Skip to main content
Log in

The Unification of Evolutionary Dynamics through the Bayesian Decay Factor in a Game on a Graph

  • Original Article
  • Published:
Bulletin of Mathematical Biology Aims and scope Submit manuscript

Abstract

We unify evolutionary dynamics on graphs in strategic uncertainty through a decaying Bayesian update. Our analysis focuses on the Price theorem of selection, which governs replicator(-mutator) dynamics, based on a stratified interaction mechanism and a composite strategy update rule. Our findings suggest that the replication of a certain mutation in a strategy, leading to a shift from competition to cooperation in a well-mixed population, is equivalent to the replication of a strategy in a Bayesian-structured population without any mutation. Likewise, the replication of a strategy in a Bayesian-structured population with a certain mutation, resulting in a move from competition to cooperation, is equivalent to the replication of a strategy in a well-mixed population without any mutation. This equivalence holds when the transition rate from competition to cooperation is equal to the relative strength of selection acting on either competition or cooperation in relation to the selection differential between cooperators and competitors. Our research allows us to identify situations where cooperation is more likely, irrespective of the specific payoff levels. This approach provides new perspectives into the intended purpose of Price’s equation, which was initially not designed for this type of analysis.

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
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8

Similar content being viewed by others

Data availability

Data sharing does not apply to this article as no new data were created or analyzed in this study.

Notes

  1. Chatterjee and Chakrabartty (2018) demonstrated that various types of evolutionary game dynamics can be considered as specific instances of a dynamical system model derived from a framework of generalized growth transforms.

  2. Previous research by Zhang et al. (2021) has demonstrated that alternative updating rules, such as the delayed group-based sequential myopic best response adjustment rule (MBRAR), have been verified to ensure convergence of the evolutionary game and enhance convergence efficiency.

  3. Weak selection on graphs refers to the situation in which the evolution of a trait on a graph is influenced by weak selective forces. In this context, the fitness of a particular node or edge is determined by its ability to contribute to the spread of the trait through the graph. Even if evolution can be quick, it requires strong selection pressure, which is seldom demonstrated (Kun 2022). This aspect gains particular significance for our investigation as the influence of weak selection escalates when other evolutionary drivers, such as mutation, exert a comparatively substantial impact on the population dynamics (Ueda et al. 2017).

  4. Whenever we discuss defection as an alternative term for competition, it is considered in the context of defecting to cooperate, as described by Nowak and Sigmund (2000). In other words, when players choose to defect, they are essentially engaging in competition.

  5. Considering the variety of terms used in the literature, it is imperative to elucidate that the concept of strategic uncertainty encompasses two distinct components: strategic risk, which pertains to uncertainty about the realization of the opponent’s mixed strategy, and strategic ambiguity, referring to the uncertainty surrounding the selection of a mixed strategy by the opponent (Calford 2020; Dragicevic 2024).

  6. Utilizing the concept of prevalence proportion, it is understood that the probability of selecting an agent at risk from the population corresponds to the population’s at-risk proportion (Dragicevic 2017). From this standpoint, the probability p of engaging with a cooperator in gameplay can be deemed proportional to the cooperator fraction \(x_i\) within the entire population. Nevertheless, a critical distinction emerges between the two: p originates from the context of an N-player game, wherein players are selected from a population characterized by a specific cooperator density \(x_i\). Therefore, although p and \(x_i\) are interrelated, they originate from distinct dimensions of the model’s framework.

  7. Recent advancements in the field of evolutionary dynamics have expanded the scope of inquiry to include not merely pairwise interactions within networks but also those entailing higher-order connectivity. These higher-order interactions, characteristic of evolutionary games, facilitate group-based play rather than the traditional dyadic exchanges. A hallmark of such interactions is the capacity of a single connection, or a hyperlink in this context, to encompass multiple individuals simultaneously, diverging from the conventional network models limited to pairwise links. This architecture permits the modeling of complex group interactions through hyperlinks, effectively capturing the essence of collective play among interconnected participants (Alvarez-Rodriguez et al. 2021).

  8. It is worth noting that at the mixed strategy Nash equilibrium, both players have the same expected payoffs from their respective strategies.

  9. The transition rate, also known as selection intensity or selection strength, represents the rate at which a strategy is replaced by another strategy due to selection pressure in the population. This means that a higher transition rate implies a higher rate of selection, and strategies that are less fit will be replaced more quickly. On the other hand, the mutation probability represents the probability that a strategy will mutate into a different strategy due to a copying error during reproduction. This means that even if a strategy is fit and successful, it can still be replaced by another strategy due to a random mutation event.

  10. It has been established that, in the absence of correlations between hyperdegrees, the dynamics of higher-order-interaction networks align with the replicator dynamics observed in homogeneously mixed populations (Alvarez-Rodriguez et al. 2021).

  11. The value of \(10^{-3}\) is chosen to be small enough to have a negligible impact on the overall result, but large enough to ensure that the denominator is non-zero.

  12. Based on the prevalence proportion principle (Dragicevic 2017), we can determine that the mutation probability of a player chosen at random from the population is equal to the proportion of cooperators in the population. In the case of a population-level mutation from type j to type i, \(E(\pi _{i}\Delta )\) represents the expected surplus from cooperating. This is assuming that the probability of mutation is certain, or in other words, \(E(\Delta )=1\).

  13. \(E(\Delta ) = 1\) implies \(E(\dot{p}) \rightarrow 0\), such that the expected change in p tends toward zero as the player population shifts toward cooperation on a particular basis.

  14. Conversely, if there is a preponderance of cooperators, the mutation is less significant.

  15. The Python codes for the numerical simulations are available upon request for interested readers.

  16. The Python code uses NumPy’s np.linspace and np.meshgrid functions to define the x and y coordinates of the plot. The ax.streamplot function from Matplotlib is used to create a plot of the vector field for the replicator dynamics model. Contour lines are added to the plot using the contourf function to color the regions of the basins of attraction. The set_xlim, set_ylim, set_xmargin, and set_ymargin functions are used to set the axis limits and margins. Finally, the plot is displayed using the plt.show function from Matplotlib.

  17. The colorplan surface plot has been obtained on Python by defining ranges for \(\pi _{i}\), \(\pi _{j}\), and p using NumPy functions, creating a meshgrid, calculating covariance and \(\delta \) values, and then plotting the result using matplotlib.pyplot functions.

  18. The Python code plots a 3D surface of a math function using the func(p,x) function, with p and x values generated by numpy.meshgrid. The imshow function from the matplotlib.pyplot library creates the plot with color scheme set by cmap and a colorbar added using colorbar.

  19. An investigation of fixed points and attractors has been carried out using Python with the aid of two tools: (1) fsolve from the scipy.optimize package and (2) the solve_ivp function from the scipy.integrate module. The analysis was performed using the numpy library, which enabled numerical calculations.

  20. When the transition rate of switching from a competitor to a cooperation is 0.5 in a birth-death process on networks, it indicates that the probability of a neighboring competitor being replaced by a cooperator is equal to the probability of staying in the competition state. This lack of preference or bias towards either state suggests that the system is equally likely to transition to either state.

References

  • Allen B, Nowak M (2014) Games on graphs. EMS Surv Math Sci 1:113–151

    Article  MathSciNet  Google Scholar 

  • Allen B, Rosenbloom D (2012) Mutation Rate Evolution in Replicator Dynamics. Bull Math Biol 74:2650–2675

    MathSciNet  Google Scholar 

  • Alvarez-Rodriguez U, Battiston F, De Arruda GF, Moreno Y, Perc M, Latora V (2021) Evolutionary dynamics of higher-order interactions in social networks. Nat Hum Behav 5:586–595

    Article  Google Scholar 

  • Ariful Kabir K, Shahidul Islam M, Nijhum S (2023) Exploring the performance of volatile mutations on evolutionary game dynamics in complex networks. Heliyon 9:e16790

    Article  Google Scholar 

  • Bach L, Helvik T, Christiansen F (2006) The evolution of N-player cooperation—threshold games and ESS bifurcations. J Theor Biol 238:426–434

    Article  MathSciNet  Google Scholar 

  • Broom M, Křivan V (2020) Two-strategy games with time constraints on regular graphs. J Theor Biol 506:110426

    Article  MathSciNet  Google Scholar 

  • Bürger R (1998) Mathematical principles of mutation-selection models. Genetica 102–103:279–298

    Article  Google Scholar 

  • Calford EM (2020) Uncertainty aversion in game theory: experimental evidence. J Econ Behav Organ 176:720–734

    Article  Google Scholar 

  • Cardinot M, Griffith J, O’Riordan C, Perc M (2018) Cooperation in the spatial Prisoner’s Dilemma game with probabilistic abstention. Sci Rep 8:14531

    Article  Google Scholar 

  • Capraro V, Perc M (2018) Grand challenges in social physics: in pursuit of moral behavior. Front. Phys. 6:107-1–6

  • Chatterjee O, Chakrabartty S (2018) A unified perspective of evolutionary game dynamics using generalized growth transforms. arXiv:1811.02010

  • Chatterjee K, Zufferey D, Nowak M (2012) Evolutionary game dynamics in populations with different learners. J Theor Biol 301:161–173

    Article  MathSciNet  Google Scholar 

  • Chen Y, McAvoy A, Nowak M (2016) Fixation probabilities for any configuration of two strategies on regular graphs. Sci Rep 6:39181

    Article  Google Scholar 

  • Collins S, Gardner A (2009) Integrating physiological, ecological and evolutionary change: a price equation approach. Ecol Lett 12:744–757

    Article  Google Scholar 

  • Cressman R, Tao Y (2014) The replicator equation and other game dynamics. Proc Natl Acad Sci 111:10810–10817

    Article  MathSciNet  Google Scholar 

  • Dercole F, Rinaldi S (2008) Analysis of evolutionary processes: the adaptive dynamics approach and its applications, Princeton series in theoretical and computational biology. Princeton University Press, Princeton

    Google Scholar 

  • De Roos A, Persson L (2005) Unstructured population models: do population-level assumptions yield general theory? Ecological paradigms lost: routes of theory change. Academic Press, Cambridge

    Google Scholar 

  • Dragicevic A (2015) Bayesian population dynamics of spreading species. Environ Model Assess 20:17–27

    Article  MathSciNet  Google Scholar 

  • Dragicevic A (2016) From robustness to resilience: a network price identity approach. Ecol Complex 28:47–53

    Article  Google Scholar 

  • Dragicevic A (2017) Option fund market dynamics for threshold public goods. Dyn Games Appl 7:21–33

    Article  MathSciNet  Google Scholar 

  • Dragicevic A (2019a) Reflective evolution under strategic uncertainty. Int J Bifurc Chaos 1950018-1–14

  • Dragicevic A (2019b) Conditional rehabilitation of cooperation under strategic uncertainty. J Math Biol 79:1973–2003

  • Dragicevic A (2024) The price identity of replicator(-mutator) dynamics on graphs with quantum strategies in a public goods game. Dyn Games Appl. https://doi.org/10.1007/s13235-024-00560-z

    Article  Google Scholar 

  • Duong M, Han T (2020) On equilibrium properties of the replicator-mutator equation in deterministic and random games. Dyn Games Appl 10:641–663

    Article  MathSciNet  Google Scholar 

  • Dyer M, Mohanaraj V (2011) Pairwise-interaction games, automata, languages and programming, vol 6755. Lecture notes in computer science. Springer, Berlin

  • Fehl K, van der Post D, Semmann D (2011) Co-evolution of behaviour and social network structure promotes human cooperation. Ecol Lett 14:546–551

    Article  Google Scholar 

  • Fox J (2006) Using the price equation to partition the effects of biodiversity loss on ecosystem function. Ecology 87:2687–2696

    Article  Google Scholar 

  • Frank S (1995) George Price’s contributions to evolutionary genetics. J Theor Biol 175:373–388

    Article  Google Scholar 

  • Harsanyi J (1973) Games with randomly disturbed payoffs: a new rationale for mixed strategy equilibrium points. Int J Game Theory 2:1–23

    Article  MathSciNet  Google Scholar 

  • Hauert C, De Monte S, Hofbauer J, Sigmund K (2002) Volunteering as Red Queen mechanism for cooperation in public goods games. Science 296:1129–1132

    Article  Google Scholar 

  • Hauert C, Holmes M, Doebeli M (2006) Evolutionary games and population dynamics: maintenance of cooperation in public goods games. Proc R Soc B 273:2565–2570

    Article  Google Scholar 

  • Helanterä H, Uller T (2010) The Price equation and extended inheritance. Philos Theory Biol 2:1–17

    Google Scholar 

  • Hilbe C (2011) Local replicator dynamics: a simple link between deterministic and stochastic models of evolutionary game theory. Bull Math Biol 73:2068–2087

    Article  MathSciNet  Google Scholar 

  • Hilbert M (2017) Complementary variety: when can cooperation in uncertain environments outperform competitive selection? Complexity 5052071:1–15

    Article  MathSciNet  Google Scholar 

  • Hofbauer J, Sigmund K (1998) Evolutionary games and population dynamics. Cambridge University Press, Cambridge

    Book  Google Scholar 

  • Kerr B, Godfrey-Smith P (2008) Generalization of the Price equation for evolutionary change. Evolution 63:531–536

    Article  Google Scholar 

  • Kim J, Chae H, Yook S, Kim Y (2015) Spatial evolutionary public goods game on complete graph and dense complex networks. Sci Rep 5:1–7

    Google Scholar 

  • Knudsen T (2004) General selection theory and economic evolution: the price equation and the replicator/interactor distinction. J Econ Methodol 11:147–173

    Article  Google Scholar 

  • Kun Á (2022) Is there still evolution in the human population? Biologia Futura 73:359–374

    Article  Google Scholar 

  • Kurokawa S (2019) The role of generosity on the evolution of cooperation. Ecol Complex 40:100778

    Article  Google Scholar 

  • Lampert A, Tlusty T (2011) Density-dependent cooperation as a mechanism for persistence and coexistence. Evolution 65:2750–2759

    Article  Google Scholar 

  • Levin S (2002) Complex adaptive systems: exploring the known, the unknown and the unknownable. Bull Am Math Soc 40:3–19

    Article  Google Scholar 

  • Lieberman E, Hauert C, Nowak M (2005) Evolutionary dynamics on graphs. Nature 433:312–316

    Article  Google Scholar 

  • Madgwick P, Wolf J (2020) Evolution of strategic cooperation. Evolut Lett 4:164–175

    Article  Google Scholar 

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

    Article  Google Scholar 

  • McAvoy A, Hauert C (2015) Asymmetric evolutionary games. PLoS Comput Biol 11:e1004349

    Article  Google Scholar 

  • Nolte D (2015) Introduction to modern dynamics, chaos, networks, space and time. Oxford University Press, Oxford

    Google Scholar 

  • Nowak M, May R (1992) Evolutionary games and spatial chaos. Nature 359:826–829

    Article  Google Scholar 

  • Nowak M, Sigmund K (1998) Evolution of indirect reciprocity by image scoring. Nature 393:573–577

    Article  Google Scholar 

  • Nowak M, Sigmund K (2000) Cooperation versus competition. Financ Anal J 56:13–22

    Article  Google Scholar 

  • Nowak M, Tarnita C, Antal T (2010) Evolutionary dynamics in structured populations. Philos Trans R Soc B 365:19–30

    Article  Google Scholar 

  • Odenbaugh J (2005) The structure of population ecology: philosophical reflections on unstructured and structured models, Paradigm’s lost: theory change in ecology. Academic Press, Cambridge

    Google Scholar 

  • Ohtsuki H, Nowak M (2006) The Replicator Equation on Graphs. J Theor Biol 243:86–97

    Article  MathSciNet  Google Scholar 

  • Ohtsuki H, Hauert C, Lieberman E, Nowak M (2006) A simple rule for the evolution of cooperation on graphs and social networks. Nature 441:502–505

    Article  Google Scholar 

  • Okushima T, Niiyama T, Ikeda K, Shimizu Y (2018) Changes of graph structure of transition probability matrices indicate the slowest kinetic relaxations. Phys Rev E 98:032304

    Article  Google Scholar 

  • Page K (2003) Unifying evolutionary dynamics. Mathematical modelling and computing in biology and medicine, Società Editrice Esculapio, Bologna

  • Page K, Nowak M (2002) Unifying evolutionary dynamics. J Theor Biol 219:93–98

    Article  MathSciNet  Google Scholar 

  • Perc M (2006) Coherence resonance in a spatial Prisoner’s Dilemma game. New J Phys 8:22-1–8

  • Perc M (2007) Transition from Gaussian to Lévy distributions of stochastic Payoff variations in the spatial Prisoner’s Dilemma game. Phys Rev E 75:022101-1–4

    Article  Google Scholar 

  • Perc M (2016) Phase transitions in models of human cooperation. Phys Lett A 380:2803–2808

    Article  Google Scholar 

  • Perc M, Marhl M (2006) Evolutionary and dynamical coherence resonances in the pair approximated Prisoner’s Dilemma game. N J Phys 8:142-1–12

  • Perc M, Szolnoki A (2010) Coevolutionary games—a mini review. Biosystems 99:109–125

    Article  Google Scholar 

  • Perc M, Gómez-Gardeñes J, Szolnoki A, Floría L, Moreno Y (2013) Evolutionary Dynamics of Group Interactions on Structured Populations: A Review. J R Soc Interface 10:20120997-1–17

  • Perc M, Jordan J, Rand D, Wang Z, Boccaletti S, Szolnoki A (2017) Statistical physics of human cooperation. Phys Rep 687:1–51

    Article  MathSciNet  Google Scholar 

  • Powers S, Taylor D, Bryson J (2012) Punishment can promote defection in group-structured populations. J Theor Biol 311:107–116

    Article  MathSciNet  Google Scholar 

  • Price G (1970) Selection and covariance. Nature 227:520–521

    Article  Google Scholar 

  • Sandholm W (2012) Evolutionary game theory, computational complexity. Springer, New York

    Google Scholar 

  • Santos F, Santos M, Pacheco J (2008) Social diversity promotes the emergence of cooperation in public goods games. Nature 454:213–216

    Article  Google Scholar 

  • Sato Y, Crutchfield J (2003) Coupled replicator equations for the dynamics of learning in multiagent systems. Phys Rev E 67:015206-1–015206-4

    Article  Google Scholar 

  • Shapley L (1971) Cores of convex games. Int J Game Theory 1:11–26

    Article  MathSciNet  Google Scholar 

  • Szolnoki A, Perc M (2012) Conditional strategies and the evolution of cooperation in spatial public goods games. Phys Rev E 85:026104-1–7

    Article  Google Scholar 

  • Szolnoki A, Perc M (2013) Correlation of positive and negative reciprocity fails to confer an evolutionary advantage: Phase transitions to elementary strategies. Phys Rev X 3:041021-1–11

    Google Scholar 

  • Szolnoki A, Perc M, Szabo G (2012) Defense mechanisms of empathetic players in the spatial ultimatum game. Phys Rev Lett 109:078701-1–4

    Article  Google Scholar 

  • Su Q, Li A, Wang L, Stanley H (2019) Spatial reciprocity in the evolution of cooperation. Proc R Soc B Biol Sci 286:20190041

    Article  Google Scholar 

  • Su Q, Zhou L, Wang L (2019) Evolutionary multiplayer games on graphs with edge diversity. PLoS Comput Biol 15:e1006947

    Article  Google Scholar 

  • Traulsen A, Röhl T, Schuster H (2004) Stochastic gain in population dynamics. Phys Rev Lett 93:028701-1–4

    Article  Google Scholar 

  • Traulsen A, Nowak M (2006) Evolution of cooperation by multilevel selection. Proc Natl Acad Sci 103:10952–10955

    Article  Google Scholar 

  • Ueda M, Takeuchi N, Kaneko K (2017) Stronger selection can slow down evolution driven by recombination on a smooth fitness landscape. PLoS ONE 12:e0183120

    Article  Google Scholar 

  • Waite A, Cannistra C, Shou W (2015) Defectors can create conditions that rescue cooperation. PLoS Comput Biol 11:e1004645

    Article  Google Scholar 

  • Wang J, Fu F, Wu T, Wang L (2009) Emergence of social cooperation in threshold public goods games with collective risk. Phys Rev E 80:016101.1–016101.11

  • Wang Z, Wang L, Szolnoki A, Perc M (2015) Evolutionary games on multilayer networks: a colloquium. Eur Phys J B 88:124-1–15

  • Yagoobi S, Sharma N, Traulsen A (2023) Categorizing update mechanisms for graph-structured metapopulations. J R Soc Interface 20:20220769

    Article  Google Scholar 

  • Yoshimura J, Jansen V (1996) Evolution and population dynamics in stochastic environments. Res Popul Ecol 38:165–182

    Article  Google Scholar 

  • Zhang H, Perc M (2016) Evolution of conditional cooperation under multilevel selection. Sci Rep 6:23006

    Article  Google Scholar 

  • Zhang J, Lou J, Qiu J, Lu J (2021) Dynamics and convergence of hyper-networked evolutionary games with time delay in strategies. Inf Sci 563:166–182

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

The author would like to express gratitude to Sayan Gupta (Center for Complex Systems and Dynamics, IIT Madras), Georges Zaccour (GERAD, HEC Montréal) and Qi Su (Department of Automation, Shanghai Jiao Tong University) for their comprehensive and insightful recommendations. The author is also grateful to the anonymous referees for their thorough comments and suggestions, which significantly contributed in elevating the overall quality of the paper.

Funding

This project received funding from Chulalongkorn University (Grant n\(^{\circ }0325/2566\)).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Arnaud Zlatko Dragicevic.

Ethics declarations

Conflict of interest

The authors have no Conflict of interest to declare.

Additional information

Publisher's Note

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

Appendices

Appendix A

Proof of Proposition 1

We observe that

$$\begin{aligned}{} & {} \pi _i \gtreqless \pi _j \\{} & {} \Leftrightarrow p\left( w(1+\epsilon )-c\right) + (1-p)\left( w-c\right) \gtreqless pw+(1-p)w \\{} & {} \Leftrightarrow pw(1+\epsilon )-pc+w-c-pw+pc \gtreqless pw + w -pw \\{} & {} \Leftrightarrow pw\epsilon -c \gtreqless 0 \\{} & {} \Leftrightarrow p \epsilon w \gtreqless c. \end{aligned}$$

\(\square \)

Appendix B

Proof of Proposition 2

We observe that

$$\begin{aligned}{} & {} \pi _i \gtreqless \pi _j \\{} & {} \Leftrightarrow p\left( w(1+\epsilon )-c\right) x_{i}^{N-1} + (1-p)\left( w-c\right) \gtreqless pw+(1-p)w\left( 1-x_{i}^{N-1}\right) \\{} & {} \Leftrightarrow p(w(1+\epsilon )-c)x_{i}^{N-1} + (1-p)wx_{i}^{N-1} \gtreqless pw +(1-p)w \\{} & {} -(1-p)w +(1-p)c \\{} & {} \Leftrightarrow x_{i}^{N-1} \left( p(w(1+\epsilon )-c)+(1-p)w \right) \gtreqless pw +(1-p)c \\{} & {} \Leftrightarrow x_{i}^{N-1} \gtreqless \frac{c+p(w-c)}{w+p(\epsilon w-c}) \\{} & {} \Leftrightarrow x_{i} \gtreqless \left( \frac{c+p(w-c)}{w+p(\epsilon w-c}) \right) ^{\frac{1}{N-1}} \end{aligned}$$

\(\square \)

Appendix C

Proof of Proposition 3

We analyzed the stability conditions of the system by examining its fixed points,Footnote 19 assuming \(\delta =0.5\).Footnote 20 To accomplish this, we utilized the Jacobian matrix. We solved the system of differential equations \(\dot{E}(p)=\{x_{i}, x_{j}\}\), which yielded the fixed point values \(x_{i}^{\star }=\{0, \frac{-2+15p-15{p^2}-10px_{j}+10{p^2}x_{j}+30{p^2}x_{j}^{99}-20{p^2}x_{j}^{100}}{20p(1-x_{j}^{99}+px_{j}^{99})}\}\) and \(x_{j}^{\star }=\{0, \frac{-2+15p-15{p^2}-10px_{i}+10{p^2}x_{i}+30{p^2}x_{i}^{99}-20{p^2}x_{i}^{100}}{20p(1-x_{i}^{99}+px_{i}^{99})}\}\). Our analysis revealed that the system possesses both a corner equilibrium and an interior equilibrium. However, we found that all fixed points are classified as saddle points due to the eigenvalues of the Jacobian matrix exhibiting opposite signs. At a saddle point, the system displays both stable and unstable behavior in different directions, causing trajectories near the point to either converge or diverge. It is worth noting that saddle points can lead to complex and unpredictable system behavior.

After analyzing the stability of fixed points in the system, we now turn our attention to the presence of attractors. In a system of differential equations, an attractor is a set of states or trajectories that the system tends to converge towards over time, regardless of its initial conditions. Fixed points can also act as attractors in a system. At \(p=0.25\), we observe an attractor with coordinates \(\{x_{i}, x_{j}\}=(0.38,0.38)\). Similarly, at \(p=0.50\) and \(p=0.75\), we have attractors with coordinates \(\{x_{i}, x_{j}\}=(0.3,0.3)\) and \(\{x_{i}, x_{j}\}=(0.13,0.13)\), respectively. At \(p=1\), we observe \(\{x_{i}, x_{j}\}=(0,0)\).

A noteworthy finding is that as the probability of encountering a cooperator increases, the population dynamics system tends to converge towards the interior equilibrium, which approaches zero as p approaches one. The system exhibits a bifurcation around the system attractor point. However, when players are guaranteed to encounter a cooperator, incentivizing free-riding behavior, only the corner equilibrium attracts the system. This results in both populations of players competing against each other, ultimately leading to the extinction of one or both populations. \(\square \)

Appendix D

Proof of Proposition 4

The stability conditions of the system were investigated by analyzing its fixed points, while assuming a value of \(\delta =0.5\). The Jacobian matrix was utilized for this purpose. We solved the differential equations \(\dot{E}(p)=\{x_{i}, x_{j}\}\) and obtained the values of the fixed points as \(x_{i}^{\star }=\{0, \frac{-2x_{j}+15px_{j}-15{p^2}x_{j}-10px_{j}^{3}+10{p^2}x_{j}^{3}+30{p^2}x_{j}^{100}-20{p^2}x_{j}^{102}}{20px_{j}^{2}(1-x_{j}^{99}+px_{j}^{99})}\}\) and \(x_{j}^{\star }=\{0, \frac{-2x_{i}+15px_{i}-15{p^2}x_{i}-10px_{i}^{3}+10{p^2}x_{i}^{3}+30{p^2}x_{i}^{100}-20{p^2}x_{i}^{102}}{20px_{i}^{2}(1-x_{i}^{99}+px_{i}^{99})}\}\). Our analysis indicated that the system possesses both a corner equilibrium and an interior equilibrium. Our analysis indicates that, apart from the coordinates of the attractor, all other fixed points were found to be saddle points, as evidenced by the opposite signs of the eigenvalues of the Jacobian matrix. At a saddle point, the system exhibits both stable and unstable behavior in different directions, leading to trajectories that either converge or diverge near the point. It should be noted that saddle points can result in complex and unpredictable system behavior.

Following our analysis of the stability of fixed points in the system, we now shift our focus to the presence of attractors. In a system of differential equations, an attractor is defined as a set of states or trajectories that the system tends to converge towards over time, regardless of its initial conditions. Fixed points can also serve as attractors in a system. At \(p=0.25\), we observe an attractor with coordinates \(\{x_{i}, x_{j}\}=(0.62,0.62)\). Similarly, at \(p=0.50\) and \(p=0.75\), we have attractors with coordinates \(\{x_{i}, x_{j}\}=(0.55,0.55)\) and \(\{x_{i}, x_{j}\}=(0.37,0.37)\), respectively. At \(p=1\), we observe \(\{x_{i}, x_{j}\}=(0.03,0.03)\).

A key observation is that the population dynamics system typically approaches the interior equilibrium in situations where there is uncertainty about encountering a cooperator. The interior equilibrium approaches zero as p approaches one. The system as a whole exhibits attraction towards a stable attractor point. Nevertheless, when players are guaranteed to interact with cooperators, thereby incentivizing free-riding behavior, only the corner equilibrium attracts the system. This leads to both player populations competing against each other, ultimately resulting in the extinction of one or both populations. \(\square \)

Appendix E

Proof of Proposition 5

The stability conditions of the system were investigated by analyzing its fixed points, while assuming a value of \(\delta =0.5\). We recall that the mutation is considered to occur for sure. The Jacobian matrix was utilized for this purpose. We solved the differential equations \(\dot{E}(p)=\{x_{i}, x_{j}\}\) and obtained the values of the fixed points as \(x_{i}^{\star }=\{0, \frac{-x_{j}+5px_{j}-5p^{2}x_{j}-5px_{j}^{2}+5p^{2}x_{j}^{2}+10p^{2}x_{j}^{100}-10{p^2}x_{j}^{101}}{5p(2x_{j}-3)(1-x_{j}^{99}+px_{j}^{99})}\}\) and \(x_{j}^{\star }=\{0, \frac{-x_{i}+5px_{i}-5p^{2}x_{i}-5px_{i}^{2}+5p^{2}x_{i}^{2}+10p^{2}x_{i}^{100}-10{p^2}x_{i}^{101}}{5p(2x_{i}-3)(1-x_{i}^{99}+px_{i}^{99})}\}\). Our analysis indicated that the system possesses an unstable corner equilibrium, for the eigenvalues of the Jacobian matrix were positive. When a fixed point is unstable, it means that small perturbations from the fixed point will cause the system to move away from the fixed point rather than converge to it. In out case, the unstable fixed point is a source, such that the system exhibits a behavior similar to that of a repeller: trajectories diverge away from the fixed point in all directions.

Following our analysis of the stability of fixed points in the system, we now shift our focus to the presence of attractors. In a system of differential equations, an attractor is defined as a set of states or trajectories that the system tends to converge towards over time, regardless of its initial conditions. In this system, the fixed points do not serve as attractors. At \(p=0.25\), we observe an attractor with coordinates \(\{x_{i}, x_{j}\}=(1,1)\). Similarly, at \(p=0.50\) and \(p=0.75\), we have attractors with coordinates \(\{x_{i}, x_{j}\}=(1,1)\) and \(\{x_{i}, x_{j}\}=(1,1)\), respectively. At \(p=1\), we observe \(\{x_{i}, x_{j}\}=(1,1)\).

The finding of this study indicates that the population dynamics system tends to converge towards an unstable attractor, regardless of the probability of encountering a cooperator. An unstable attractor is a point in a dynamical system that attracts nearby trajectories, but any slight perturbation to the system can cause the trajectories to diverge and move away from it. Assuming a certain probability of mutation towards cooperation during the game, the system continuously eliminates competitors who keep returning as free-riders because they are certain to encounter cooperators, thus driving the system towards a transient full population density. However, when players are guaranteed to interact with cooperators, incentivizing free-riding behavior, only mutation can bring them back to cooperation. \(\square \)

Appendix F

Proof of Proposition 6

The stability of the system was investigated by examining its fixed points, while assuming a fixed value of \(\delta =0.5\) and assuming that mutation towards cooperation always occurs. The Jacobian matrix was utilized to determine the stability conditions. The differential equations \(\dot{E}(p)=\{x_{i}, x_{j}\}\) were solved, and the fixed points were obtained as \(x_{i}^{\star }=\{0, \frac{-x_{j}+5px_{j}-5p^{2}x_{j}-5px_{j}^{3}+5p^{2}x_{j}^{3}+10p^{2}x_{j}^{100}-10{p^2}x_{j}^{102}}{5p(2x_{j}^{2}-3)(1-x_{j}^{99}+px_{j}^{99})}\}\) and \(x_{j}^{\star }=\{0, \frac{-x_{i}+5px_{i}-5p^{2}x_{i}-5px_{i}^{3}+5p^{2}x_{i}^{3}+10p^{2}x_{i}^{100}-10{p^2}x_{i}^{102}}{5p(2x_{i}^{2}-3)(1-x_{i}^{99}+px_{i}^{99})}\}\). Our analysis revealed the presence of an unstable corner equilibrium, as the eigenvalues of the Jacobian matrix were found to be positive. An unstable fixed point implies that small perturbations from the fixed point will cause the system to move away from it instead of converging towards it. In our case, the unstable fixed point is a source, resulting in a behavior similar to that of a repeller, whereby trajectories diverge away from the fixed point in all directions.

After analyzing the stability of the fixed points in the system, we proceed to investigate the presence of attractors. An attractor is a set of states or trajectories to which a system of differential equations converges over time, regardless of its initial conditions. Fixed points can also serve as attractors in a system, but such is not the case in this particular instance. At \(p=0.25\), we observe an attractor located at the coordinates \(\{x_{i}, x_{j}\}=(1,1)\). Similarly, at \(p=0.50\) and \(p=0.75\), we find attractors at \(\{x_{i}, x_{j}\}=(1,1)\) and \(\{x_{i}, x_{j}\}=(1,1)\), respectively. When \(p=1\), the attractor is located at \(\{x_{i}, x_{j}\}=(1,1)\).

The research finding suggests that the population dynamics system tends to converge towards an unstable attractor, regardless of the probability of encountering a cooperator. An unstable attractor is a point in a dynamical system that attracts nearby trajectories, but any slight perturbation to the system may cause the trajectories to diverge and move away from it. Assuming a certain probability of mutation toward cooperation during the game, the system continuously eliminates competitors who persist as free-riders due to their certain encounters with cooperators, driving the system towards a transient full population density. When players are guaranteed to interact with cooperators, which incentivizes free-riding behavior, only mutation can prompt them to revert to cooperation. Interestingly, Bayesian updating has no effect on population dynamics in the case of certain mutation. \(\square \)

Appendix G

Proof of Proposition 7

Through the combination of Eqs. (9) and (10), a general equality can be obtained. It is known that \(\Sigma _{i} px_{i}\pi _{i}=cov(\pi _{i},p) + \bar{\pi }\bar{p}\), where covariance is given by the expression \(cov(\pi _{i},p)=\Sigma _{i} px_{i}\pi _{i}-\bar{\pi }\bar{p}\). The identification of all the covariances that are implicit in the equation allows for the derivation of \(\delta \), which is obtained through the solution of \(cov(\pi _{i},p)\delta = cov(\pi _{j},p)(1+\delta ) \Leftrightarrow \delta = \frac{cov(\pi _{j},p)}{cov(\pi _{i}-\pi _{j},p)}\). In the context of the Price equation, the covariance term represents the impact of natural selection on the alteration in the trait values. Selection is interpreted as the difference between the expected payoff of a model-player and the average payoff within the population. It can be concluded that the equivalence holds when the rate of transition from competition to cooperation corresponds to the relative strength of selection exerted on competition in proportion to the selection differential between the cooperators and competitors. \(\square \)

Appendix H

Proof of Proposition 8

The combination of Eqs. (6) and (11) leads to a general equality. Specifically, it is established that \(\Sigma _{i} px_{i}\pi _{i}=cov(\pi _{i},p) + \bar{\pi }\bar{p}\), where covariance is expressed as \(cov(\pi _{i},p)=\Sigma _{i} px_{i}\pi _{i}-\bar{\pi }\bar{p}\). By identifying the implicit covariances in the equation, we can solve for \(\delta \) using \(cov(\pi _{i},p)(1+\delta ) = cov(\pi _{j},p)\delta \Leftrightarrow \delta = \frac{cov(\pi _{i},p)}{cov(\pi _{j}-\pi _{i},p)}\). In the context of the Price equation, the covariance term indicates the effect of natural selection on changes in trait values. Selection is represented as the difference between the expected payoff of a model-player and the average payoff within the population. The equivalence is valid when the transition rate from competition to cooperation corresponds to the relative strength of selection acting on cooperation relative to the selection differential between competitors and cooperators.\(\square \)

Appendix I

Proof of Corollary 1

In order to determine the range of values for \(\pi _i\) and \(\pi _j\) that satisfy \(x_i^{\star } = \frac{\pi _j}{\pi _j-\pi _i} \in [0,1]\), we must solve the inequalities \(0 \le x_i^{\star } \le 1\). Substituting the expression for \(x_i^{\star }\) into the inequalities, we obtain \(0 \le \frac{\pi _j}{\pi _j-\pi _i} \le 1\). If \(\pi _i > \pi _j\), the denominator is negative, and \(x_i^{\star }\) is undefined. Similarly, if \(\pi _i = \pi _j\), the denominator is zero, and \(x_i^{\star }\) is undefined. Therefore, we require \(\pi _i < \pi _j\) to ensure that \(x_i^{\star } \in [0,1]\). By combining these conditions, we can conclude that \(x_i^{\star } \in [0,1]\) if and only if \(\pi _i=0\). In this case, \(x_i^{\star } =1\), which implies full density of cooperators. While a zero payoff for cooperation might initially discourage cooperative behavior, the full density of cooperators can still emerge under certain conditions in an evolutionary context. This typically requires a sufficiently large population, the ability to adopt a mixed strategy that includes cooperation, and the benefits of mutual cooperation outweighing the costs of being exploited by defectors. This implies that \(w(1+\epsilon )-c=w-c\). This equation leads to the conclusion that the reward for cooperation is zero. Another way to come to this result is by posing \(w(1+\epsilon )-c=0 \Leftrightarrow \epsilon = \frac{c-w}{w}\). Since \(\epsilon \) cannot be negative, it follows that \(w=c\) or \(\epsilon =0\), which suggests that the benefits of mutual cooperation outweigh the costs of being exploited by competitors. Given the conditions described, cooperators are expected to form clusters within the network where the need for rewards is eliminated, as they will interact primarily with other cooperators, thus avoiding exploitation. This phenomenon can only occur when the game is played on a graph with a structured population. In light of the fact that \(x_i^{\star }=1\), we can infer that the entire network can be viewed as a singular, vast cluster of cooperators. \(\square \)

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) 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

Dragicevic, A.Z. The Unification of Evolutionary Dynamics through the Bayesian Decay Factor in a Game on a Graph. Bull Math Biol 86, 69 (2024). https://doi.org/10.1007/s11538-024-01299-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11538-024-01299-9

Keywords

Navigation