Física de São Carlos, Universidade de São Paulo, 13560-970 São Carlos, São Paulo, Brazil Departament de Genètica i de Microbiologia, Grup de Genòmica, Bioinformàtica i Biologia Evolutiva (GBBE), Universitat Autònoma de Barcelona, Spain
cE3c - Centre for Ecology, Evolution and Environmental Changes & CHANGE - Global Change and Sustainability Institute, Lisboa, Portugal
Abstract

Explaining the evolution of cooperation in the strong altruism scenario, where a cooperator does not benefit from her contribution to the public goods, is a challenging problem that requires positive assortment among cooperators (i.e., cooperators must tend to associate with other cooperators) or punishment of defectors. The need for these drastic measures stems from the analysis of a group selection model of temporarily formed random groups introduced by Hamilton nearly fifty years ago to describe the fate of altruistic behavior in a population. Challenging conventional wisdom, we show analytically here that strong altruism evolves in Hamilton’s original model in the case of biparental sexual reproduction. Moreover, when the cost of cooperation is small and the amplified contribution shared by group members is large, cooperation is the only stable strategy in equilibrium. Thus, our results provide a solution to the ‘problem of origination’ of strong altruism, i.e. how cooperation can take off from an initial low frequency of cooperators. We discuss a possible reassessment of cooperation in cases of viral co-infection, as cooperation may even be favored in situations where the prisoner’s dilemma applies.

1 Introduction

As noted by Darwin in his discussion of courage and self-sacrifice, traits associated with altruistic behavior are counter-selected within a group, but increase the group’s productivity and the group’s chances of survival in the case of intergroup fighting [1]. Recognizing that the possible outcomes of this conflict between different levels of selection could only be determined by mathematical analysis, two slightly different models of group selection were proposed nearly fifty years ago to study the conditions for the evolution of altruistic behavior [2, 3]. Both models assume temporary random grouping of individuals, but differ in the way in which within-group competition between selfish and altruistic individuals is implemented. Hamilton’s model follows the Darwinian view that altruists are always at disadvantage in local competition with selfish individuals, but groups with more altruists contribute more offspring to a pool of dispersers that will eventually form the new groups [2]. Wilson’s trait group model ignores within-group competition altogether: although it assumes that individuals acquire their fitness locally, the competition between altruistic and selfish individuals takes place globally, i.e., in the population at large [3, 4]. Recently, we have pointed out the equivalence between Wilson’s model and the evolutionary game approach to studying the emergence of cooperation in n𝑛nitalic_n-person public goods games, which is based on the replicator equation framework [5]. We are aware of the distinction between cooperation, which involves mutual benefit, and altruism, which implies ‘others-only’ benefit [6, 7]. However, we will use both terms, cooperation and altruism, more or less interchangeably here.

The outcome of these group selection models depends on whether an altruist benefits from her altruistic behavior or not. When the return from altruistic behavior exceeds the cost of performing it, we have the scenario of weak altruism [4, 8]. This is still altruism in the sense that selfish individuals benefit the most from this behavior. The traditional n𝑛nitalic_n-person prisoner’s dilemma [9, 10, 11] and the n𝑛nitalic_n-person snowdrift game [5, 12, 13, 14] fall into this category. Both Hamilton’s and Wilson’s models of group selection with randomly formed groups predict that altruism will evolve in this scenario, a clearly impossible outcome for homogeneous or unstructured populations, which explains the notoriety of these models. When an altruist has no return from her altruistic behavior, as in the others-only variant of the n𝑛nitalic_n-person prisoner’s dilemma [2, 15, 16], we have the more challenging scenario of strong altruism. Hamilton’s seminal paper addressed this scenario and concluded that altruism cannot progress in randomly formed groups: cooperation is impossible unless there is some form of positive assortment among cooperators [2]. Wilson came to the same conclusion when groups are formed by binomially sampling individuals from the total population, as in Hamilton’s model, but mistakenly suggested that environmental heterogeneity could produce a sufficiently large excess of variance over the binomial expectation that strong altruism might be favored [3]. This error was quickly corrected [17], so that both models now point to the impossibility of strong altruism evolving in the absence of explicit positive assortment among altruists, i.e., altruists must have a tendency to group with other altruists. Nonetheless, we show here, using both analytical and simulation results, that strong altruism can evolve in Hamilton’s model without assuming positive assortment. This is achieved by not taking the phenotypic gambit in full [18].

The phenotypic gambit, often used in evolutionary game theory, assumes, among other things, the simplest genetic system (e.g., haploid individuals). Although, this is usually wrong, we can generally ignore genetics to study the evolutionary fate of behavioral strategies, except in the case of diploidy, when the heterozygote is the fittest genotype [18] (see also [19]). Let us accept the gambit and assume that organisms are haploid and that each chromosome consists of L𝐿Litalic_L loci, one of which controls one of the only two pure strategies, altruistic (cooperator) and selfish (defector). Now consider three possible modes of reproduction that are commonly used in theoretical models: asexual reproduction, sexual reproduction (production of haploid gametes that combine in pairs to produce a recombinant offspring chromosome) that allows self-mating, and sexual reproduction that requires two different individuals, what Williams calls euphrasic [20]. Thus, mating pairs are formed by randomly drawing individuals (with or without replacement) according to their relative fitness, and each pair produces a single offspring. This is still a simplified picture of a more complex reality, where organisms can be diploid and the behavioral strategy is likely to be controlled by many loci with possible epistatic interactions. The question here is whether the phenotypic gambit, which in this case would also ignore the complications introduced by the mode of reproduction, is still valid. If so, we would conclude that strong altruism cannot evolve in Hamilton’s model regardless of how individuals reproduce.

Here, we show that the way we model reproduction can have a qualitative impact on the evolution of altruism in ways that the phenotypic approach cannot predict or explain. In particular, we show that strong altruism can evolve under biparental sexual reproduction in Hamilton’s model [2].

The rest of the paper is organized as follows. In Section 2 we describe Hamilton’s model of temporarily formed random groups, assuming G𝐺Gitalic_G groups of size n𝑛nitalic_n, using the terminology of the n𝑛nitalic_n-person prisoner’s dilemma. In Section 3 we develop an analytical treatment to solve Hamilton’s model in the deterministic limit G𝐺G\to\inftyitalic_G → ∞ and show that in the case of asexual reproduction, cooperation is doomed to extinction in the strong altruism scenario, but not in the weak altruism scenario. In addition, we show that sexual reproduction, allowing self-mating, leads to the same conclusion as in the asexual case. We then consider biparental sexual reproduction and, challenging conventional wisdom, derive the conditions that guarantee the emergence of cooperation in the strong altruism scenario. In all cases, we follow the fate of the altruistic or selfish strategy, which for sexual reproduction would be analogous to following the allelic changes in the recombinant offspring chromosomes hosting the behavioral gene. In Section 4 we use simulations to prove the robustness of the conclusions of the deterministic analysis to the effect of demographic noise that arises when G𝐺Gitalic_G is finite. Finally, in Section 5 we summarize our results and also suggest a possible reassessment of cooperation in cases of viral co-infection, as cooperation may even be favored in situations where the prisoner’s dilemma applies. Returning to the phenotypic gambit, we cannot avoid ending this section by recalling a quote from Albert Einstein (possibly misattributed; see [21]): “Make things as simple as possible, but not simpler.”.

2 Model

Here we recast Hamilton’s model of temporarily formed random groups [2] using the terminology of public goods games [9]. Consider a population of Gn𝐺𝑛Gnitalic_G italic_n individuals, divided into G𝐺Gitalic_G groups, each of which contains n𝑛nitalic_n individuals. Within each group, individuals play a n𝑛nitalic_n-person prisoner’s dilemma game to determine their fitness: each individual may (or may not) contribute an amount c>0𝑐0c>0italic_c > 0 to the public goods, which is then multiplied by a factor r>1𝑟1r>1italic_r > 1, and the resulting amount rc𝑟𝑐rcitalic_r italic_c is divided among the n1𝑛1n-1italic_n - 1 other players. This is the others-only variant of the n𝑛nitalic_n-person prisoner’s dilemma, where contributors do not benefit from their own contributions [15]. We refer to individuals that contribute to the public goods as cooperators and those that do not as defectors.

The fitness (or payoff) of a cooperator in a group with i1𝑖1i-1italic_i - 1 other cooperators is

fc(i)=1c+(i1)rcn1,subscript𝑓𝑐𝑖1𝑐𝑖1𝑟𝑐𝑛1f_{c}(i)=1-c+(i-1)\frac{rc}{n-1},italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) = 1 - italic_c + ( italic_i - 1 ) divide start_ARG italic_r italic_c end_ARG start_ARG italic_n - 1 end_ARG , (1)

whereas the fitness of a defector in a group with i𝑖iitalic_i cooperators is

fd(i)=1+ircn1.subscript𝑓𝑑𝑖1𝑖𝑟𝑐𝑛1f_{d}(i)=1+i\frac{rc}{n-1}.italic_f start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_i ) = 1 + italic_i divide start_ARG italic_r italic_c end_ARG start_ARG italic_n - 1 end_ARG . (2)

These are the payoffs of the popular n𝑛nitalic_n-person prisoner’s dilemma [9, 15, 16] with a constant baseline fitness value of 1111. They are identical to the fitness attributed to altruistic and selfish individuals in Hamilton’s model of randomly formed groups: in Hamilton’s terminology, an altruist gives up k=c𝑘𝑐k=citalic_k = italic_c units of her own fitness in order to add K=rc𝐾𝑟𝑐K=rcitalic_K = italic_r italic_c units to the joint fitness of her n1𝑛1n-1italic_n - 1 companions [2]. We will assume c[0,1]𝑐01c\in[0,1]italic_c ∈ [ 0 , 1 ] to ensure that the fitness values are not negative, so c=1𝑐1c=1italic_c = 1 is the worst scenario for cooperation. The parallel between the n𝑛nitalic_n-person prisoner’s dilemma and Hamilton’s model ends here, however. In fact, in Hamilton’s model the fitness of a group with i𝑖iitalic_i cooperators,

π(i)=n+ic(r1),𝜋𝑖𝑛𝑖𝑐𝑟1\pi(i)=n+ic(r-1),italic_π ( italic_i ) = italic_n + italic_i italic_c ( italic_r - 1 ) , (3)

determines the number of offspring m𝑚mitalic_m produced by the group. In particular, we assume that m𝑚mitalic_m is a random variable distributed by a Poisson distribution with mean π(i)𝜋𝑖\pi(i)italic_π ( italic_i ). Group fitness plays no role in the n𝑛nitalic_n-person prisoner’s dilemma, where individuals decide whether or not to keep their strategies by comparing their payoffs to those of randomly selected individuals in the population at large [22, 23], as in Wilson’s trait group formulation [3, 4], so there is effectively no within-group competition.

In Hamilton’s model, within-group competition takes place during the reproduction stage. As mentioned before, we will consider three modes of reproduction. In the asexual reproduction mode, which is the one considered in Hamilton’s seminal paper, we select one parent within each group with probability equal to her relative fitness, and the single offspring inherits the strategy of the parent. For each group, exactly m𝑚mitalic_m parents are selected in this way, and the resulting m𝑚mitalic_m offspring are added to the offspring pool. Hamilton refers to the offspring pool as the migrant pool, which is formed by young individuals that reach maturity [2]. In sexual reproduction with selfing allowed, we randomly select two not necessarily different parents, and the single offspring inherits the strategy of either parent with equal probability. This selection process is repeated m𝑚mitalic_m times resulting in the addition of m𝑚mitalic_m offspring to the offspring pool. This mode is clearly equivalent to asexual reproduction, as we will show in the following section. Finally, in the biparental sexual reproduction mode, the two randomly chosen parents must be different and, as before, the single offspring inherits the strategy of either parent with equal probability. In this mode, it is guaranteed that in a group with i=n1𝑖𝑛1i=n-1italic_i = italic_n - 1 cooperators, one cooperator will be chosen as a parent. Note that the groups contribute different numbers of offspring to the offspring pool, not only because m𝑚mitalic_m is a Poisson distributed random variable, but also because the parameter of the Poisson, i.e. π(i)𝜋𝑖\pi(i)italic_π ( italic_i ), depends on the number of cooperators within the group, which varies from group to group. Finally, each group of the next generation is formed by sampling n𝑛nitalic_n individuals with replacement from the offspring pool, and the whole process is repeated until one of the strategies disappears and evolution stops.

In the others-only variant of the n𝑛nitalic_n-person prisoner’s dilemma, altruistic behavior (i.e., contribution to public goods) reduces the fitness of the cooperator, corresponding to the strong altruism scenario. As noted above, the conventional wisdom in this case is that cooperation cannot evolve in randomly formed groups that last less than one generation [24]. The more traditional variant of the n𝑛nitalic_n-person prisoner’s dilemma allows the amplified amount rc𝑟𝑐rcitalic_r italic_c to be redistributed to all players, so that the cooperator gets the return cr/n𝑐𝑟𝑛cr/nitalic_c italic_r / italic_n [10]. If cr/nc>0𝑐𝑟𝑛𝑐0cr/n-c>0italic_c italic_r / italic_n - italic_c > 0 or r>n𝑟𝑛r>nitalic_r > italic_n, then altruistic behavior confers a net benefit to the cooperator, corresponding to the weak altruism scenario. A similar situation occurs in the n𝑛nitalic_n-person snowdrift game when the payoff for completing a task is greater than the effort expended [5]. In this case, the cooperator may have a net fitness advantage over the population as a whole, despite being at a disadvantage relative to defectors within the same group. This is the only situation where Wilson’s trait group formulation [3, 4] or its game theoretic counterparts [10] can explain the maintenance of cooperation [17].

3 Deterministic limit

In his seminal paper [2], Hamilton studied the asexual reproduction mode in the case there are infinitely many groups using Price’s equation [25]. Here we offer an elementary derivation of a recurrence for the frequency of cooperators qtsubscript𝑞𝑡q_{t}italic_q start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT in the offspring pool in generation t𝑡titalic_t, which can be easily generalized to more complex scenarios and reproduction modes [26].

Given qtsubscript𝑞𝑡q_{t}italic_q start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT, the proportion of groups with i𝑖iitalic_i cooperators among the infinitely many groups is given by the binomial distribution

Yi(t)=(ni)qti(1qt)ni,subscript𝑌𝑖𝑡binomial𝑛𝑖superscriptsubscript𝑞𝑡𝑖superscript1subscript𝑞𝑡𝑛𝑖Y_{i}(t)=\binom{n}{i}q_{t}^{i}(1-q_{t})^{n-i},italic_Y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) = ( FRACOP start_ARG italic_n end_ARG start_ARG italic_i end_ARG ) italic_q start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ( 1 - italic_q start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_n - italic_i end_POSTSUPERSCRIPT , (4)

and our problem is to determine how many cooperators, on average, a group with i𝑖iitalic_i cooperators contributes to the offspring pool. We recall that, according to Section 2, a group with i𝑖iitalic_i cooperators contributes m𝑚mitalic_m offspring to the offspring pool with m𝑚mitalic_m distributed by the Poisson distribution

p[m;π(i)]=exp[π(i)][π(i)]mm!.𝑝𝑚𝜋𝑖𝜋𝑖superscriptdelimited-[]𝜋𝑖𝑚𝑚p[m;\pi(i)]=\exp[-\pi(i)]~{}\frac{[\pi(i)]^{m}}{m!}.italic_p [ italic_m ; italic_π ( italic_i ) ] = roman_exp [ - italic_π ( italic_i ) ] divide start_ARG [ italic_π ( italic_i ) ] start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT end_ARG start_ARG italic_m ! end_ARG . (5)

Let Qc(i)subscript𝑄𝑐𝑖Q_{c}(i)italic_Q start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) denote the probability that an offspring is a cooperator. Calculating this probability requires to be explicit about the mode of reproduction, and we will do this calculation separately for each mode in the following, but for now let us assume that Qc(i)subscript𝑄𝑐𝑖Q_{c}(i)italic_Q start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) is known. The probability that there are j𝑗jitalic_j cooperators among the m𝑚mitalic_m offspring produced by a group with i𝑖iitalic_i cooperators is given by the binomial distribution

B[m,Qc(i)]=(mj)[Qc(i)]j[1Qc(i)]mj.𝐵𝑚subscript𝑄𝑐𝑖binomial𝑚𝑗superscriptdelimited-[]subscript𝑄𝑐𝑖𝑗superscriptdelimited-[]1subscript𝑄𝑐𝑖𝑚𝑗B[m,Q_{c}(i)]=\binom{m}{j}\left[Q_{c}(i)\right]^{j}\left[1-Q_{c}(i)\right]^{m-% j}.italic_B [ italic_m , italic_Q start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) ] = ( FRACOP start_ARG italic_m end_ARG start_ARG italic_j end_ARG ) [ italic_Q start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) ] start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT [ 1 - italic_Q start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) ] start_POSTSUPERSCRIPT italic_m - italic_j end_POSTSUPERSCRIPT . (6)

Therefore, a group with i𝑖iitalic_i cooperators contributes, on average, with

m=0p[m;π(i)]j=0mB[m,Qc(i)]j=m=0p[m;π(i)]mQc(i)=π(i)Qc(i)superscriptsubscript𝑚0𝑝𝑚𝜋𝑖superscriptsubscript𝑗0𝑚𝐵𝑚subscript𝑄𝑐𝑖𝑗superscriptsubscript𝑚0𝑝𝑚𝜋𝑖𝑚subscript𝑄𝑐𝑖𝜋𝑖subscript𝑄𝑐𝑖\sum_{m=0}^{\infty}p[m;\pi(i)]~{}\sum_{j=0}^{m}B[m,Q_{c}(i)]j=\sum_{m=0}^{% \infty}p[m;\pi(i)]mQ_{c}(i)=\pi(i)Q_{c}(i)∑ start_POSTSUBSCRIPT italic_m = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_p [ italic_m ; italic_π ( italic_i ) ] ∑ start_POSTSUBSCRIPT italic_j = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_m end_POSTSUPERSCRIPT italic_B [ italic_m , italic_Q start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) ] italic_j = ∑ start_POSTSUBSCRIPT italic_m = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_p [ italic_m ; italic_π ( italic_i ) ] italic_m italic_Q start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) = italic_π ( italic_i ) italic_Q start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) (7)

cooperators to the offspring pool. Note that only the mean of the distribution of the number of offspring m𝑚mitalic_m goes into this calculation, so any distribution with mean π(i)𝜋𝑖\pi(i)italic_π ( italic_i ) gives the same result.

The proportion of cooperators in the offspring pool in generation t+1𝑡1t+1italic_t + 1 is given by the ratio between the mean number of cooperator offspring produced by the groups formed by sampling the offspring pool in generation t𝑡titalic_t and the mean total number of offspring produced by these groups, i.e.

qn+1=i=0nYi(t)π(i)Qc(i)i=0nYi(t)π(i).subscript𝑞𝑛1superscriptsubscript𝑖0𝑛subscript𝑌𝑖𝑡𝜋𝑖subscript𝑄𝑐𝑖superscriptsubscript𝑖0𝑛subscript𝑌𝑖𝑡𝜋𝑖q_{n+1}=\frac{\sum_{i=0}^{n}Y_{i}(t)\pi(i)Q_{c}(i)}{\sum_{i=0}^{n}Y_{i}(t)\pi(% i)}.italic_q start_POSTSUBSCRIPT italic_n + 1 end_POSTSUBSCRIPT = divide start_ARG ∑ start_POSTSUBSCRIPT italic_i = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_Y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) italic_π ( italic_i ) italic_Q start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) end_ARG start_ARG ∑ start_POSTSUBSCRIPT italic_i = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_Y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) italic_π ( italic_i ) end_ARG . (8)

The sum in the denominator can be easily performed and yields n[1+c(r1)qt]𝑛delimited-[]1𝑐𝑟1subscript𝑞𝑡n[1+c(r-1)q_{t}]italic_n [ 1 + italic_c ( italic_r - 1 ) italic_q start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ]. The last step to obtain the recurrence for qtsubscript𝑞𝑡q_{t}italic_q start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT is to calculate Qc(i)subscript𝑄𝑐𝑖Q_{c}(i)italic_Q start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) for the different reproduction modes.

3.1 Asexual reproduction

In this case, we randomly select a parent within a group with i𝑖iitalic_i cooperators according to her relative fitness, and the single offspring inherits her strategy. Thus, the probability of an offspring being a cooperator is simply

Qc(i)=ifc(i)π(i).subscript𝑄𝑐𝑖𝑖subscript𝑓𝑐𝑖𝜋𝑖Q_{c}(i)=\frac{if_{c}(i)}{\pi(i)}.italic_Q start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) = divide start_ARG italic_i italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) end_ARG start_ARG italic_π ( italic_i ) end_ARG . (9)

Inserting this probability into Eq, (8) yields

qt+1=qt1c+rcqt1+cqt(r1),subscript𝑞𝑡1subscript𝑞𝑡1𝑐𝑟𝑐subscript𝑞𝑡1𝑐subscript𝑞𝑡𝑟1q_{t+1}=q_{t}\frac{1-c+rcq_{t}}{1+cq_{t}(r-1)},italic_q start_POSTSUBSCRIPT italic_t + 1 end_POSTSUBSCRIPT = italic_q start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT divide start_ARG 1 - italic_c + italic_r italic_c italic_q start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT end_ARG start_ARG 1 + italic_c italic_q start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ( italic_r - 1 ) end_ARG , (10)

which is consistent with the recurrence derived using Price’s equation, provided that the mean fitness in [2] is corrected to w=1+qt(Kk)=1+qtc(r1)𝑤1subscript𝑞𝑡𝐾𝑘1subscript𝑞𝑡𝑐𝑟1w=1+q_{t}(K-k)=1+q_{t}c(r-1)italic_w = 1 + italic_q start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ( italic_K - italic_k ) = 1 + italic_q start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_c ( italic_r - 1 ). The only fixed points of recurrence (11) are q=0superscript𝑞0q^{*}=0italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0 and q=1superscript𝑞1q^{*}=1italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 1. For qt1much-less-thansubscript𝑞𝑡1q_{t}\ll 1italic_q start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ≪ 1 it is rewritten as qt+1(1c)qtsubscript𝑞𝑡11𝑐subscript𝑞𝑡q_{t+1}\approx(1-c)q_{t}italic_q start_POSTSUBSCRIPT italic_t + 1 end_POSTSUBSCRIPT ≈ ( 1 - italic_c ) italic_q start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT, which implies that qt+1<qtsubscript𝑞𝑡1subscript𝑞𝑡q_{t+1}<q_{t}italic_q start_POSTSUBSCRIPT italic_t + 1 end_POSTSUBSCRIPT < italic_q start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT and so q=0superscript𝑞0q^{*}=0italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0 is always stable and q=1superscript𝑞1q^{*}=1italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 1 is always unstable. This conclusion agrees with Hamilton’s that cooperation cannot progress in such a model [2].

It is instructive to briefly discuss the results for the weak altruism version of the n𝑛nitalic_n-prisoner’s dilemma. In this case, the fitness of a cooperator is f^c(i)=1c+irc/nsubscript^𝑓𝑐𝑖1𝑐𝑖𝑟𝑐𝑛\hat{f}_{c}(i)=1-c+irc/nover^ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) = 1 - italic_c + italic_i italic_r italic_c / italic_n and the fitness of a defector is f^d(i)=1+irc/nsubscript^𝑓𝑑𝑖1𝑖𝑟𝑐𝑛\hat{f}_{d}(i)=1+irc/nover^ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_i ) = 1 + italic_i italic_r italic_c / italic_n, where i𝑖iitalic_i is the number of cooperators in the group of size n𝑛nitalic_n. The fitness of this group is still given by Eq. (3). Thus, the recurrence for the frequency of cooperators q^tsubscript^𝑞𝑡\hat{q}_{t}over^ start_ARG italic_q end_ARG start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT in the weak altruism scenario is

q^t+1=q^t1c+rc/n+(n1)rcq^t/n1+cq^t(r1),subscript^𝑞𝑡1subscript^𝑞𝑡1𝑐𝑟𝑐𝑛𝑛1𝑟𝑐subscript^𝑞𝑡𝑛1𝑐subscript^𝑞𝑡𝑟1\hat{q}_{t+1}=\hat{q}_{t}\frac{1-c+rc/n+(n-1)rc\hat{q}_{t}/n}{1+c\hat{q}_{t}(r% -1)},over^ start_ARG italic_q end_ARG start_POSTSUBSCRIPT italic_t + 1 end_POSTSUBSCRIPT = over^ start_ARG italic_q end_ARG start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT divide start_ARG 1 - italic_c + italic_r italic_c / italic_n + ( italic_n - 1 ) italic_r italic_c over^ start_ARG italic_q end_ARG start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT / italic_n end_ARG start_ARG 1 + italic_c over^ start_ARG italic_q end_ARG start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ( italic_r - 1 ) end_ARG , (11)

whose only fixed points are q^=0superscript^𝑞0\hat{q}^{*}=0over^ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0 and q^=1superscript^𝑞1\hat{q}^{*}=1over^ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 1. For q^t1much-less-thansubscript^𝑞𝑡1\hat{q}_{t}\ll 1over^ start_ARG italic_q end_ARG start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ≪ 1 we have q^t+1(1c+rc/n)q^tsubscript^𝑞𝑡11𝑐𝑟𝑐𝑛subscript^𝑞𝑡\hat{q}_{t+1}\approx(1-c+rc/n)\hat{q}_{t}over^ start_ARG italic_q end_ARG start_POSTSUBSCRIPT italic_t + 1 end_POSTSUBSCRIPT ≈ ( 1 - italic_c + italic_r italic_c / italic_n ) over^ start_ARG italic_q end_ARG start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT so q^=0superscript^𝑞0\hat{q}^{*}=0over^ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0 is unstable (and thus q^=1superscript^𝑞1\hat{q}^{*}=1over^ start_ARG italic_q end_ARG start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 1 is stable) for r>n𝑟𝑛r>nitalic_r > italic_n. This means that cooperators at low frequency can invade a resident population of defectors if the amplification factor r𝑟ritalic_r is sufficiently large.

3.2 Sexual reproduction with self-mating

In this case, we randomly select two parents according to their relative fitness, and the single offspring inherits the strategy of either parent with equal probability. Since there are no restrictions on the choice of parents, the probability that an offspring is a cooperator is

Qc(i)subscript𝑄𝑐𝑖\displaystyle Q_{c}(i)italic_Q start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) =\displaystyle== (ifc(i)π(i))2+12ifc(i)π(i)(ni)fd(i)π(i)+12(ni)fd(i)π(i)ifc(i)π(i)superscript𝑖subscript𝑓𝑐𝑖𝜋𝑖212𝑖subscript𝑓𝑐𝑖𝜋𝑖𝑛𝑖subscript𝑓𝑑𝑖𝜋𝑖12𝑛𝑖subscript𝑓𝑑𝑖𝜋𝑖𝑖subscript𝑓𝑐𝑖𝜋𝑖\displaystyle\left(\frac{if_{c}(i)}{\pi(i)}\right)^{2}+\frac{1}{2}\frac{if_{c}% (i)}{\pi(i)}\frac{(n-i)f_{d}(i)}{\pi(i)}+\frac{1}{2}\frac{(n-i)f_{d}(i)}{\pi(i% )}\frac{if_{c}(i)}{\pi(i)}( divide start_ARG italic_i italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) end_ARG start_ARG italic_π ( italic_i ) end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + divide start_ARG 1 end_ARG start_ARG 2 end_ARG divide start_ARG italic_i italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) end_ARG start_ARG italic_π ( italic_i ) end_ARG divide start_ARG ( italic_n - italic_i ) italic_f start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_i ) end_ARG start_ARG italic_π ( italic_i ) end_ARG + divide start_ARG 1 end_ARG start_ARG 2 end_ARG divide start_ARG ( italic_n - italic_i ) italic_f start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_i ) end_ARG start_ARG italic_π ( italic_i ) end_ARG divide start_ARG italic_i italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) end_ARG start_ARG italic_π ( italic_i ) end_ARG (12)
=\displaystyle== ifc(i)π(i),𝑖subscript𝑓𝑐𝑖𝜋𝑖\displaystyle\frac{if_{c}(i)}{\pi(i)},divide start_ARG italic_i italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) end_ARG start_ARG italic_π ( italic_i ) end_ARG ,

which is identical to the result for asexual reproduction, as expected. Thus, cooperation cannot progress also in the model where sexual reproduction with selfing is allowed.

3.3 Biparental sexual reproduction

The difference to the previous case is that the mating parents within the group of size n𝑛nitalic_n must be different individuals. Their single offspring inherits the strategy of either parent with equal probability. So the problem we have to solve now is to find the probability that the offspring will be a cooperator when the two parents are chosen from a group with i𝑖iitalic_i cooperators and ni𝑛𝑖n-iitalic_n - italic_i defectors. There are three possibilities we need to consider, which are described below.

  1. (a)

    Both parents are cooperators The probability that the first selected parent is a cooperator is simply

    ifc(i)ifc(i)+(ni)fd(i).𝑖subscript𝑓𝑐𝑖𝑖subscript𝑓𝑐𝑖𝑛𝑖subscript𝑓𝑑𝑖\frac{if_{c}(i)}{if_{c}(i)+(n-i)f_{d}(i)}.divide start_ARG italic_i italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) end_ARG start_ARG italic_i italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) + ( italic_n - italic_i ) italic_f start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_i ) end_ARG . (13)

    Given that the first selected parent is a cooperator, the probability that the second is also a cooperator is

    (i1)fc(i)(i1)fc(i)+(ni)fd(i),𝑖1subscript𝑓𝑐𝑖𝑖1subscript𝑓𝑐𝑖𝑛𝑖subscript𝑓𝑑𝑖\frac{(i-1)f_{c}(i)}{(i-1)f_{c}(i)+(n-i)f_{d}(i)},divide start_ARG ( italic_i - 1 ) italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) end_ARG start_ARG ( italic_i - 1 ) italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) + ( italic_n - italic_i ) italic_f start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_i ) end_ARG , (14)

    since the fitness fc(i)subscript𝑓𝑐𝑖f_{c}(i)italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) and fd(i)subscript𝑓𝑑𝑖f_{d}(i)italic_f start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_i ) given by Eqs. (1) and (2) are computed before the selection procedure. Thus, the probability that the two selected parents are cooperators is

    (i1)fc(i)(i1)fc(i)+(ni)fd(i)×ifc(i)ifc(i)+(ni)fd(i).𝑖1subscript𝑓𝑐𝑖𝑖1subscript𝑓𝑐𝑖𝑛𝑖subscript𝑓𝑑𝑖𝑖subscript𝑓𝑐𝑖𝑖subscript𝑓𝑐𝑖𝑛𝑖subscript𝑓𝑑𝑖\frac{(i-1)f_{c}(i)}{(i-1)f_{c}(i)+(n-i)f_{d}(i)}\times\frac{if_{c}(i)}{if_{c}% (i)+(n-i)f_{d}(i)}.divide start_ARG ( italic_i - 1 ) italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) end_ARG start_ARG ( italic_i - 1 ) italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) + ( italic_n - italic_i ) italic_f start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_i ) end_ARG × divide start_ARG italic_i italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) end_ARG start_ARG italic_i italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) + ( italic_n - italic_i ) italic_f start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_i ) end_ARG . (15)

    Of course, the offspring resulting from this mating is a cooperator with probability one. The probabilities of the other events are calculated using the same reasoning.

  2. (b)

    The first chosen parent is a cooperator and the second is a defector. The probability of this event is

    (ni)fd(i)(i1)fc(i)+(ni)fd(i)×ifc(i)ifc(i)+(ni)fd(i)𝑛𝑖subscript𝑓𝑑𝑖𝑖1subscript𝑓𝑐𝑖𝑛𝑖subscript𝑓𝑑𝑖𝑖subscript𝑓𝑐𝑖𝑖subscript𝑓𝑐𝑖𝑛𝑖subscript𝑓𝑑𝑖\frac{(n-i)f_{d}(i)}{(i-1)f_{c}(i)+(n-i)f_{d}(i)}\times\frac{if_{c}(i)}{if_{c}% (i)+(n-i)f_{d}(i)}divide start_ARG ( italic_n - italic_i ) italic_f start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_i ) end_ARG start_ARG ( italic_i - 1 ) italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) + ( italic_n - italic_i ) italic_f start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_i ) end_ARG × divide start_ARG italic_i italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) end_ARG start_ARG italic_i italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) + ( italic_n - italic_i ) italic_f start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_i ) end_ARG (16)

    and the resulting offspring is a cooperator with probability 1/2121/21 / 2.

  3. (c)

    The first chosen parent is a defector and the second is a cooperator. The probability of this event is

    ifc(i)ifc(i)+(ni1)fd(i)×(ni)fd(i)ifc(i)+(ni)fd(i)𝑖subscript𝑓𝑐𝑖𝑖subscript𝑓𝑐𝑖𝑛𝑖1subscript𝑓𝑑𝑖𝑛𝑖subscript𝑓𝑑𝑖𝑖subscript𝑓𝑐𝑖𝑛𝑖subscript𝑓𝑑𝑖\frac{if_{c}(i)}{if_{c}(i)+(n-i-1)f_{d}(i)}\times\frac{(n-i)f_{d}(i)}{if_{c}(i% )+(n-i)f_{d}(i)}divide start_ARG italic_i italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) end_ARG start_ARG italic_i italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) + ( italic_n - italic_i - 1 ) italic_f start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_i ) end_ARG × divide start_ARG ( italic_n - italic_i ) italic_f start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_i ) end_ARG start_ARG italic_i italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) + ( italic_n - italic_i ) italic_f start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_i ) end_ARG (17)

    and the resulting offspring is a cooperator with probability 1/2121/21 / 2.

Combining these probabilities, we obtain that the probability the offspring is a cooperator when two parents are chosen from a group with i𝑖iitalic_i cooperators and ni𝑛𝑖n-iitalic_n - italic_i defectors is

Qc(i)subscript𝑄𝑐𝑖\displaystyle Q_{c}(i)italic_Q start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) =\displaystyle== i(i1)[fc(i)]2[(i1)fc(i)+(ni)fd(i)][ifc(i)+(ni)fd(i)]𝑖𝑖1superscriptdelimited-[]subscript𝑓𝑐𝑖2delimited-[]𝑖1subscript𝑓𝑐𝑖𝑛𝑖subscript𝑓𝑑𝑖delimited-[]𝑖subscript𝑓𝑐𝑖𝑛𝑖subscript𝑓𝑑𝑖\displaystyle\frac{i(i-1)[f_{c}(i)]^{2}}{\left[(i-1)f_{c}(i)+(n-i)f_{d}(i)% \right]\left[if_{c}(i)+(n-i)f_{d}(i)\right]}divide start_ARG italic_i ( italic_i - 1 ) [ italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG [ ( italic_i - 1 ) italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) + ( italic_n - italic_i ) italic_f start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_i ) ] [ italic_i italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) + ( italic_n - italic_i ) italic_f start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_i ) ] end_ARG (18)
+12i(ni)fc(i)fd(i)[(i1)fc(i)+(ni)fd(i)][ifc(i)+(ni)fd(i)]12𝑖𝑛𝑖subscript𝑓𝑐𝑖subscript𝑓𝑑𝑖delimited-[]𝑖1subscript𝑓𝑐𝑖𝑛𝑖subscript𝑓𝑑𝑖delimited-[]𝑖subscript𝑓𝑐𝑖𝑛𝑖subscript𝑓𝑑𝑖\displaystyle+\frac{1}{2}\frac{i(n-i)f_{c}(i)f_{d}(i)}{\left[(i-1)f_{c}(i)+(n-% i)f_{d}(i)\right]\left[if_{c}(i)+(n-i)f_{d}(i)\right]}+ divide start_ARG 1 end_ARG start_ARG 2 end_ARG divide start_ARG italic_i ( italic_n - italic_i ) italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) italic_f start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_i ) end_ARG start_ARG [ ( italic_i - 1 ) italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) + ( italic_n - italic_i ) italic_f start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_i ) ] [ italic_i italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) + ( italic_n - italic_i ) italic_f start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_i ) ] end_ARG
+12i(ni)fc(i)fd(i)[ifc(i)+(ni1)fd(i)][ifc(i)+(ni)fd(i)].12𝑖𝑛𝑖subscript𝑓𝑐𝑖subscript𝑓𝑑𝑖delimited-[]𝑖subscript𝑓𝑐𝑖𝑛𝑖1subscript𝑓𝑑𝑖delimited-[]𝑖subscript𝑓𝑐𝑖𝑛𝑖subscript𝑓𝑑𝑖\displaystyle+\frac{1}{2}\frac{i(n-i)f_{c}(i)f_{d}(i)}{\left[if_{c}(i)+(n-i-1)% f_{d}(i)\right]\left[if_{c}(i)+(n-i)f_{d}(i)\right]}.+ divide start_ARG 1 end_ARG start_ARG 2 end_ARG divide start_ARG italic_i ( italic_n - italic_i ) italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) italic_f start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_i ) end_ARG start_ARG [ italic_i italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) + ( italic_n - italic_i - 1 ) italic_f start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_i ) ] [ italic_i italic_f start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) + ( italic_n - italic_i ) italic_f start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( italic_i ) ] end_ARG .

Clearly, Qc(0)=0subscript𝑄𝑐00Q_{c}(0)=0italic_Q start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( 0 ) = 0 and Qc(n)=1subscript𝑄𝑐𝑛1Q_{c}(n)=1italic_Q start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_n ) = 1. Inserting Qc(i)subscript𝑄𝑐𝑖Q_{c}(i)italic_Q start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) into the recurrence (8) yields

qt+1=Fn(qt)subscript𝑞𝑡1subscript𝐹𝑛subscript𝑞𝑡q_{t+1}=F_{n}(q_{t})italic_q start_POSTSUBSCRIPT italic_t + 1 end_POSTSUBSCRIPT = italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_q start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ) (19)

with

Fn(qt)=1n[1+c(r1)qt]i=0n(ni)qti(1qt)niπ(i)Qc(i),subscript𝐹𝑛subscript𝑞𝑡1𝑛delimited-[]1𝑐𝑟1subscript𝑞𝑡superscriptsubscript𝑖0𝑛binomial𝑛𝑖superscriptsubscript𝑞𝑡𝑖superscript1subscript𝑞𝑡𝑛𝑖𝜋𝑖subscript𝑄𝑐𝑖F_{n}(q_{t})=\frac{1}{n[1+c(r-1)q_{t}]}\sum_{i=0}^{n}\binom{n}{i}q_{t}^{i}(1-q% _{t})^{n-i}\pi(i)Q_{c}(i),italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_q start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ) = divide start_ARG 1 end_ARG start_ARG italic_n [ 1 + italic_c ( italic_r - 1 ) italic_q start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ] end_ARG ∑ start_POSTSUBSCRIPT italic_i = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ( FRACOP start_ARG italic_n end_ARG start_ARG italic_i end_ARG ) italic_q start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ( 1 - italic_q start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_n - italic_i end_POSTSUPERSCRIPT italic_π ( italic_i ) italic_Q start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_i ) , (20)

where the dependence on qtsubscript𝑞𝑡q_{t}italic_q start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT is made explicit. Since Qc(0)=0subscript𝑄𝑐00Q_{c}(0)=0italic_Q start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( 0 ) = 0 we have Fn(0)=0subscript𝐹𝑛00F_{n}(0)=0italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( 0 ) = 0 and so q=0superscript𝑞0q^{*}=0italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0 is a fixed point of the recurrence (19). In addition, noting that Qc(n)=1subscript𝑄𝑐𝑛1Q_{c}(n)=1italic_Q start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_n ) = 1 and π(n)=n[1+c(r1)]𝜋𝑛𝑛delimited-[]1𝑐𝑟1\pi(n)=n[1+c(r-1)]italic_π ( italic_n ) = italic_n [ 1 + italic_c ( italic_r - 1 ) ] we obtain Fn(1)=1subscript𝐹𝑛11F_{n}(1)=1italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( 1 ) = 1, which implies that q=1superscript𝑞1q^{*}=1italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 1 is also a fixed point of the recurrence (19). Whenever these two extreme fixed points are stable, there must also be an unstable coexistence fixed point. If both extreme fixed points are unstable, then the coexistence fixed point is stable, but this situation never happens, as we will see below. We recall that the condition for the local stability of a fixed point is |Fn(q)|<1superscriptsubscript𝐹𝑛superscript𝑞1|F_{n}^{\prime}(q^{*})|<1| italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) | < 1, where the prime indicates the derivative of Fn(q)subscript𝐹𝑛𝑞F_{n}(q)italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_q ), as usual [27].

Before considering the general case of group size n𝑛nitalic_n, it is instructive to write the recurrence (19) for n=2𝑛2n=2italic_n = 2 explicitly. We have

qt+1=qt1+c(r1)qt[1+c(r1)2(1+qt)],subscript𝑞𝑡1subscript𝑞𝑡1𝑐𝑟1subscript𝑞𝑡delimited-[]1𝑐𝑟121subscript𝑞𝑡q_{t+1}=\frac{q_{t}}{1+c(r-1)q_{t}}\left[1+\frac{c(r-1)}{2}(1+q_{t})\right],italic_q start_POSTSUBSCRIPT italic_t + 1 end_POSTSUBSCRIPT = divide start_ARG italic_q start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT end_ARG start_ARG 1 + italic_c ( italic_r - 1 ) italic_q start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT end_ARG [ 1 + divide start_ARG italic_c ( italic_r - 1 ) end_ARG start_ARG 2 end_ARG ( 1 + italic_q start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ) ] , (21)

so that the only fixed points are q=0superscript𝑞0q^{*}=0italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0 and q=1superscript𝑞1q^{*}=1italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 1. Noting that

F2(0)=1+c(r1)2>1superscriptsubscript𝐹201𝑐𝑟121F_{2}^{\prime}(0)=1+\frac{c(r-1)}{2}>1italic_F start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( 0 ) = 1 + divide start_ARG italic_c ( italic_r - 1 ) end_ARG start_ARG 2 end_ARG > 1 (22)

and

F2(1)=112c(r1)[1+c(r1)]<1,superscriptsubscript𝐹21112𝑐𝑟1delimited-[]1𝑐𝑟11F_{2}^{\prime}(1)=1-\frac{1}{2}\frac{c(r-1)}{[1+c(r-1)]}<1,italic_F start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( 1 ) = 1 - divide start_ARG 1 end_ARG start_ARG 2 end_ARG divide start_ARG italic_c ( italic_r - 1 ) end_ARG start_ARG [ 1 + italic_c ( italic_r - 1 ) ] end_ARG < 1 , (23)

we conclude that the all-defectors fixed point is unstable and the all-cooperators fixed point is stable if r>1𝑟1r>1italic_r > 1. This is expected, since the advantage of the selfish strategy comes only from within-group competition, which for n=2𝑛2n=2italic_n = 2 only occurs in groups with one cooperator and one defector. However, the requirement that mating involves both individuals gives Qc(1)=1/2subscript𝑄𝑐112Q_{c}(1)=1/2italic_Q start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( 1 ) = 1 / 2, so this advantage is lost for n=2𝑛2n=2italic_n = 2. Now, if we consider that a group formed by two cooperators contributes on average 2+2c(r1)22𝑐𝑟12+2c(r-1)2 + 2 italic_c ( italic_r - 1 ) offspring to the offspring pool, while a group formed by two defectors contributes on average only 2222 offspring, the dominance of the cooperators is not surprising (see also [28]). However, for groups of size n>2𝑛2n>2italic_n > 2 the situation is not so clear-cut, as we will discuss in the following.

Let us first consider the all-defectors fixed point, q=0superscript𝑞0q^{*}=0italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0, for general n𝑛nitalic_n. A trite calculation yields Fn(0)=π(1)Qc(1)>0superscriptsubscript𝐹𝑛0𝜋1subscript𝑄𝑐10F_{n}^{\prime}(0)=\pi(1)Q_{c}(1)>0italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( 0 ) = italic_π ( 1 ) italic_Q start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( 1 ) > 0, which is rewritten as

Fn(0)=12(1c)[1+n1+rcn1+rccrc/(n1)],superscriptsubscript𝐹𝑛0121𝑐delimited-[]1𝑛1𝑟𝑐𝑛1𝑟𝑐𝑐𝑟𝑐𝑛1F_{n}^{\prime}(0)=\frac{1}{2}(1-c)\left[1+\frac{n-1+rc}{n-1+rc-c-rc/(n-1)}% \right],italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( 0 ) = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( 1 - italic_c ) [ 1 + divide start_ARG italic_n - 1 + italic_r italic_c end_ARG start_ARG italic_n - 1 + italic_r italic_c - italic_c - italic_r italic_c / ( italic_n - 1 ) end_ARG ] , (24)

where we have used Eqs. (3) and (18) with i=1𝑖1i=1italic_i = 1. The condition Fn(0)<1superscriptsubscript𝐹𝑛01F_{n}^{\prime}(0)<1italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( 0 ) < 1 is satisfied when

c>12n3𝑐12𝑛3c>\frac{1}{2n-3}italic_c > divide start_ARG 1 end_ARG start_ARG 2 italic_n - 3 end_ARG (25)

regardless of the value of the amplification factor r𝑟ritalic_r. However, if this inequality is violated, then the condition Fn(0)<1superscriptsubscript𝐹𝑛01F_{n}^{\prime}(0)<1italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( 0 ) < 1 is satisfied, provided that r<rs𝑟subscript𝑟𝑠r<r_{s}italic_r < italic_r start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT where

rs=(n1)2n3c1(2n3)c.subscript𝑟𝑠𝑛12𝑛3𝑐12𝑛3𝑐r_{s}=(n-1)\frac{2n-3-c}{1-(2n-3)c}.italic_r start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = ( italic_n - 1 ) divide start_ARG 2 italic_n - 3 - italic_c end_ARG start_ARG 1 - ( 2 italic_n - 3 ) italic_c end_ARG . (26)

For small c𝑐citalic_c, this equation reduces to rs=(n1)(2n3)+4(n2)(n1)2csubscript𝑟𝑠𝑛12𝑛34𝑛2superscript𝑛12𝑐r_{s}=(n-1)(2n-3)+4(n-2)(n-1)^{2}citalic_r start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = ( italic_n - 1 ) ( 2 italic_n - 3 ) + 4 ( italic_n - 2 ) ( italic_n - 1 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_c. The important result from the ‘problem of origination’ perspective [29] is that the all-defectors fixed point is unstable if inequalities (25) and r<rs𝑟subscript𝑟𝑠r<r_{s}italic_r < italic_r start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT are violated simultaneously. For example, for n=4𝑛4n=4italic_n = 4 the all-defectors fixed point is stable for c>0.2𝑐0.2c>0.2italic_c > 0.2, regardless the value of r𝑟ritalic_r, but for c=0.1𝑐0.1c=0.1italic_c = 0.1 it is unstable if r>29.4𝑟29.4r>29.4italic_r > 29.4. It is clear from these inequalities that increasing the group size n𝑛nitalic_n favors the selfish strategy.

The analysis of the all-cooperators fixed point, q=1superscript𝑞1q^{*}=1italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 1, is more laborious. In this case we have

Fn(1)=nc(r1)1+c(r1)π(n1)Qc(n1)1+c(r1),superscriptsubscript𝐹𝑛1𝑛𝑐𝑟11𝑐𝑟1𝜋𝑛1subscript𝑄𝑐𝑛11𝑐𝑟1F_{n}^{\prime}(1)=n-\frac{c(r-1)}{1+c(r-1)}-\frac{\pi(n-1)Q_{c}(n-1)}{1+c(r-1)},italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( 1 ) = italic_n - divide start_ARG italic_c ( italic_r - 1 ) end_ARG start_ARG 1 + italic_c ( italic_r - 1 ) end_ARG - divide start_ARG italic_π ( italic_n - 1 ) italic_Q start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_n - 1 ) end_ARG start_ARG 1 + italic_c ( italic_r - 1 ) end_ARG , (27)

which can be easily evaluated numerically using Eqs. (3) and (18) to determine the regions in the parameter space where the all-cooperators fixed point is stable, i.e., where the condition |Fn(1)|<1superscriptsubscript𝐹𝑛11|F_{n}^{\prime}(1)|<1| italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( 1 ) | < 1 is satisfied. However, in the limit c0𝑐0c\to 0italic_c → 0 we can write down an explicit expression for Fn(1)superscriptsubscript𝐹𝑛1F_{n}^{\prime}(1)italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( 1 ), viz.,

Fn(1)=1+c2n32(n1)rc12(n1)2superscriptsubscript𝐹𝑛11𝑐2𝑛32𝑛1𝑟𝑐12superscript𝑛12F_{n}^{\prime}(1)=1+c\frac{2n-3}{2(n-1)}-rc\frac{1}{2(n-1)^{2}}italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( 1 ) = 1 + italic_c divide start_ARG 2 italic_n - 3 end_ARG start_ARG 2 ( italic_n - 1 ) end_ARG - italic_r italic_c divide start_ARG 1 end_ARG start_ARG 2 ( italic_n - 1 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG (28)

from which it follows that the all-cooperators fixed point is stable for r>ra𝑟subscript𝑟𝑎r>r_{a}italic_r > italic_r start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT where ra=(n1)(2n3)subscript𝑟𝑎𝑛12𝑛3r_{a}=(n-1)(2n-3)italic_r start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT = ( italic_n - 1 ) ( 2 italic_n - 3 ) in this limit. Note that rarssubscript𝑟𝑎subscript𝑟𝑠r_{a}\to r_{s}italic_r start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT → italic_r start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT for c0𝑐0c\to 0italic_c → 0. Although for c=0𝑐0c=0italic_c = 0 the recurrence (19) reduces to qt+1=qtsubscript𝑞𝑡1subscript𝑞𝑡q_{t+1}=q_{t}italic_q start_POSTSUBSCRIPT italic_t + 1 end_POSTSUBSCRIPT = italic_q start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT so that the dynamics freezes at the initial conditions, a vanishingly small value of c𝑐citalic_c is sufficient to lead the dynamics to q=1superscript𝑞1q^{*}=1italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 1 if r>ra𝑟subscript𝑟𝑎r>r_{a}italic_r > italic_r start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT or to q=0superscript𝑞0q^{*}=0italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0 if r<rs𝑟subscript𝑟𝑠r<r_{s}italic_r < italic_r start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT.

Refer to caption
Figure 1: Phase diagram for n=4𝑛4n=4italic_n = 4 showing the stability regions of the different fixed points. In region C, only the all-cooperators fixed point q=1superscript𝑞1q^{*}=1italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 1 is stable, while in region D, only the all-defectors fixed point q=0superscript𝑞0q^{*}=0italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0 is stable. In region C+D both fixed points are stable. The regions C and C+D are separated by the curve rs=rs(c)subscript𝑟𝑠subscript𝑟𝑠𝑐r_{s}=r_{s}(c)italic_r start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = italic_r start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_c ) and the regions D and C+D are separated by the curve ra=ra(c)subscript𝑟𝑎subscript𝑟𝑎𝑐r_{a}=r_{a}(c)italic_r start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT = italic_r start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( italic_c ).

The results of the stability analysis are conveniently summarized in a phase diagram in the plane (c,r)𝑐𝑟(c,r)( italic_c , italic_r ) as shown in Fig.  1 for n=4𝑛4n=4italic_n = 4. Only the fixed point q=1superscript𝑞1q^{*}=1italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 1 is stable in region C and only the fixed point q=0superscript𝑞0q^{*}=0italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 0 is stable in region D. In region C+D we have the bistability situation, where the outcome of the dynamics depends on the initial frequency of cooperators q0subscript𝑞0q_{0}italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. In particular, the unstable coexistence fixed point 0<q<10superscript𝑞10<q^{*}<10 < italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT < 1 determines the domains of attraction of the two stable fixed points: the dynamics leads to the all-cooperators regime when q0>qsubscript𝑞0superscript𝑞q_{0}>q^{*}italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT > italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT and to the all-defectors regime when q0<qsubscript𝑞0superscript𝑞q_{0}<q^{*}italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT < italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT. An important feature of this phase diagram, which also holds for general group sizes n>2𝑛2n>2italic_n > 2, is that there is no stable coexistence solution. In fact, a stable coexistence fixed point would occur if rs<rasubscript𝑟𝑠subscript𝑟𝑎r_{s}<r_{a}italic_r start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT < italic_r start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT in the limit c0𝑐0c\to 0italic_c → 0, so there would be a region where neither of the extreme fixed points is stable, but we have analytically ruled out this possibility by showing that rarssubscript𝑟𝑎subscript𝑟𝑠r_{a}\to r_{s}italic_r start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT → italic_r start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT in this limit.

Another important limit where we can explicitly rewrite Eq. (27) is for c=1𝑐1c=1italic_c = 1, which corresponds to the worst-case scenario for cooperation and will be the main focus of our analysis henceforth. In this case, we find

Fn(1)=n1+12rn1r(n2)r2(2n310n2+18n11)n1+r(n23n+3),superscriptsubscript𝐹𝑛1𝑛112𝑟𝑛1𝑟𝑛2superscript𝑟22superscript𝑛310superscript𝑛218𝑛11𝑛1𝑟superscript𝑛23𝑛3F_{n}^{\prime}(1)=n-1+\frac{1}{2r}\frac{n-1-r\left(n-2\right)-r^{2}\left(2n^{3% }-10n^{2}+18n-11\right)}{n-1+r\left(n^{2}-3n+3\right)},italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( 1 ) = italic_n - 1 + divide start_ARG 1 end_ARG start_ARG 2 italic_r end_ARG divide start_ARG italic_n - 1 - italic_r ( italic_n - 2 ) - italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 2 italic_n start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT - 10 italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 18 italic_n - 11 ) end_ARG start_ARG italic_n - 1 + italic_r ( italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 3 italic_n + 3 ) end_ARG , (29)

so that the condition |Fn(1)|<1superscriptsubscript𝐹𝑛11|F_{n}^{\prime}(1)|<1| italic_F start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( 1 ) | < 1 is satisfied for r>ra𝑟subscript𝑟𝑎r>r_{a}italic_r > italic_r start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT, where rasubscript𝑟𝑎r_{a}italic_r start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT is the sole positive solution of the quadratic equation

ra2ra(2n27n+6)(n1)=0.superscriptsubscript𝑟𝑎2subscript𝑟𝑎2superscript𝑛27𝑛6𝑛10r_{a}^{2}-r_{a}\left(2n^{2}-7n+6\right)-\left(n-1\right)=0.italic_r start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_r start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( 2 italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 7 italic_n + 6 ) - ( italic_n - 1 ) = 0 . (30)

For large n𝑛nitalic_n we have ra2n2subscript𝑟𝑎2superscript𝑛2r_{a}\approx 2n^{2}italic_r start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ≈ 2 italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, and for the group size n=4𝑛4n=4italic_n = 4 used in drawing the phase diagram in Fig. 1 we have ra10.29subscript𝑟𝑎10.29r_{a}\approx 10.29italic_r start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ≈ 10.29.

Refer to caption
Figure 2: Unstable coexistence fixed point qsuperscript𝑞q^{*}italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT as a function of the amplification factor r𝑟ritalic_r for c=1𝑐1c=1italic_c = 1 and (from left to right) n=3,4,5,6𝑛3456n=3,4,5,6italic_n = 3 , 4 , 5 , 6. If q0>qsubscript𝑞0superscript𝑞q_{0}>q^{*}italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT > italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT the dynamics is driven to the all-cooperators fixed point and to the all-defectors fixed point if q0<qsubscript𝑞0superscript𝑞q_{0}<q^{*}italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT < italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT. The value of r𝑟ritalic_r at which q=1superscript𝑞1q^{*}=1italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 1 signals the instability of the all-cooperators fixed point and is given by the solution of Eq. (30). The all-defectors fixed point is stable for these parameter settings.

Figure 2 shows the unstable coexistence fixed point qsuperscript𝑞q^{*}italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT for c=1𝑐1c=1italic_c = 1 and several group sizes n𝑛nitalic_n. In case of bistability, this fixed point determines the domains of attraction of the two stable fixed points. The coexistence and all-cooperators fixed points coincide (i.e., q=1superscript𝑞1q^{*}=1italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT = 1) at r=ra𝑟subscript𝑟𝑎r=r_{a}italic_r = italic_r start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT, which is given for c=1𝑐1c=1italic_c = 1 by the solution of Eq. (30). We use a log-log scale in this figure to emphasize the power-law decay of qsuperscript𝑞q^{*}italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT with increasing r𝑟ritalic_r. More explicitly, for c=1𝑐1c=1italic_c = 1 and large r𝑟ritalic_r we have an analytical expression for the unstable coexistence fixed point, viz. q(4n6)/rsuperscript𝑞4𝑛6𝑟q^{*}\approx(4n-6)/ritalic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ≈ ( 4 italic_n - 6 ) / italic_r, which is in perfect agreement with the numerical results shown in Fig. 2. This result is important because it shows that at maximum cost to cooperators (i.e., c=1𝑐1c=1italic_c = 1), they can take over the population even at low initial frequencies, provided that the amplification factor r𝑟ritalic_r is sufficiently large.

4 Simulations for finite number of groups

Here we consider the robustness of our deterministic results to demographic noise, which arises when the population is finite, i.e., when G𝐺Gitalic_G is finite. We focus only in the case of biparental sexual reproduction, which can maintain a stable population of cooperators in the limit G𝐺G\to\inftyitalic_G → ∞, as shown in the previous section.

Refer to caption
Figure 3: Frequency of cooperators as a function of time for r=5𝑟5r=5italic_r = 5 and q0=0.5subscript𝑞00.5q_{0}=0.5italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.5 (left panel), r=15𝑟15r=15italic_r = 15, q0=0.5subscript𝑞00.5q_{0}=0.5italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.5 and q0=0.7subscript𝑞00.7q_{0}=0.7italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.7 (middle panel), r=25𝑟25r=25italic_r = 25 and q0=0.5subscript𝑞00.5q_{0}=0.5italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.5 (right panel). The jagged thin black curves are runs of the simulations for G=104𝐺superscript104G=10^{4}italic_G = 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT and the smooth thick colored curves are the deterministic results. The other parameters are n=4𝑛4n=4italic_n = 4 and c=0.05𝑐0.05c=0.05italic_c = 0.05.

To validate our analytical formulation, we compare in Fig. 3 the results obtained by iterating the recurrence (19) and the simulation results for G=104𝐺superscript104G=10^{4}italic_G = 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT and c=0.05𝑐0.05c=0.05italic_c = 0.05. We choose this particular small value of the contribution to the public goods because by varying r𝑟ritalic_r we can go through the three different equilibrium regimes shown in the phase diagram of Fig. 1. In addition to the good agreement between the analytical results and the individual runs of the simulations (the figure shows four runs for each parameter setting), we can see that the fluctuations due to demographic noise are higher in the bistability regime (middle panel of Fig. 3), where the dynamics take a much longer time to reach equilibrium than in the other two regimes. Moreover, the convergence to the all-defectors fixed point (left panel of Fig. 3) is typically much faster than to the all-cooperators fixed point (right panel of Fig. 3). For not too large G𝐺Gitalic_G, in the bistability regime we can have fixation to different fixed points even starting from the same initial condition q0subscript𝑞0q_{0}italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT due to demographic noise. This is the problem we address next, considering a worst-case scenario for cooperation by setting c=1𝑐1c=1italic_c = 1.

To quantify the effect of demographic noise, we consider the probability of fixation of the cooperators ΠΠ\Piroman_Π, which is estimated as the fraction of 104superscript10410^{4}10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT independent runs for which this strategy fixates, i.e., for which the stochastic dynamics is attracted to the all-cooperators absorbing state. In the initial population, each individual is randomly assigned to one of the strategies with equal probability, corresponding to the choice q0=0.5subscript𝑞00.5q_{0}=0.5italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.5 in the deterministic analysis. In the bistability regime, the relevant deterministic result is given in Fig. 2, which can be interpreted in two different ways: for fixed r𝑟ritalic_r there is a minimum initial frequency of cooperators q0=qsubscript𝑞0superscript𝑞q_{0}=q^{*}italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_q start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT above which the recurrence (19) converges to the all-cooperators fixed point, and for fixed q0subscript𝑞0q_{0}italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT there is a minimum value of r=rc𝑟subscript𝑟𝑐r=r_{c}italic_r = italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT above which the recurrence (19) converges to the all-cooperators fixed point. Here we choose the later interpretation and, in particular, for q0=0.5subscript𝑞00.5q_{0}=0.5italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.5 and n=4𝑛4n=4italic_n = 4 we have rc=20.37subscript𝑟𝑐20.37r_{c}=20.37italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 20.37. Accordingly, Fig.  4 shows the probability of fixation of the cooperators as function of the amplification factor r𝑟ritalic_r. As noted above, for the same parameter setting, the demographic noise can tilt the dynamics into one absorbing state or the other. The strength of this effect depends on the number of groups and on the proximity of r𝑟ritalic_r to rcsubscript𝑟𝑐r_{c}italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. In fact, the scaling assumption Π(rrc)G1/2Π𝑟subscript𝑟𝑐superscript𝐺12\Pi\approx(r-r_{c})G^{1/2}roman_Π ≈ ( italic_r - italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) italic_G start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT perfectly describes this dependence for large G𝐺Gitalic_G, implying that the width of the transition region around rcsubscript𝑟𝑐r_{c}italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT shrinks as G1/2superscript𝐺12G^{-1/2}italic_G start_POSTSUPERSCRIPT - 1 / 2 end_POSTSUPERSCRIPT as G𝐺Gitalic_G increases (see [30, 31] for other applications of curve collapsing to characterize threshold phenomena).

Refer to caption
Figure 4: Probability of fixation ΠΠ\Piroman_Π of the cooperators as a function of the amplification factor r𝑟ritalic_r (left panel) for (from bottom to top at r=25𝑟25r=25italic_r = 25) G=400,800,1600,3200𝐺40080016003200G=400,800,1600,3200italic_G = 400 , 800 , 1600 , 3200. The vertical dashed line indicates the threshold rc=20.37subscript𝑟𝑐20.37r_{c}=20.37italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 20.37 beyond which the all-cooperators fixed point attracts the orbits starting at q0=0.5subscript𝑞00.5q_{0}=0.5italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.5 for G𝐺G\to\inftyitalic_G → ∞. The right panel shows ΠΠ\Piroman_Π as a function of the scaled variable (rrc)G1/2𝑟subscript𝑟𝑐superscript𝐺12(r-r_{c})G^{1/2}( italic_r - italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) italic_G start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT for G=800,1600,3200𝐺80016003200G=800,1600,3200italic_G = 800 , 1600 , 3200. The other parameters are n=4𝑛4n=4italic_n = 4 and c=1𝑐1c=1italic_c = 1.

As already hinted at in Fig. 3, another key quantity to characterize the stochastic dynamics is the mean fixation time Tfsubscript𝑇𝑓T_{f}italic_T start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT, i.e., the mean time for the stochastic dynamics to reach the all-cooperators or the all-defectors absorbing states. Figure 5 shows that in the all-cooperators regime (r>rc𝑟subscript𝑟𝑐r>r_{c}italic_r > italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT) the fixation time diverges with aglnGsubscript𝑎𝑔𝐺a_{g}\ln Gitalic_a start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT roman_ln italic_G, where the prefactor ag=ag(n)subscript𝑎𝑔subscript𝑎𝑔𝑛a_{g}=a_{g}(n)italic_a start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT = italic_a start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ( italic_n ) is an increasing function of the group size n𝑛nitalic_n. However, in the all-defectors regime (r<rc𝑟subscript𝑟𝑐r<r_{c}italic_r < italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT) the fixation time tends to a finite limit as G𝐺Gitalic_G increases. Note that for finite G𝐺Gitalic_G, the maximum of Tfsubscript𝑇𝑓T_{f}italic_T start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT does not occur at r=rc𝑟subscript𝑟𝑐r=r_{c}italic_r = italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, but it moves in the direction of the threshold rcsubscript𝑟𝑐r_{c}italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT as G𝐺Gitalic_G increases.

Refer to caption
Figure 5: Mean time for fixation of cooperators or defectors Tfsubscript𝑇𝑓T_{f}italic_T start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT as a function of amplification factor r𝑟ritalic_r (left panel) for n=4𝑛4n=4italic_n = 4 and (from bottom to top at r=25𝑟25r=25italic_r = 25) G=400,800,1600,3200𝐺40080016003200G=400,800,1600,3200italic_G = 400 , 800 , 1600 , 3200. The vertical dashed line indicates the threshold rc=20.37subscript𝑟𝑐20.37r_{c}=20.37italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 20.37 beyond which the all-cooperators fixed point attracts the orbits starting at q0=0.5subscript𝑞00.5q_{0}=0.5italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.5 for G𝐺G\to\inftyitalic_G → ∞. The right panel shows Tfsubscript𝑇𝑓T_{f}italic_T start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT at r=rc𝑟subscript𝑟𝑐r=r_{c}italic_r = italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT as function of G𝐺Gitalic_G for (from bottom to top) n=3,4,5,6𝑛3456n=3,4,5,6italic_n = 3 , 4 , 5 , 6. The lines are the fits Tf=bf+aflnGsubscript𝑇𝑓subscript𝑏𝑓subscript𝑎𝑓𝐺T_{f}=b_{f}+a_{f}\ln Gitalic_T start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = italic_b start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT + italic_a start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT roman_ln italic_G where bfsubscript𝑏𝑓b_{f}italic_b start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT and afsubscript𝑎𝑓a_{f}italic_a start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT are fit parameters. The public goods contribution is c=1𝑐1c=1italic_c = 1.

Figure 6 shows the effect of the group size n𝑛nitalic_n on the threshold transition for finite G𝐺Gitalic_G. Although Fig. 3 already showed that the threshold rcsubscript𝑟𝑐r_{c}italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT increases with n𝑛nitalic_n, we can now see that the width of the transition region also increases with n𝑛nitalic_n, making the fixation of cooperators less likely for r>rc𝑟subscript𝑟𝑐r>r_{c}italic_r > italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT and finite G𝐺Gitalic_G. Indeed, the right panel of Fig. 5 shows that the mean fixation time also increases with n𝑛nitalic_n, confirming that increasing group size makes within-group competition more favorable for defectors, reducing the chances of the cooperators to contribute to the offspring pool. Finally, we note that although in most of the results in this section we have fixed the initial frequency of cooperators at q0=0.5subscript𝑞00.5q_{0}=0.5italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.5, we get qualitatively similar results if we decrease q0subscript𝑞0q_{0}italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, provided that we increase the range of r𝑟ritalic_r according to Fig. 2.

Refer to caption
Figure 6: Probability of fixation of the cooperators as a function of the amplification factor r𝑟ritalic_r for (from left to right at Π=0.8Π0.8\Pi=0.8roman_Π = 0.8) n=3,4,5,6𝑛3456n=3,4,5,6italic_n = 3 , 4 , 5 , 6 (left panel). The vertical dashed lines indicate the thresholds rc(3)=9.62subscript𝑟𝑐39.62r_{c}(3)=9.62italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( 3 ) = 9.62, rc(4)=20.37subscript𝑟𝑐420.37r_{c}(4)=20.37italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( 4 ) = 20.37, rc(5)=34.32subscript𝑟𝑐534.32r_{c}(5)=34.32italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( 5 ) = 34.32 and rc(6)=51.83subscript𝑟𝑐651.83r_{c}(6)=51.83italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( 6 ) = 51.83 beyond which the all-cooperators fixed point attracts the orbits starting at q0=0.5subscript𝑞00.5q_{0}=0.5italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.5 for G𝐺G\to\inftyitalic_G → ∞. The right panel shows the same data with the curves shifted by rc(n)subscript𝑟𝑐𝑛r_{c}(n)italic_r start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_n ). The other parameters are G=400𝐺400G=400italic_G = 400 and c=1𝑐1c=1italic_c = 1.

5 Discussion

There is a consensus that strong altruism, where a cooperator does not benefit from her costly altruistic behavior, cannot evolve in randomly formed, transient groups unless there is some sort of positive assortment among cooperators [2, 3, 4, 24, 32, 33] or punishment for defectors [10, 15, 16]. Note that reputation and reciprocity foster cooperation through the creation of positive assortment among players [34]. Even with these fixes, one difficulty remains: the so-called ‘origination problem’, i.e. how cooperation can take off from an initial low frequency of cooperators. However, it has been argued that this problem is an artifact of assuming, as ours and most models do, that strategies exist as discrete categories rather than as quantitative traits, as is most likely the case [29]. Alas, the quantitative strategy trait fix has only been tested in the case of weak altruism, where a cooperator also benefits from her altruistic behavior and the benefit exceeds the cost of cooperation [29]. In this case, even assuming discrete traits, there are conditions under which defectors can be invaded by rare cooperators, as we showed in Section 3.1. For the others-only variant of the n𝑛nitalic_n-prisoner’s dilemma, and for scenarios of strong altruism in general, the problem of origination remains.

The main purpose of this paper was to highlight the importance that the mode of reproduction can have on the fate of strong altruism in populations that are frequently divided into small temporary groups, in which individuals interact locally and fitness-related processes such as mating take place. The selection dynamics resulting from the different forms of reproduction considered here were solved analytically in the limit of infinitely many groups, assuming, as usual, that individuals colonize the groups according to a binomial distribution, so that cooperators do not cluster into groups, and that local interactions follow the others-only variant of the n𝑛nitalic_n-prisoner’s dilemma. Under biparental sexual reproduction, the results are quite striking, challenging conventional wisdom by showing that strong altruism can indeed evolve in the absence of assortment and punishment in temporary groups if the amplification factor r𝑟ritalic_r is sufficiently large. The basic reason is that biparental sexual reproduction may enable a resident population of cooperators to resist invasion by rare defectors, since the fitness advantage of a single defector in a group with n1𝑛1n-1italic_n - 1 cooperators is smothered by the requirement that the recombinant offspring chromosome has a one-half chance of being a cooperator in any mating involving the defector. However, and most surprisingly, biparental sexual reproduction can also enable the invasion of a resident population of defectors by rare cooperators if the cost of cooperation c𝑐citalic_c is not large, thus solving the problem of origination. These results contrast strikingly with asexual reproduction and sexual reproduction with self-mating modes, for which cooperation is always doomed and the amplification factor r𝑟ritalic_r can only slow the rate at which cooperators disappear from the population [2].

These findings are summarized in Fig. 1, which shows a region, labeled C, where cooperation is the only evolutionary outcome. For large group size n𝑛nitalic_n, this region exists for c<1/2n𝑐12𝑛c<1/2nitalic_c < 1 / 2 italic_n and r>2n2𝑟2superscript𝑛2r>2n^{2}italic_r > 2 italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. The only modification we have made to Hamilton’s original model [2] is to change asexual reproduction to biparental sexual reproduction, which is how most sexual reproduction occurs [20]. This result provides a neat solution to both the problem of origination and the conditions under which strong altruism can evolve, thus avoiding the ‘inverse genetic fallacy’ [35], i.e. the inappropriate attribution of mechanisms that might support cooperation (e.g., positive assortment and punishment) to the explanation of its origin. For large cooperation costs (i.e., c1𝑐1c\approx 1italic_c ≈ 1) and large n𝑛nitalic_n, we find a wide region in the parameter space, labeled C+D in Fig. 1, bounded at the bottom by r=2n2𝑟2superscript𝑛2r=2n^{2}italic_r = 2 italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, where cooperation can evolve provided that q0>4n/rsubscript𝑞04𝑛𝑟q_{0}>4n/ritalic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT > 4 italic_n / italic_r, where q0subscript𝑞0q_{0}italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the initial frequency of cooperators. These estimates for large n𝑛nitalic_n are worst-case scenarios for cooperation, since increasing group size n𝑛nitalic_n favors defection. In addition, we have shown that the results of our analytical study for infinitely many groups are robust to the demographic noise that arises when the number of groups is finite.

How realistic are the conditions in our model for the evolution of cooperation in nature? One possible example that comes to mind is viral co-infection, which complicates the symptoms and diagnosis of disease [36], and also leads to some interesting evolutionary outcomes [37, 38, 39]. Viral co-infections can be modeled by analogy with Wilson’s trait groups [40]. In RNA viruses, recombination can occur via a copy-choice mechanism during viral replication, in which the viral RNA polymerase switches templates during strand synthesis [41]. Virus-virus interactions after co-infection can be modeled using game theory [39], as in the classic example of the RNA phage ϕ6italic-ϕ6\phi 6italic_ϕ 6, where competitive interactions were shown to follow the rules of the prisoner’s dilemma [37]. In this case, the fixation of high-multiplicity phages led to a decrease in population fitness. To escape the dilemma, it has been suggested that fitness payoffs evolve from the prisoner’s dilemma to the snowdrift game, where coexistence between cooperators and defectors is possible [42]. How general this result can be in viral co-infections is unknown, but as shown here, there may be circumstances where the payoffs are consistent with the prisoner’s dilemma and yet cooperation can still evolve.

Finally, it is an open question whether our conclusions apply to diploid sexual organisms under the more realistic assumption that altruistic and selfish strategies are quantitative traits rather than discrete behaviors. As Maynard Smith wrote [43]: “Most populations of interest have sexual diploid inheritance. In many cases this does not matter.” Apart from the obvious case where a pure phenotypic strategy is produced by a heterozygous genotype and cannot be fixed in a random mating population because of Mendelian inheritance, we believe there is no simple answer to the question of whether it does or does not matter.

Acknowledgments

JFF is partially supported by Conselho Nacional de Desenvolvimento Científico e Tecnológico (Brazil), grant number 305620/2021-5. MS is funded by grant PID2021-127107NB-I00 from Ministerio de Ciencia e Innovación (Spain) and grant 2021 SGR 00526 from Generalitat de Catalunya.

Declaration of interest:

None.

References

  • [1] C. Darwin, The Descent of Man, and Selection in Relation to Sex, John Murray, London, 1871.
  • [2] W.D. Hamilton, Innate social aptitudes of man: an approach from evolutionary genetics, in: R. Fox R (ed), ASA Studies 4: Biological Anthropology, Malaby Press, London, 1975, pp. 133–153.
  • [3] D.S. Wilson, A theory of group selection, Proc. Nat. Acad. Sci. USA 72 (1975) 143–146.
  • [4] D.S. Wilson, Weak altruism, strong group selection, Oikos 59 (1990) 135–140.
  • [5] J.F. Fontanari, M. Santos, The dynamics of casual groups can keep free-riders at bay, Math. Biosc. 372 (2024) 109188.
  • [6] J.W. Pepper, Relatedness in trait group models of social evolution, J. Theor. Biol. 206 (2000) 355–368.
  • [7] S.A. West, A.S. Griffin, A. Gardner, Social semantics: altruism, cooperation, mutualism, strong reciprocity and group selection, J. Evol. Biol. 20 (2007) 415–432.
  • [8] S. Okasha, Evolution and the Levels of Selection, Oxford University Press, Oxford, 2009.
  • [9] J. Fox, M. Guyer, Public Choice and cooperation in N-person Prisoner’s Dilemma, J. Conflict Resolut. 22 (1978) 469–481.
  • [10] R. Boyd, R.J. Richerson, Punishment allows the evolution of cooperation (and anything else), in sizable group, Ethol. Sociobiol. 13 (1992) )171–195.
  • [11] E. Fehr, S. Gächter, Cooperation and punishment in public goods experiments, Am. Econ. Rev. 90 (2000) 980–994.
  • [12] D.F. Zheng, H.P. Yin, C.H. Chan, P.M. Hui, Cooperative behavior in a model of evolutionary snowdrift games with N𝑁Nitalic_N-person interactions, Europhys. Lett. 80 (2007) 18002.
  • [13] Santos MD, Pinheiro FL, Santos FC, Pacheco JM (2012) Dynamics of N-person snowdrift games in structured populations. J. Theor. Biol. 315: 81–86.
  • [14] M. Archetti, I. Scheuring, Review: Game theory of public goods in one-shot social dilemmas without assortment, J. Theor. Biol. 299 (2012) 9–20.
  • [15] H. De Silva, C. Hauert, A. Traulsen, K. Sigmund, Freedom, enforcement, and the social dilemma of strong altruism, Evol. Econ. 20 (2010) 203–217.
  • [16] K. Sigmund, H. de Silva, A. Traulsen, C. Hauert, Social learning promotes institutions for governing the commons, Nature 466 (2010) 861–863.
  • [17] B. Charlesworth, A Note on the Evolution of Altruism in Structured Demes, Am. Nat. 113 (1979) 601–630.
  • [18] A. Grafen, Natural selection, kin selection and group selection, in: J. R. Krebs, N. B. Davies (eds), Behavioural Ecology: An Evolutionary Approach, Blackwell Scientific Publications, Oxford, UK, 1984, pp. 62–84.
  • [19] M. Broom, J. Rychtář, Game-Theoretical Models in Biology, CRC Press, Boca Raton, FL, 2013.
  • [20] G.C. Williams, Sex and Evolution, Princeton University Press, Princeton, 1975, p. 115.
  • [21] A. Robinson, Einstein said that - didn’t he? Nature 557 (2018) 30.
  • [22] A. Traulsen, J.C. Claussen, C. Hauert, Coevolutionary Dynamics: From Finite to Infinite Populations, Phys. Rev. Lett. 95 (2005) 238701.
  • [23] J.F. Fontanari, Imitation dynamics and the replicator equation, Europhys. Lett. 146 (2024) 47001.
  • [24] J.A. Fletcher, M. Zwick, Strong altruism can evolve in randomly formed groups. J. Theor. Biol. 228 (2004) 303–313.
  • [25] G.R. Price, Selection and covariance, Nature 227 (1970) 520–521.
  • [26] D. Alves, P.R,A, Campos, A.T.C. Silva, J.F. Fontanari, Group selection models in prebiotic evolution, Phys. Rev. E 63 (2000) 011911.
  • [27] N.F. Britton, Essential Mathematical Biology, Springer, London, 2003.
  • [28] G.C. Williams, D.C. Williams, Natural selection of individually harmful social adaptations among sibs with special reference to social insects, Evolution 11 (1957) 32–39.
  • [29] D.S. Wilson, L. A. Dugatkin, Group selection and assortative interactions, Am. Nat. 149 (1997) 336–351.
  • [30] K. Binder, The Monte Carlo method for the study of phase transitions: A review of some recent progress, J. Comp. Phys. 59 (1985) 1–55.
  • [31] P.R.A. Campos, J.F. Fontanari, Finite-size scaling of the error threshold transition in finite populations, J. Phys. A: Math. Gen. 32 (1999) L1–L7.
  • [32] L. Nunney, Group selection, altruism, and structured-deme models, Am. Nat. 126 (1985) 212–230.
  • [33] J. Maynard Smith, The origin of altruism, Nature 393 (1998) 639–640.
  • [34] C. Xia, J. Wang, M. Perc, Z. Wang, Reputation and reciprocity, Phys. Life Rev. 46 (2023) 8–45.
  • [35] A.J. Field, Altruistically Inclined?: The Behavioral Sciences, Evolutionary Theory, and the Origins of Reciprocity, The University of Michigan Press, Ann Arbor, 2001.
  • [36] Y. Du, C. Wang, Y. Zhang, Viral coinfections, Viruses 14 (2022) 2645.
  • [37] P.E. Turner, L. Chao, Prisoner’s Dilemma in an RNA virus, Nature 398 (1999) 441–443.
  • [38] P.E. Turner, L. Chao, Escape from Prisoner’s Dilemma in RNA phage ϕ6italic-ϕ6\phi 6italic_ϕ 6, Am. Nat. 161 (2003) 497–505.
  • [39] S.F. Elena, G.P. Bernet, J.L. Carrasco, The games plant viruses play, Curr. Opin. Virol. 8 (2014) 62–87.
  • [40] E. Szathmáry, Natural selection and dynamical coexistence of defective and complementing virus segments, J. Theor. Biol. 157(1992) 383–406.
  • [41] T.C. Jarvis, K. Kirkegaard, The polymerase in its labyrinth: mechanisms and implications of RNA recombination, Trends Genet. 7 (1991) 186–191.
  • [42] L. Chao, S.F. Elena, Nonlinear trade-offs allow the cooperation game to evolve from Prisoner’s Dilemma to Snowdrift, Proc. R. Soc. B 284 (2017) 20170228.
  • [43] J. Maynard Smith, Evolution and the Theory of Games, Cambridge University Press, Cambridge, 1982, p. 40.