Skip to main content
Log in

Average abundancy of cooperation in multi-player games with random payoffs

  • Published:
Journal of Mathematical Biology Aims and scope Submit manuscript

Abstract

We consider interactions between players in groups of size \(d\ge 2\) with payoffs that not only depend on the strategies used in the group but also fluctuate at random over time. An individual can adopt either cooperation or defection as strategy and the population is updated from one time step to the next by a birth-death event according to a Moran model. Assuming recurrent symmetric mutation and payoffs to cooperators and defectors according to the composition of the group whose expected values, variances, and covariances are of the same small order, we derive a first-order approximation for the average abundance of cooperation in the selection-mutation equilibrium. In general, we show that increasing the variance of any payoff for defection or decreasing the variance of any payoff for cooperation increases the average abundance of cooperation. As for the effect of the covariance between any payoff for cooperation and any payoff for defection, we show that it depends on the number of cooperators in the group associated with these payoffs. We study in particular the public goods game, the stag hunt game, and the snowdrift game, all social dilemmas based on random benefit b and random cost c for cooperation, which lead to correlated payoffs to cooperators and defectors within groups. We show that a decrease in the scaled variance of b or c, or an increase in their scaled covariance, makes it easier for weak selection to favor the abundance of cooperation in the stag hunt game and the snowdrift game. The same conclusion holds for the public goods game except that the variance of b has no effect on the average abundance of C. Moreover, while the mutation rate has little effect on which strategy is more abundant at equilibrium, the group size may change it at least in the stag hunt game with a larger group size making it more difficult for cooperation to be more abundant than defection under weak selection.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1

Similar content being viewed by others

References

  • Antal T, Nowak MA, Traulsen A (2009) Strategy abundance in \(2\times 2\) games for arbitrary mutation rates. J Theor Biol 257:340–344

    MATH  Google Scholar 

  • Andrews GE, Askey R, Roy R (1999) Special Functions. Cambridge University Press, New York

    MATH  Google Scholar 

  • Archetti M, Scheuring I (2012) Review: Game theory of public goods in one-shot social dilemmas without assortment. J Theor Biol 299:9–20

    MathSciNet  MATH  Google Scholar 

  • Balakrishnan N, Nevzorov VB (2003) A Primer on Statistical Distributions. John Wiley and Sons, Hoboken, New Jersey

    MATH  Google Scholar 

  • Christiansen FB (1991) On conditions for evolutionary stability for a continuous varying character. Am Nat 138:37–50

    Google Scholar 

  • Conway JH, Guy RK (1995) The Book of Numbers. Copernicus Press, New York

    MATH  Google Scholar 

  • Eshel I (1983) Evolutionary and continuous stability. J Theor Biol 103:99–111

    MathSciNet  Google Scholar 

  • Evans SN, Hening A, Schreiber SJ (2015) Protected polymorphisms and evolutionary stability of patch-selection strategies in stochastic environments. J Math Biol 71:325–359

    MathSciNet  MATH  Google Scholar 

  • Ewens WJ (1972) The sampling theory of selectively neutral alleles. Theor Popul Biol 3:87–112

    MathSciNet  MATH  Google Scholar 

  • Ewens WJ (1990) Population genetics theory - The past and the future. In: Lessard S (ed) Mathematical and Statistical Developments of Evolutionary Theory, NATO ASI Series C: Mathematical and Physical Sciences, vol 299. Kluwer Academic Publishers, Dordrecht, The Netherlands, pp 177–227

    Google Scholar 

  • Ewens WJ (2004) Mathematical Population Genetics 1: Theoretical Introduction. Springer-Verlag, New York

    MATH  Google Scholar 

  • Fox J, Guyer M (1978) Public choice and cooperation in \(N\)-person prisoner’s dilemma. J Conflict Resolut 22:469–481

    Google Scholar 

  • Frank SA, Slatkin M (1990) Evolution in a variable environment. Am Nat 136:244–260

    Google Scholar 

  • Fudenberg D, Harris C (1992) Evolutionary dynamics with aggregate shocks. Journal of Economic Theory 57:420–441

    MathSciNet  MATH  Google Scholar 

  • Fudenberg D, Imhof LA (2006) Imitation processes with small mutations. Journal of Economic Theory 131:251–262

    MathSciNet  MATH  Google Scholar 

  • Gillespie JH (1973) Natural selection with varying selection coefficients - a haploid model. Genet Res 21:115–120

    Google Scholar 

  • Gillespie JH (1974) Natural selection for within-generation variance in offspring number. Genetics 76:601–606

    MathSciNet  Google Scholar 

  • Gokhale C, Traulsen A (2011) Strategy abundance in evolutionary many-player games with multiple strategies. J Theor Biol 283:180–191

    MathSciNet  MATH  Google Scholar 

  • Griffiths RC (1980) Lines of descent in the diffusion approximation of neutral Wright-Fisher models. Theor Popul Biol 17:37–50

    MathSciNet  MATH  Google Scholar 

  • Griffiths RC, Lessard S (2005) Ewens’ sampling formula and related formulae: combinatorial proofs, extensions to variable population size and applications to ages of alleles. Theor Popul Biol 68:167–177

    MATH  Google Scholar 

  • Hamburger H (1973) \(N\)-person prisoner’s dilemma. J Math Sociol 3:27–48

    MathSciNet  MATH  Google Scholar 

  • Hofbauer J, Sigmund K (1988) The Theory of Evolution and Dynamical Systems. Cambridge University Press, Cambridge

    MATH  Google Scholar 

  • Hoppe FM (1984) Polya-like urns and the Ewens’ sampling formula. J Math Biol 20:91–94

    MathSciNet  MATH  Google Scholar 

  • Imhof L, Nowak MA (2006) Evolutionary game dynamics in a Wright-Fisher process. J Math Biol 52:667–681

    MathSciNet  MATH  Google Scholar 

  • Imhof LA (2005) The long-run behavior of the stochastic replicator dynamics. Annals of Applied Probability 15:1019–1045

    MathSciNet  MATH  Google Scholar 

  • Joyce P, Tavaré S (1987) Cycles, permutations and the structures of the Yule process with immigration. Stochastic Processes and its Applications 25:309–314

    MathSciNet  MATH  Google Scholar 

  • Kaplan H, Hill K, Hurtado AM (1990) Risk, foraging and food sharing among the Ache. In: Cashdan E (ed) Risk and Uncertainty in Tribal and Peasant Economies. Westview Press, Boulder, Colorado, pp 107–144

    Google Scholar 

  • Karlin S, Levikson B (1974) Temporal fluctuations in selection intensities: case of small population size. Theor Popul Biol 6:383–412

    MathSciNet  MATH  Google Scholar 

  • Karlin S, Liberman U (1974) Random temporal variation in selection intensities: case of large population size. Theor Popul Biol 6:355–382

    MathSciNet  MATH  Google Scholar 

  • Karlin S, Taylor HM (1975) A First Course in Stochastic Processes, 2nd edn. Academic Press, New York

    MATH  Google Scholar 

  • Kingman JFC (1982) The coalescent. Stochastic Processes and its Applications 13:235–248

    MathSciNet  MATH  Google Scholar 

  • Kroumi D, Lessard S (2015) Evolution of cooperation in a multidimensional phenotype space. Theor Popul Biol 102:60–75

    MATH  Google Scholar 

  • Kroumi D, Lessard S (2015) Strong migration limit for games in structured populations: Applications to dominance hierarchy and set structure. Games 6:318–346

    MathSciNet  MATH  Google Scholar 

  • Kroumi D, Lessard S (2021) The effect of variability in payoffs on average abundance in two-player linear games under symmetric mutation. J Theor Biol 513:110569

    MathSciNet  MATH  Google Scholar 

  • Kroumi D, Martin É, Li C, Lessard S (2021) The effect of variability in payoffs on conditions for the evolution of cooperation in a small population. Dynamic Games and Applications 11:803–834

    MathSciNet  MATH  Google Scholar 

  • Kurokawa S, Ihara Y (2009) Emergence of cooperation in public goods games. Proc R Soc B 276:1379–1384

    Google Scholar 

  • Lande R, Engen S, Saether BE (2003) Stochastic Population Dynamics in Ecology and Conservation. Oxford University Press, Oxford

    MATH  Google Scholar 

  • Lambert A (2006) Probability of fixation under weak selection: a branching process unifying approach. Theor Popul Biol 69:419–441

    MATH  Google Scholar 

  • Li C, Lessard S (2020) Randomized matrix game in a finite population: effect of stochastic fluctuations in the payoffs on the evolution of cooperation. Theor Popul Biol 134:77–91

    MATH  Google Scholar 

  • Lynch M, Walsh B (1998) Genetics and Analysis of Quantitative Traits, 1st edn. Sinauer Associates, Sunderland, Massachusetts

    Google Scholar 

  • May RM (1973) Stability and Complexity in Model Ecosystems. Princeton University Press, Princeton

    Google Scholar 

  • Maynard Smith J (1982) Evolution and the Theory of Games. Cambridge University Press, Cambridge

    MATH  Google Scholar 

  • Maynard Smith J, Price GR (1973) The logic of animal conflict. Nature 246:15–18

    MATH  Google Scholar 

  • Moran PAP (1958) Random processes in genetics. Math Proc Cambridge Philos Soc 54:60–71

    MathSciNet  MATH  Google Scholar 

  • Nowak MA, Sasaki A, Taylor C, Fudenberg D (2004) Emergence of cooperation and evolutionary stability in finite populations. Nature 428:646–650

    Google Scholar 

  • Otto SP, Whitlock MC (1997) The probability of fixation in populations of changing size. Genetics 146:723–733

    Google Scholar 

  • Pacheco JM, Santos FC, Souza MO, Skyrms B (2009) Evolutionary dynamics of collective action in \(N\)-person stag hunt dilemmas. Proc R Soc B 276:1655

    Google Scholar 

  • Parsons TL, Quince C (2007) Fixation in haploid populations exhibiting density dependence II: the non-neutral case. Theor Popul Biol 72:121–135

    MATH  Google Scholar 

  • Parsons TL, Quince C (2007) Fixation in haploid populations exhibiting density dependence II: the quasi-neutral case. Theor Popul Biol 72:468–479

    MATH  Google Scholar 

  • Rice SH, Papadopoulos A (2009) Evolution with stochastic fitness and stochastic migration. PLoS ONE 4:e7130

    Google Scholar 

  • Rychtar J, Taylor DT (2019) Moran process and Wright-Fisher process favor low variability. Discrete and Continuous Dynamical Systems B. https://doi.org/10.3934/dcdsb.2020242

    Article  MATH  Google Scholar 

  • Santos FC, Pacheco JM (2011) Risk of collective failure provides an escape from the tragedy of the commons. Proceedings of the National Academy of Sciences USA 108:10421–10425

    Google Scholar 

  • Schreiber SJ (2012) The evolution of patch selection in stochastic environments. Am Nat 180:17–34

    Google Scholar 

  • Schreiber SJ (2015) Unifying within- and between-generation bet-hedging theories: An ode to J. H. Gillespie. Am Nat 186:792–796

    Google Scholar 

  • Skyrms B (2004) The Stag Hunt and Evolution of Social Structure. Cambridge University Press, Cambridge

    Google Scholar 

  • Souza MO, Pacheco JM, Santos FC (2009) Evolution of cooperation under \(N\)-person snowdrift games. J Theor Biol 260:581–588

    MathSciNet  MATH  Google Scholar 

  • Starrfelt J, Kokko H (2012) Bet-hedging-a triple trade-off between means, variances and correlations. Biol Rev Camb Philos Soc 87:742–55

    Google Scholar 

  • Tavaré S (1984) Line-of-descent and genealogical processes, and their application in population genetics models. Theor Popul Biol 26:119–164

    MathSciNet  MATH  Google Scholar 

  • Taylor P, Jonker L (1978) Evolutionary stable strategies and game dynamics. Math Biosci 40:145–156

    MathSciNet  MATH  Google Scholar 

  • Taylor P (1989) Evolutionary stability in one-parameter models under weak selection. Theor Popul Biol 36:125–143

    MathSciNet  MATH  Google Scholar 

  • Uecker H, Hermisson J (2011) On the fixation process of a beneficial mutation in a variable environment. Genetics 188:915–930

    Google Scholar 

  • Zeeman RC (1980) Populations dynamics from game theory. In: Nitecki ZH, Robinson RC (eds) Global Theory of Dynamical Systems, Lecture Notes in Mathematics, Vol 819, Springer, New York

  • Zheng DF, Yin H, Chan CH, Hui PM (2007) Cooperative behavior in a model of evolutionary snowdrift games with \(N\)-person interactions. Europhys Lett 80:18002

    MathSciNet  Google Scholar 

  • Zheng XD, Li C, Lessard S, Tao Y (2017) Evolutionary stability concepts in a stochastic environment. Phys Rev E 96:032414

    Google Scholar 

  • Zheng XD, Li C, Lessard S, Tao Y (2018) Environmental noise could promote stochastic local stability of behavioral diversity evolution. Phys Rev Lett 120:218101

    Google Scholar 

Download references

Acknowledgements

S. Lessard is supported by NSERC of Canada, grant no. 8833.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Dhaker Kroumi.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A: Conditional expected frequency change

For \(i/N=x\), we have

$$\begin{aligned} \frac{\left( {\begin{array}{c}i-1\\ k\end{array}}\right) \left( {\begin{array}{c}N-i\\ n-k\end{array}}\right) }{\left( {\begin{array}{c}N-1\\ n\end{array}}\right) }&=\left( {\begin{array}{c}n\\ k\end{array}}\right) \frac{(i-1)\times \cdots \times (i-k)\times (N-i)\times \cdots \times (N-i-n+k+1)}{(N-1)\times \cdots \times (N-n)}\nonumber \\&=\left( {\begin{array}{c}n\\ k\end{array}}\right) \frac{(x-\frac{1}{N})\times \cdots \times (x-\frac{k}{N})\times (1-x)\times \cdots \times (1-x-\frac{n-k-1}{N})}{(1-\frac{1}{N})\times \cdots \times (1-\frac{n}{N})}\nonumber \\&=\left( {\begin{array}{c}n\\ k\end{array}}\right) x^{k}(1-x)^{n-k}+O(N^{-1}). \end{aligned}$$
(75)

Therefore, in the limit of a large population size N, the average payoffs to C and D in (3) can be written as

$$\begin{aligned} P_C(x)&=\sum _{k=0}^{d-1}\left( {\begin{array}{c}d-1\\ k\end{array}}\right) x^k(1-x)^{d-k-1}a_k, \end{aligned}$$
(76a)
$$\begin{aligned} P_D(x)&=\sum _{k=0}^{d-1}\left( {\begin{array}{c}d-1\\ k\end{array}}\right) x^k(1-x)^{d-k-1}b_k. \end{aligned}$$
(76b)

Using (1), the first two moments of the average payoff to C in (3) when the frequency of C is x are given by

$$\begin{aligned} E\Big [P_C(x)\Big ]&=\delta \sum _{k=0}^{d-1}\left( {\begin{array}{c}d-1\\ k\end{array}}\right) x^k(1-x)^{d-k-1}\mu _{C,k}+o(\delta ), \end{aligned}$$
(77a)
$$\begin{aligned} E\Big [P_C^2(x)\Big ]&=\delta \sum _{k,l=0}^{d-1}\left( {\begin{array}{c}d-1\\ k\end{array}}\right) \left( {\begin{array}{c}d-1\\ l\end{array}}\right) x^{k+l}(1-x)^{2d-k-l-2}\sigma _{CC,kl}+o(\delta ), \end{aligned}$$
(77b)

and similarly the first two moments of \(P_D(x)\) by

$$\begin{aligned} E\Big [P_D(x)\Big ]&=\delta \sum _{k=0}^{d-1}\left( {\begin{array}{c}d-1\\ k\end{array}}\right) x^k(1-x)^{d-k-1}\mu _{D,k}+o(\delta ), \end{aligned}$$
(78a)
$$\begin{aligned} E\Big [P_D^2(x)\Big ]&=\delta \sum _{k,l=0}^{d-1}\left( {\begin{array}{c}d-1\\ k\end{array}}\right) \left( {\begin{array}{c}d-1\\ l\end{array}}\right) x^{k+l}(1-x)^{2d-k-l-2}\sigma _{DD,kl}+o(\delta ). \end{aligned}$$
(78b)

Moreover, we have

$$\begin{aligned} E\Big [P_C(x)P_D(x)\Big ]\!=\!\delta \sum _{k,l=0}^{d-1}\left( {\begin{array}{c}d-1\\ k\end{array}}\right) \left( {\begin{array}{c}d-1\\ l\end{array}}\right) x^{k+l}(1\!-\!x)^{2d-k-l-2}\sigma _{CD,kl}\!+\!o(\delta ).\qquad \end{aligned}$$
(79)

We are interested in

$$\begin{aligned} E\left[ \frac{f_C(x)-f_D(x)}{xf_C(x)+(1-x)f_D(x)}\right] =E\left[ \frac{P_C(x)-P_D(x)}{1+{\bar{P}}(x)}\right] , \end{aligned}$$
(80)

where \({\bar{P}}(x)=xP_C(x)+(1-x)P_D(x)\) is the average payoff in the population. We expand the last expression by the delta-method (Lynch and Walsh 1998; Rice and Papadopoulos 2009), which gives

(81)

for two random variables Y and Z with

(82)

and

(83)

for \(k \ge 1\). For \(Y=P_C(x)-P_D(x)\) and \(Z=1+{\bar{P}}(x)\), using the facts that

(84)

for \(k\ge 2\) and

(85)

for \(k\ge 1\), we obtain

(86)

Note that

(87)

and \(\left( 1+E[{\bar{P}}(x)]\right) ^{2}=1+O(\delta )\), so that

$$\begin{aligned} E\left[ \frac{P_C(x)-P_D(x)}{1+{\bar{P}}(x)}\right]&=E[P_C(x)]-E[P_D(x)] -E\left[ \left( P_C(x)-P_D(x)\right) {\bar{P}}(x)\right] +o(\delta )\nonumber \\&=E[P_C(x)]-E[P_D(x)]-xE[P^2_1(x)]+(2x-1)E[P_C(x)P_D(x)]\nonumber \\&\quad +(1-x)E[P^2_2(x)]+o(\delta )\nonumber \\&=m(x)\delta +o(\delta ), \end{aligned}$$
(88)

where

$$\begin{aligned} m(x)&=\sum _{k=0}^{d-1}\left( {\begin{array}{c}d-1\\ k\end{array}}\right) x^k(1-x)^{d-k-1}\left( \mu _{C,k}-\mu _{D,k}\right) +\sum _{k,l=0}^{d-1}\left( {\begin{array}{c}d-1\\ k\end{array}}\right) \left( {\begin{array}{c}d-1\\ l\end{array}}\right) \nonumber \\&\quad \times x^{k+l}(1-x)^{2d-k-l-2}\Big [-x\sigma _{CC,kl}+(1-x)\sigma _{DD,kl}+((x-(1-x))\sigma _{CD,kl}\Big ]. \end{aligned}$$
(89)

Appendix B: Moments of the Dirichlet distribution

The ancestry of a random sample of n individuals is described backward in time by a coalescent tree with every pair of lines at any given time back coalescing at rate 1 independently of all the others (Kingman 1982). Moreover, mutations occur independently on each lines of the coalescent tree according to a Poisson process of intensity \(\theta >0\). When there is mutation, the mutant type is 1 or 2 with probability 1/2 for each type independently of the parental type. A line is said to be ancestral to the sample as long as no mutation has occurred on it. We are interested in the probability distribution of the sample configuration, more precisely the probability for k labeled individuals to be of type 1 and the \(n-k\) others to be of type 2, denoted by \(\psi _n^k\), for \(0\le k\le n\).

In order to determine the sample probability distribution, we will extend an approach used in Griffiths and Lessard (2005) to show the Ewens sampling formula (Ewens 1972) in the case where the mutant type is always a novel type. See also Hoppe (1984) and Joyce and Tavaré (1987) for related approaches based on urn models and cycles in permutations.

Note first that the number of ancestral lines of a sample of n individuals is a death process backward in time, where ancestral lines are lost by either mutation or coalescence. This death process was studied in Griffiths (1980) and Tavaré (1984), and the events in this death process called defining events in Ewens (1990).

Label the n sampled individuals and list them in the order in which their ancestral lines are lost backward in time, following either a mutation or a coalescence. In the case of coalescence, one of the two lines involved is chosen at random to be the one that is lost, the other one being the continuing line. There are n! different orders.

Let us first consider the probability for the n sampled individuals in a given order to be all of type 1. Note that this event occurs if and only if all ancestral lines lost by mutation lead to type 1. Now let us look at the probability of each defining event. When i ancestral lines remain, the total rate of mutation is \(i\theta \) and the total rate of coalescence is \(i(i-1)/2\), while the rate of mutation leading to type 1 on any particular ancestral line is \(\theta /2\) and the rate of coalescence involving any particular ancestral line and leading to its loss is \((i-1)/2\). Therefore, the probability that a particular ancestral line is the next one lost and that it is lost by mutation leading to type 1 is \((\theta /2)/[(i\theta +i(i-1)/2]\) for \(i\ge 1\). Similarly, the probability that a particular ancestral line is the next one lost and that it is lost by coalescence is \([(i-1)/2]/[(i\theta +i(i-1)/2]\) for \(i\ge 1\). Summing these probabilities, multiplying the sums for \(i=n, n-1, \ldots , 1\), and considering all possible orders, we get

$$\begin{aligned} \psi _n^0&=n! \, \prod _{i=1}^{n} \left( \frac{\theta /2 + (i-1)/2}{i\theta +i(i-1)/2} \right) = \frac{\prod _{i=1}^{n}(\theta + i-1)}{ \prod _{i=1}^{n}(2\theta +i-1)} \end{aligned}$$
(90)

as probability for the n sampled individuals to be of type 1.

Now let us look at the general case of k labeled individuals of type 1 and \(n-k\) of type 2. When i ancestral lines of individuals of type 1 and j ancestral lines of individuals of type 2 remain, the total rate of mutation is \((i+j)\theta \) and the total rate of coalescence is \((i+j)(i+j-1)/2\), while any particular ancestral line of an individual is lost by mutation to the type of the individual, whose rate is \(\theta /2\), or by coalescence with ancestral lines of individuals of the same type, whose rate is \((i-1)/2\) for type 1 and \((j-1)/2\) for type 2. Considering i and j decreasing from k and \(n-k\) to 0 or 1 with \(i+j=n, n-1, \ldots , 1\), we get

$$\begin{aligned} \psi _n^k&=n! \, \frac{\prod _{i=1}^{k}(\theta /2 + (i-1)/2) \prod _{j=1}^{n-k}(\theta /2 + (j-1)/2)}{ \prod _{l=1}^{n}(l\theta +l(l-1)/2)} \nonumber \\&=\frac{\prod _{i=1}^{k}(\theta +i-1)\prod _{j=1}^{n-k}(\theta +j-1)}{\prod _{l=1}^{n}(2\theta +l-1)}\nonumber \\&=\frac{\Gamma (2\theta )}{ \Gamma (2\theta +n) } \left( \frac{\Gamma (\theta +k)\Gamma (\theta +n-k)}{ \Gamma (\theta )^2}\right) \end{aligned}$$
(91)

as probability for k labeled individuals to be of type 1 and the \(n-k\) others to be of type 2 in a random sample of size n. Here, we use \(\Gamma (\beta +1 )= \beta \Gamma (\beta )\) for \(\beta >0\). Note that \(\psi _n^k=\psi _{n-k}^k\).

Applying the same approach as above for K types with rate of mutation to type k given by \(\alpha _k/2 >0\) for \(k=1, \ldots , K\) with \(\alpha _1 + \cdots + \alpha _K=\alpha \), the probability for \(n_k\) labeled individuals to be of type k for \(k=1, \ldots , K\) in a random sample of size \(n=n_1 + \cdots + n_K\) is given by

$$\begin{aligned} \psi _n^{n_1, \ldots , n_K}&= \frac{\prod _{k=1}^{K}\prod _{i_k=1}^{n_k}(\alpha _k + i_k-1)}{ \prod _{i=1}^{n}(\alpha +i -1)} \nonumber \\&=\frac{\Gamma (\alpha )}{\Gamma (\alpha + n)}\prod _{k=1}^{K} \frac{\Gamma (\alpha _k + n_k)}{ \Gamma (\alpha _k)}. \end{aligned}$$
(92)

These are the moments of the Dirichlet distribution (see, e.g., Balakrishnan and Nevzorov 2003).

Rights and permissions

Springer Nature or its licensor holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kroumi, D., Lessard, S. Average abundancy of cooperation in multi-player games with random payoffs. J. Math. Biol. 85, 27 (2022). https://doi.org/10.1007/s00285-022-01789-1

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00285-022-01789-1

Keywords

Mathematics Subject Classification

Navigation