Skip to main content

Advertisement

Log in

Extinction scenarios in evolutionary processes: a multinomial Wright–Fisher approach

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

Abstract

We study a discrete-time multi-type Wright–Fisher population process. The mean-field dynamics of the stochastic process is induced by a general replicator difference equation. We prove several results regarding the asymptotic behavior of the model, focusing on the impact of the mean-field dynamics on it. One of the results is a limit theorem that describes sufficient conditions for an almost certain path to extinction, first eliminating the type which is the least fit at the mean-field equilibrium. The effect is explained by the metastability of the stochastic system, which under the conditions of the theorem spends almost all time before the extinction event in a neighborhood of the equilibrium. In addition to the limit theorems, we propose a maximization principle for a general deterministic replicator dynamics and study its implications for the stochastic model.

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

Similar content being viewed by others

References

  • Akhmet M, Fen MO (2016) Input-output mechanism of the discrete chaos extension. In: Afraimovich V, Machado JAT, Zhang J (eds) Complex motions and chaos in nonlinear systems, Nonlinear Syst. Complex., vol 15. Springer, Berlin, pp 203–233

  • Antal T, Traulsen A, Ohtsuki H, Tarnita CE, Nowak MA (2009) Mutation-selection equilibrium in games with multiple strategies. J Theor Biol 258:614–622

    MathSciNet  MATH  Google Scholar 

  • Ao P (2005) Laws in Darwinian evolutionary theory. Phys Life Rev 2:117–156

    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 

  • Arnoldt H, Timme M, Grosskinsky S (2012) Frequency-dependent fitness induces multistability in coevolutionary dynamics. J R Soc Interface 9:3387–3396

    Google Scholar 

  • Ashcroft P, Michor F, Galla T (2015) Stochastic tunneling and metastable states during the somatic evolution of cancer. Genetics 199:1213–1228

    Google Scholar 

  • Assaf M, Meerson B (2008) Noise enhanced persistence in a biochemical regulatory network with feedback control. Phys Rev Lett 100:058105

    Google Scholar 

  • Assaf M, Meerson B (2010) Extinction of metastable stochastic populations. Phys Rev E 81:021116

    Google Scholar 

  • Assaf M, Mobilia M (2011) Fixation of a deleterious allele under mutation pressure and finite selection intensity. Phys Rev E 275:93–103

    MathSciNet  MATH  Google Scholar 

  • Auletta V, Ferraioli D, Pasquale F, Persiano G (2018) Metastability of logit dynamics for coordination games. Algorithmica 80:3078–3131

    MathSciNet  MATH  Google Scholar 

  • Aydogmus O (2016) On extinction time of a generalized endemic chain-binomial model. Math Biosci 279:38–42

    MathSciNet  MATH  Google Scholar 

  • Banks J, Brooks J, Cairns G, Davis G, Stacey P (1992) On Devaney’s definition of chaos. Am Math Mon 99:332–334

    MathSciNet  MATH  Google Scholar 

  • Barton NH, Coe JB (2009) On the application of statistical physics to evolutionary biology. J Theor Biol 259:317–324

    MathSciNet  MATH  Google Scholar 

  • Barton JP, Goonetilleke N, Butler TC, Walker BD, McMichael AJ, Chakraborty AK (2016) Relative rate and location of intra-host HIV evolution to evade cellular immunity are predictable. Nat Commun 7:11660

    Google Scholar 

  • Basener WF (2013) Limits of chaos and progress in evolutionary dynamics, In: Marks II RJ, Behe MJ, Dembski WA, Gordon BL, Sanford JC (eds) Biological information. New Perspectives, Proceedings of the Symposium (Cornell University, 2011). World Scientific, Singapore, pp 87–104

  • Basener WF, Sanford JC (2018) The fundamental theorem of natural selection with mutations. J Math Biol 76:1589–1622

    MathSciNet  MATH  Google Scholar 

  • Beerenwinkel N, Antal T, Dingli D, Traulsen A, Kinzler KW, Velculescu VE, Vogelstein B, Nowak MA (2007) Genetic progression and the waiting time to cancer. PLoS Comput Biol 3:2239–2246

    MathSciNet  Google Scholar 

  • Benaïm M, Schreiber SJ (2019) Persistence and extinction for stochastic ecological models with internal and external variables. J Math Biol 79:393–431

    MathSciNet  MATH  Google Scholar 

  • Bernardi O, Florio A (2019) Existence of Lipschitz continuous Lyapunov functions strict outside the strong chain recurrent set. Dyn Syst 34:71–92

    MathSciNet  MATH  Google Scholar 

  • Birch J (2016) Natural selection and the maximization of fitness. Biol Rev 91:712–727

    Google Scholar 

  • Block GL, Allen LJS (2000) Population extinction and quasi-stationary behavior in stochastic density-dependent structured models. Bull Math Biol 62:199–228

    MATH  Google Scholar 

  • Block L, Franke JE (1985) The chain recurrent set, attractors, and explosions. Ergodic Theory Dyn Syst 5:321–327

    MathSciNet  MATH  Google Scholar 

  • Boenkost F, González Casanova A, Pokalyuk C, Wakolbinger A (2019) Haldane’s formula in Cannings models: the case of moderately weak selection, preprint is available at arXiv:1907.10049

  • Bof N, Carli R, Schenato L (2018) Lyapunov theory for discrete time systems, technical report is available at arXiv:1809.05289

  • Bovier A, den Hollander F (2015) Metastability: a potential-theoretic approach. Springer, Berlin

    MATH  Google Scholar 

  • Bratus AS, Semenov YS, Novozhilov AS (2018) Adaptive fitness landscape for replicator systems: to maximize or not to maximize. Math Model Nat Phenom 13:25

    MathSciNet  MATH  Google Scholar 

  • Brémaud P (1998) Markov Chains: Gibbs fields, Monte Carlo simulation, and queues. Springer, Berlin

    MATH  Google Scholar 

  • Broekman MJE, Muller-Landau HC, Visser MD, Jongejans E, Wright SJ, de Kroon H (2019) Signs of stabilisation and stable coexistence. Ecol Lett 22:1957–1975

    Google Scholar 

  • Broom M, Rychtář J (2013) Game-Theoretical Models in Biology. CRC Press, Cambridge

    MATH  Google Scholar 

  • Buckley FM, Pollett PK (2010) Limit theorems for discrete-time metapopulation models. Probab Surv 7:53–83

    MathSciNet  MATH  Google Scholar 

  • Cai J, Tan T, Chan SHJ (2019) Bridging traditional evolutionary game theory and metabolic models for predicting Nash equilibrium of microbial metabolic interactions, preprint is available at https://www.biorxiv.org/content/early/2019/05/14/623173.full.pdf

  • Carja O, Liberman U, Feldman MW (2014) Evolution in changing environments: modifiers of mutation, recombination, and migration. Proc Natl Acad Sci USA 111:17935–17940

    Google Scholar 

  • Caulin AF, Graham TA, Wang LS, Maley CC (2015) Solutions to Peto’s paradox revealed by mathematical modelling and cross-species cancer gene analysis. Philos Trans R Soc Lond Ser B 370:20140222

    Google Scholar 

  • Cavaliere M, Feng S, Soyer OS, Jiménez JI (2017) Cooperation in microbial communities and their biotechnological applications. Environ Microbiol 19:2949–2963

    Google Scholar 

  • Chalub FACC, Souza MO (2014) The frequency-dependent Wright–Fisher model: diffusive and non-diffusive approximations. J Math Biol 68:1089–1133

    MathSciNet  MATH  Google Scholar 

  • Chalub FACC, Souza MO (2017) On the stochastic evolution of finite populations. J Math Biol 75:1735–1774

    MathSciNet  MATH  Google Scholar 

  • Chapkin RS et al. Propagation of a signal within a gut gene-microbial ecological network with an interface layer, work in progress

  • Chen H, Kardar M (2019) Mean-field computational approach to HIV dynamics on a fitness landscape, preprint is available at https://www.biorxiv.org/content/10.1101/518704v2

  • Chesson P (2018) Updates on mechanisms of maintenance of species diversity. J Ecol 106:1773–1794

    Google Scholar 

  • Christie JR, Beekman M (2017) Selective sweeps of mitochondrial DNA can drive the evolution of uniparental inheritance. Evolution 71:2090–2099

    Google Scholar 

  • Chuang JS, Rivoire O, Leibler S (2009) Simpson’s paradox in a synthetic microbial system. Science (N S) 323:272–275

    Google Scholar 

  • Conley C (1978) Isolated invariant sets and the Morse index. In: C. B. M. S. Regional Conference Series in Math, vol 38. American Mathematical Society

  • Constable GWA, Rogers T, McKane AJ, Tarnita CE (2016) Demographic noise can reverse the direction of deterministic selection. Proc Natl Acad Sci USA 113:E4745–E4754

    Google Scholar 

  • Coron C, Méléard S, Villemonais D (2019) Impact of demography on extinction/fixation events. J Math Biol 78:549–577

    MathSciNet  MATH  Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

  • Cvijović I, Good BH, Desai MM (2018) The effect of strong purifying selection on genetic diversity. Genetics 209:1235–1278

    Google Scholar 

  • Doorn EV, Pollett P (2013) Quasi-stationary distributions for discrete state models. Eur J Oper Res 230:1–14

    MathSciNet  MATH  Google Scholar 

  • Datta RS, Gutteridge A, Swanton C, Maley CC, Graham TA (2013) Modelling the evolution of genetic instability during tumour progression. Evol Appl 6:20–33

    Google Scholar 

  • Dean A, Shnerb NM (2019) Stochasticity-induced stabilization in ecology and evolution, preprint is available at https://www.biorxiv.org/content/10.1101/725341v1

  • Der R, Epstein CL, Plotkin JB (2011) Generalized population models and the nature of genetic drift. Theor Popul Biol 80:80–99

    MATH  Google Scholar 

  • de Visser JAGM, Krug J (2014) Empirical fitness landscapes and the predictability of evolution. Nat Rev Genet 15:480–490

    Google Scholar 

  • Dewar RC, Sherwin WB, Thomas E, Holleley CE, Nichols RA (2011) Predictions of single-nucleotide polymorphism differentiation between two populations in terms of mutual information. Mol Ecol 20:3156–66

    Google Scholar 

  • Doebeli M, Ispolatov Y, Simon B (2017) Towards a mechanistic foundation of evolutionary theory. Elife 6:e23804

    Google Scholar 

  • dos Reis M (2015) How to calculate the non-synonymous to synonymous rate ratio of protein-coding genes under the Fisher-Wright mutation-selection framework. Biol Lett 11:20141031

    Google Scholar 

  • Edwards AWF (2016) Analysing nature’s experiment: Fisher’s inductive theorem of natural selection. Theor Popul Biol 109:1–5

    Google Scholar 

  • Ewens WJ (2011) What is the gene trying to do? Br J Philos Sci 62:155–176

    MathSciNet  Google Scholar 

  • Ewens WJ (2012) Mathematical population genetics (I. Theoretical introduction), 2nd edn. Interdisciplinary Applied Mathematics Series, vol 27. Springer, Berlin

  • Ewens WJ, Lessard S (2015) On the interpretation and relevance of the Fundamental Theorem of Natural Selection. Theor Popul Biol 104:59–67

    MATH  Google Scholar 

  • Fathi A, Pageault P (2015) Aubry–Mather theory for homeomorphisms. Ergodic Theory Dyn Syst 35:1187–1207

    MathSciNet  MATH  Google Scholar 

  • Fathi A, Pageault P (2019) Smoothing Lyapunov functions. Trans Am Math Soc 371:1677–1700

    MathSciNet  MATH  Google Scholar 

  • Faure M, Schreiber S (2014) Quasi-stationary distributions for randomly perturbed dynamical systems. Ann Appl Probab 24:553–598

    MathSciNet  MATH  Google Scholar 

  • Ferguson TS (1967) Mathematical statistics: a decision theoretic approach. Academic Press, New York

    MATH  Google Scholar 

  • Franke JE, Selgrade JF (1976) Abstract \(\omega \)-limit sets, chain recurrent sets, and basic sets for flows. Proc Am Math Soc 60:309–316

    MathSciNet  MATH  Google Scholar 

  • Franks J Notes on chain recurrence and Lyapunonv functions, (unpublished) lecture notes, available at https://arxiv.org/pdf/1704.07264.pdf

  • Garay BM, Hofbauer J (2003) Robust permanence for ecological differential equations, minimax, and discretizations. SIAM J Math Anal 34:1007–1039

    MathSciNet  MATH  Google Scholar 

  • Garcia V, Glassberg EC, Harpak A, Feldman MW (2018) Clonal interference can cause wavelet-like oscillations of multilocus linkage disequilibrium. J R Soc Interface 15:20170921

    Google Scholar 

  • Gavrilets S (2010) High-dimensional fitness landscapes and speciation. In: Pigliucci M, Müller GB (eds) Evolution–the extended synthesis. MIT Press, Cambridge, pp 45–79

    Google Scholar 

  • Gokhale CS, Traulsen A (2014) Evolutionary multiplayer games. Dyn Games Appl 4:468–488

    MathSciNet  MATH  Google Scholar 

  • Gompert Z (2016) Bayesian inference of selection in a heterogeneous environment from genetic time-series data. Mol Ecol 25:121–34

    Google Scholar 

  • Good BH, Martis S, Hallatschek O (2018) Directional selection limits ecological diversification and promotes ecological tinkering during the competition for substitutable resources, preprint is available at https://www.biorxiv.org/content/10.1101/292821v1

  • Goudenège L, Zitt P-A (2015) A Wright–Fisher model with indirect selection. J Math Biol 71:1411–1450

    MathSciNet  MATH  Google Scholar 

  • Grosjean N, Huillet T (2017) Wright–Fisher-like models with constant population size on average. Int J Biomath 10:1750078

    MathSciNet  MATH  Google Scholar 

  • Gyllenberg M, Silvestrov DS (2008) Quasi-stationary phenomena in nonlinearly perturbed stochastic systems, De Gruyter Expositions in Mathematics, vol 44. Walter de Gruyter

  • Harmand N, Federico V, Hindre T, Lenormand T (2019) Nonlinear frequency-dependent selection promotes long-term coexistence between bacteria species. Ecol Lett 22:1192–1202

    Google Scholar 

  • Harper M, Fryer D (2016) Stationary stability for evolutionary dynamics in finite populations. Entropy 18:316

    Google Scholar 

  • Hartle DL, Clark AG (2007) Principles of population genetics, 4th edn. Sinauer, Sunderland

    Google Scholar 

  • Hintze A, Olson RS, Adami C, Hertwig R (2015) Risk sensitivity as an evolutionary adaptation. Sci Rep 5:8242

    Google Scholar 

  • Hobolth A, Sireén J (2016) The multivariate Wright-Fisher process with mutation: moment-based analysis and inference using a hierarchical Beta model. Theor Popul Biol 108:36–50

    MATH  Google Scholar 

  • Hoeffding W (1963) Probability inequalities for sums of bounded random variables. J Am Stat Assoc 58:13–30

    MathSciNet  MATH  Google Scholar 

  • Hofbauer J (1985) The selection mutation equation. J Math Biol 23:41–53

    MathSciNet  MATH  Google Scholar 

  • Hofbauer J (2011) Deterministic evolutionary game dynamics. In: Sigmund K (ed) Evolutionary game dynamics. Proc. Sympos. Appl. Math., vol. 69, AMS Short Course Lecture Notes. Amer. Math. Soc, pp 61–79

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

    MATH  Google Scholar 

  • Hofbauer J, Sigmund K (2003) Evolutionary game dynamics. Bull Am Math Soc (N S) 40:479–519

    MathSciNet  MATH  Google Scholar 

  • Hofrichter J, Jost J, Tran T (2017) Information geometry and population genetics. The mathematical structure of the Wright–Fisher model. Springer, Berlin

    MATH  Google Scholar 

  • Houchmandzadeh B (2015) Fluctuation driven fixation of cooperative behavior. Biosystems 127:60–66

    Google Scholar 

  • Hummert S, Bohl K, Basanta D, Deutsch A, Werner S, Theißen G, Schroeter A, Schuster S (2014) Evolutionary game theory: cells as players. Mol BioSyst 10:3044

    Google Scholar 

  • Hummert S, Glock C, Lang SN, Hummert C, Skerka C, Zipfel PF, Germerodt S, Schuster S (2018) Playing “hide-and-seek’’ with factor H: game-theoretical analysis of a single nucleotide polymorphism. J R Soc Interface 15:20170963

    Google Scholar 

  • Hurley M (1998) Lyapunov functions and attractors in arbitrary metric spaces. Proc Am Math Soc 126:245–256

    MathSciNet  MATH  Google Scholar 

  • Hutchinson GE (1961) The paradox of the plankton. Am Nat 95:137–145

    Google Scholar 

  • Huttegger S, Skyrms B, Tarrès P, Wagner E (2014) Some dynamics of signaling games. Proc Natl Acad Sci USA 111:10873–10880

    MathSciNet  MATH  Google Scholar 

  • Hutson V, Schmitt K (1992) Permanence and the dynamics of biological systems. Math Biosci 111:1–71

    MathSciNet  MATH  Google Scholar 

  • Iglehart DL (1972) Extreme values in the \(GI/ G/ 1\) queue. Ann Math Stat 43:627–635

  • Iizuka M (2010) Effective population size of a population with stochastically varying size. J Math Biol 61:359–375

    MathSciNet  MATH  Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

  • Iwasa Y, Michor F (2011) Evolutionary dynamics of intratumor heterogeneity. PLoS ONE 6:e17866

    Google Scholar 

  • John S, Seetharaman S (2016) Exploiting the adaptation dynamics to predict the distribution of beneficial fitness effects. PLoS ONE 11:e0151795

    Google Scholar 

  • Kang Y, Chesson P (2010) Relative nonlinearity and permanence. Theor Popul Biol 78:26–35

    MATH  Google Scholar 

  • Karlin S, McGregor J (1964) Direct product branching processes and related Markov chains. Proc Natl Acad Sci USA 51:598–602

    MathSciNet  MATH  Google Scholar 

  • Karr AF (1975) Weak convergence of a sequence of Markov chains. Z Wahrsch Verw Gebiete 33:41–48

    MathSciNet  MATH  Google Scholar 

  • Kingman JFC (1961) A mathematical problem in population genetics. Proc Camb Phil Soc 57:574–582

    MathSciNet  MATH  Google Scholar 

  • Klebaner FC, Nerman O (1994) Autoregressive approximation in branching processes with a threshold. Stochastic Process Appl 51:1–7

    MathSciNet  MATH  Google Scholar 

  • Koopmann B, Müller J, Tellier A, Živković D (2017) Fisher–Wright model with deterministic seed bank and selection. Theor Popul Biol 114:29–39

    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 

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

    Google Scholar 

  • Landim C (2019) Metastable Markov chains. Probab Surv 16:143–227

    MathSciNet  MATH  Google Scholar 

  • Lawler GF (2006) Introduction to stochastic processes, 2nd edn. Chapman & Hall/CRC, Cambridge

    MATH  Google Scholar 

  • Li D, Cheng F (2019) The extinction and persistence of tumor evolution influenced by external fluctuations and periodic treatment. Qual Theory Dyn Syst (to appear)

  • Li L, Chesson P (2016) The effects of dynamical rates on species coexistence in a variable environment: the paradox of the plankton revisited. Am Nat 188:46–58

    Google Scholar 

  • Losert V, Akin E (1983) Dynamics of games and genes: discrete versus continuous time. J Math Biol 17:241–251

    MathSciNet  MATH  Google Scholar 

  • Mahdipour-Shirayeh A, Darooneh AH, Long AD, Komarova NL, Kohandel M (2017) Genotype by random environmental interactions gives an advantage to non-favored minor alleles. Sci Rep 7:5193

    Google Scholar 

  • Mandel SPH (1959) The stability of a multiple allelic system. Heredity 13:289–302

    Google Scholar 

  • McAvoy A, Fraiman N, Hauert C, Wakeley J, Nowak MA (2018) Public goods games in populations with fluctuating size. Theor Popul Biol 121:72–84

    MATH  Google Scholar 

  • Méléard S, Villemonais D (2012) Quasi-stationary distributions and population processes. Probab Surv 9:340–410

    MathSciNet  MATH  Google Scholar 

  • Nagylaki T (1986) The Gaussian approximation for random genetic drift. In: Karlin S, Nevo E (eds) Evolutionary processes and theory. Academic Press, New York, pp 629–642

    Google Scholar 

  • Nagylaki T (1990) Models and approximations for random genetic drift. Theor Popul Biol 37:192–212

    MathSciNet  MATH  Google Scholar 

  • Norton DE (1995) The fundamental theorem of dynamical systems. Comment Math Univ Carolin 36:585–597

    MathSciNet  MATH  Google Scholar 

  • Nourmohammad A, Otwinowski J, Plotkin JB (2016) Host-pathogen coevolution and the emergence of broadly neutralizing antibodies in chronic infections. PLoS Genet 12:e1006171

    Google Scholar 

  • Nowak MA, Sigmund K (2004) Evolutionary dynamics of biological games. Science 303:793–799

    Google Scholar 

  • Obolski U, Ram Y, Hadany L (2018) Key issues review: evolution on rugged adaptive landscapes. Rep Prog Phys 81:012602

    Google Scholar 

  • Orr HA (2009) Fitness and its role in evolutionary genetics. Nat Rev Genet 10:531–539

    Google Scholar 

  • Pageault P (2009) Conley barriers and their applications: chain-recurrence and Lyapunov functions. Topol Appl 156:2426–2442

    MathSciNet  MATH  Google Scholar 

  • Panageas I, Srivastava P, Vishnoi NK (2016) Evolutionary dynamics in finite populations mix rapidly, In: Krauthgamer R (ed) Proceedings of the twenty-seventh annual ACM-SIAM symposium on discrete algorithms. SIAM, pp 480–497

  • Papadimitriou C, Piliouras G (2016) From Nash equilibria to chain recurrent sets: solution concepts and topology. In: Sudan M (ed) ITCS’16-Proceedings of the 2016 ACM conference on innovations in theoretical computer science. ACM, New York, pp 227–235

  • Park S-C, Krug J (2007) Clonal interference in large populations. Proc Natl Acad Sci USA 104:18135

    Google Scholar 

  • Park HJ, Traulsen A (2017) Extinction dynamics from metastable coexistences in an evolutionary game. Phys Rev E 96:042412

    Google Scholar 

  • Parra-Rojas C, Challenger JD, Fanelli D, McKane AJ (2014) Intrinsic noise and two-dimensional maps: Quasicycles, quasiperiodicity, and chaos. Phys Rev E 90:032135

    Google Scholar 

  • Perko L (1991) Differential equations and dynamical systems, 3rd edn. Springer, Berlin

    MATH  Google Scholar 

  • Proulx SR (2011) The rate of multi-step evolution in Moran and Wright–Fisher populations. Theor Popul Biol 80:197–207

    MATH  Google Scholar 

  • Queller DC (2017) Fundamental theorems of evolution. Am Nat 189:345–353

    Google Scholar 

  • Rao CR (2001) Linear statistical inference and its applications, 2nd edn. Wiley, New York

    Google Scholar 

  • Rohlfs M, Hoffmeist TS (2003) An evolutionary explanation of the aggregation model of species coexistence. Proc Biol Sci 270:33–35

    Google Scholar 

  • Ruppin E, Papin JA, de Figueiredo LF, Schuster S (2010) Metabolic reconstruction, constraint-based analysis and game theory to probe genomescale metabolic networks. Curr Opin Biotechnol 21:502–510

    Google Scholar 

  • Samoilov MS, Arkin AP (2006) Deviant effects in molecular reaction pathways. Nat Biotechnol 24:1235–40

    Google Scholar 

  • Sandholm WH (2010) Population games and evolutionary dynamics, series on economic learning and social evolution, vol 8. MIT Press, Cambridge

    Google Scholar 

  • Schneider KA (2010) Maximization principles for frequency-dependent selection II: the one-locus multiallele case. J Math Biol 61:95–132

    MathSciNet  MATH  Google Scholar 

  • Schreiber SJ, Levine JM, Godoy O, Kraft NJB, Hart SP (2018) Does deterministic coexistence theory matter in a finite world? Insights from serpentine annual plants, preprint is available at https://www.biorxiv.org/content/10.1101/290882v1

  • Sigmund K (1986) A survey of replicator equations. In: Casti JL, Karlqvist A (eds) Complexity, language, and life: mathematical approaches, Biomathematics, vol 16. Springer, Berlin, pp 88–104

  • Silverman S (1992) On maps with dense orbits and the definition of chaos. Rocky Mountain J Math 22:353–375

    MathSciNet  MATH  Google Scholar 

  • Simonsen Speed N, Balding DJ, Hobolth A (2019) A general framework for moment-based analysis of genetic data. J Math Biol 78:1727–1769

    MathSciNet  MATH  Google Scholar 

  • Spichtig M, Kawecki TJ (2004) The maintenance (or not) of polygenic variation by soft selection in heterogeneous environments. Am Nat 164:70–84

    Google Scholar 

  • Steinsaltz D, Evans SN (2004) Markov mortality models: implications of quasistationarity and varying initial distributions. Theor Popul Biol 65:319–337

    MATH  Google Scholar 

  • Stephens C (2014) Selection, drift, and the “forces’’ of evolution. Philos Sci 71:550–570

    Google Scholar 

  • Svensson EI, Connallon T (2019) How frequency-dependent selection affects population fitness, maladaptation and evolutionary rescue. Evol Appl 12:1243–1258

    Google Scholar 

  • Tarnita CE (2017) The ecology and evolution of social behavior in microbes. J Exp Biol 220:18–24

    Google Scholar 

  • Taylor C, Nowak M (2006) Evolutionary game dynamics with non-uniform interaction rates. Theor Popul Biol 69:243–252

    MATH  Google Scholar 

  • Traulsen A, Claussen JC, Hauert C (2005) Coevolutionary dynamics: from finite to infinite populations. Phys Rev Lett 95:238701

    Google Scholar 

  • Traulsen A, Reed FA (2012) From genes to games: cooperation and cyclic dominance in meiotic drive. J Theor Biol 299:120–125

    MathSciNet  MATH  Google Scholar 

  • Vallier M, Abou Chakra M, Hindersin L, Linnenbrink M, Traulsen A, Baines JF (2017) Evaluating the maintenance of disease-associated variation at the blood group-related gene B4galnt2 in house mice. BMC Evol Biol 17:187

    Google Scholar 

  • Wasserman L (2004) All of statistics: a concise course in statistical inference. Springer texts in statistics. Springer, Berlin

    MATH  Google Scholar 

  • Waxman D, Loewe L (2010) A stochastic model for a single click of Muller’s ratchet. J Theor Biol 264:1120–1132

    MathSciNet  MATH  Google Scholar 

  • West SA, Diggle SP, Buckling A, Gardner A, Griffin AS (2007) The social lives of microbes. Annu Rev Ecol Evol Syst 38:53–77

    Google Scholar 

  • Xu S, Jiao S, Jiang P, Ao P (2014) Two-time-scale population evolution on a singular landscape. Phys Rev E 89:012724

    Google Scholar 

  • Zeng Q, Wu S, Sukumaran J, Rodrigo A (2017) Models of microbiome evolution incorporating host and microbial selection. Microbiome 5:127

    Google Scholar 

  • Zhang Y, Fu F, Wu T, Xie G, Wang L (2013) A tale of two contribution mechanisms for nonlinear public goods. Sci Rep 3:2021

    Google Scholar 

  • Zhou D, Qian H (2011) Fixation, transient landscape, and diffusion dilemma in stochastic evolutionary game dynamics. Phys Rev E 84:31907

    Google Scholar 

  • Zinger T, Gelbart M, Miller D, Pennings PS, Stern A (2019) Inferring population genetics parameters of evolving viruses using time-series data. Virus Evol 5:vez011

    Google Scholar 

  • Zomorrodi AR, Segrè D (2017) Genome-driven evolutionary game theory helps understand the rise of metabolic interdependencies in microbial communities. Nat Commun 8:1563

    Google Scholar 

Download references

Acknowledgements

We would like to thank Ozgur Aydogmus for a number of illuminating discussions. This work was supported by the Allen Endowed Chair in Nutrition and Chronic Disease Prevention, and the National Institute of Health (R35-CA197707 and P30E5029067). A. R. is supported by the Nutrition, Statistics & Bioinformatics Training Grant T32CA090301.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Ivan Ivanov.

Additional information

Publisher's Note

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

Appendices

Appendices

1.1 Appendix A. Absorption states and stochastic equilibria

In this appendix we explore the connection between basic geometric properties of the vector field \({\varvec{\Gamma }}\) and the structure of closed communication classes of the Markov chain \(\textbf{X}^{(N)}.\) In other words, we study the configuration of absorbing states and identify possible supports of stationary distributions according to the properties of \({\varvec{\Gamma }}\). We remark that the results of this section are included for the sake of completeness and are not used anywhere else in the remainder of the paper. We have made the section self-contained and included the necessary background on the Markov chain theory.

The main result of the appendix is stated in Theorem 7.3 where a geometric characterization of recurrent classes is described. The general structure of recurrent classes of \(\textbf{X}^{(N)}\) in this theorem holds universally for the Wright–Fisher model (7) with any update function \({\varvec{\Gamma }}.\) Thus, in Appendix A we deviate from assumption (10) that we have used throughout the paper and to which we will return in the subsequent Appendix B, and adopt instead the following general one:

Assumption 7.1

\(\textbf{X}^{(N)}\) is a Markov chain on \(\Delta _M\) with transition kernel defined by (7) and an arbitrary update rule \({\varvec{\Gamma }}:\Delta _M\rightarrow \Delta _M.\)

Recall (5) and (13). The discussion in this section is largely based on the fact that

$$\begin{aligned} P_N(\textbf{x},\textbf{y})>0 \quad \Longleftrightarrow {{\mathcal {C}}}(\textbf{y})\subset {{\mathcal {C}}}\big ({\varvec{\Gamma }}(\textbf{x})\big )\quad \Longleftrightarrow \quad P_N(\textbf{x},\textbf{z})>0~\text{ for } \text{ all }~\textbf{z}\in \Delta _{[{{\mathcal {C}}}(\textbf{y})],M}. \nonumber \\ \end{aligned}$$
(64)

This simple observation allows one to relate the geometry of zero patterns in the vector field to communication properties, and hence asymptotic behavior, of \(\textbf{X}^{(N)}.\)

From the point of view of the model’s adequacy in potential applications, (64) is a direct consequence of the fundamental assumption of the Wright–Fisher model that particles in the population update their phenotypes independently of each other and follow the same stochastic protocol. Mathematically, this assumption is expressed in identity (45), whose right-hand side is a sum of independent and identically distributed random indicators. To appreciate this feature of the Wright–Fisher model, we remark that while (64) holds true for an arbitrary model (7), that feature does not hold true, in general, for two most natural generalization of the Wright–Fisher model, namely the Cannings exchangeable model (Ewens 2012) and the pure-drift Generalized Wright–Fisher process of Der et al. (2011).

Recall that two states \(\textbf{x},\textbf{y}\in \Delta _{M,N}\) of the finite-state Markov chain \(\textbf{X}^{(N)}\) are said to communicate if there exist \(i,j\in {{\mathbb {N}}}\) such that \(P_N^i(\textbf{x},\textbf{y})>0\) and \(P_N^j(\textbf{y},\textbf{x})>0.\) If \(\textbf{x}\) and \(\textbf{y}\) communicate, we write \(\textbf{x}\leftrightarrow \textbf{y}.\) The binary relation \(\leftrightarrow \) partitions the state space \(\Delta _{M,N}\) into a finite number of disjoint equivalence classes, namely \(\textbf{x},\textbf{y}\) belong to the same class if and only if \(\textbf{x}\leftrightarrow \textbf{y}.\) A state \(\textbf{x}\in \Delta _{M,N}\) is called absorbing if \(P_N(\textbf{x},\textbf{x})=1,\) recurrent if \(\textbf{X}^{(N)}\) starting at \(\textbf{x}\) returns to \(\textbf{x}\) infinitely often with probability one, and transient if, with probability one, \(\textbf{X}^{(N)}\) starting at \(\textbf{x}\) revisits \(\textbf{x}\) only a finite number of times. It turns out that each communication class consists of either (1) exactly one absorbing state; or (2) recurrent non-absorbing states only; or (3) transient states only. No other behavior is possible due to zero–one laws enforced by the Markov property (Lawler 2006). A communication class is called closed or recurrent if it belongs to the first or second category, and transient if it belongs to the third one. The chain is called irreducible if there is only one (hence, recurrent) communication class, and aperiodic if \(P_N^j(\textbf{x},\textbf{y})>0\) for some \(j\in {{\mathbb {N}}}\) and all \(\textbf{x},\textbf{y}\in \Delta _{M,N}.\) If \(C_1,\ldots ,C_r\) are disjoint recurrent classes of \(\textbf{X}^{(N)},\) then the general form of its stationary distribution is \(\pi (\textbf{x})=\sum _{i=1}^r \alpha _i \pi _i(\textbf{x})\) where \(\pi _i\) is the unique stationary distribution supported on \(C_i\) (that is \(\pi _i(\textbf{x})>0\) if and only if \(\textbf{x}\in C_i\)) and \(\alpha _i\) are arbitrary non-negative numbers adding up to one. If the chain is irreducible, then the stationary distribution is unique and is strictly positive on the whole state space \(\Delta _{M,N}.\) An irreducible finite-state Markov chain is aperiodic if and only if \(\pi (\textbf{x})=\lim _{k\rightarrow \infty } P_N^k(\textbf{y},\textbf{x})\) for any \(\textbf{x},\textbf{y}\in \Delta _{M,N},\) where \(\pi \) is the unique stationary distribution of the chain (Lawler 2006).

Let \({{\mathcal {A}}}_N \subset \Delta _M\) denote the set of absorbing states of the Markov chain \(\textbf{X}^{(N)}.\) It readily follows from (7) that \({{\mathcal {A}}}_N\subset V_M,\) and a vertex \(\textbf{e}_j\in V_M\) is an absorbing (fixation) state if and only if \({\varvec{\Gamma }}(\textbf{e}_j)=\textbf{e}_j.\) In particular, \({{\mathcal {A}}}_N\) depends only on the update rule \({\varvec{\Gamma }}\) and is independent of the population size N. We summarize this observation as follows:

$$\begin{aligned} {{\mathcal {A}}}_N=\big \{\textbf{e}_j\in V_M:{\varvec{\Gamma }}(\textbf{e}_j)=\textbf{e}_j\big \}\qquad \qquad \forall \, N\in {{\mathbb {N}}}. \end{aligned}$$
(65)

The following lemma asserts that if \(\textbf{e}_j\) is an absorbing state of \(\textbf{X}^{(N)},\) then the type j cannot be mixed into any stochastic equilibrium of the model, namely any other than \(\textbf{e}_j\) state \(\textbf{x}\in \Delta _{M,N}\) with \(x(j)>0\) is transient.

Lemma 7.2

Let Assumption 7.1 hold. If \(\textbf{x}\in \Delta _{M,N}\) is a recurrent state of the Markov chain \(\textbf{X}^{(N)}\) and \(x(j)>0,\) then \(\textbf{e}_j\) is a recurrent state belonging to the same closed communication class as \(\textbf{x}.\)

Proof

By (65), a state \(\textbf{x}\in \Delta _{M,N}\) cannot be absorbing if \(\textbf{x}\ne \textbf{e}_j\) and \(x(j)>0.\) If \(\textbf{x}\) is recurrent and non-absorbing, then there is a state \(\textbf{y}\in \Delta _{M,N}\) that belongs to the same closed communication class as \(\textbf{x}\) and such that \(P_N(\textbf{y},\textbf{x})>0.\) In view of (7) this implies that \(\Gamma _j(\textbf{y})>0\) and hence \(P_N(\textbf{y},\textbf{e}_j)>0.\) Since the state \(\textbf{y}\) is recurrent, \(\textbf{e}_j\) must belong to the same communication class as \(\textbf{y}\) and \(\textbf{x},\) and in particular is recurrent. The proof is complete. \(\square \)

Recall (5). The same argument as in the above lemma shows that if an interior point \(\textbf{x}\) of a simplex \(\Delta _{[J],N}\) is recurrent for \(\textbf{X}^{(N)},\) then the whole simplex \(\Delta _{[J],N}\) belongs to the (recurrent) closed communication class of \(\textbf{x}.\) One can rephrase this observation as follows:

Theorem 7.3

Let Assumption 7.1 hold. Then, any recurrent class of the Markov chain \(\textbf{X}^{(N)}\) has the form of a simplicial complex \(\bigcup _\ell \Delta _{[J_\ell ],N},\) where \(J_\ell \) are (possibly overlapping) subsets (possibly singletons) of \(S_M.\)

With applications in mind, we collect some straightforward implications of this proposition in the next corollary.

Corollary 7.4

Let Assumption 7.1 hold. Then:

  1. (i)

    If there is a stationary distribution \(\pi \) of \(\textbf{X}^{(N)}\) and \(J\subset S_M\) such that \(\pi (\textbf{x})>0\) for some interior point \(\textbf{x}\in \Delta _{[J],N}^\circ ,\) then \(\pi (\textbf{y})>0\) for all \(\textbf{y}\in \Delta _{[J],N}.\) Furthermore, the following holds in this case unless \(\Gamma (\textbf{y})\cdot \textbf{y}=0\) (that is, \(\textbf{y}\) and \({\varvec{\Gamma }}(\textbf{y})\) are orthogonal) for all \(\textbf{y}\in \Delta _{[J],N}:\)

    1. (a)

      Let C be the (closed) communication class to which the above \(\textbf{x}\) (and hence the entire \(\Delta _{[J],N}\)) belongs. Then the Markov chain \(\textbf{X}^{(N)}\) restricted to C is aperiodic.

    2. (b)

      Let T be the first hitting time of C,  namely \(T=\inf \{k\in {{\mathbb {Z}}}_+:\textbf{X}^{(N)}_k\in C\}.\) As usual, we use here the convention that \(\inf \emptyset =+\infty .\) Then

      $$\begin{aligned} \pi (\textbf{y})=\lim _{k\rightarrow \infty } P(\textbf{X}^{(N)}_k=\textbf{y}\mid \textbf{X}^{(N)}_0=\textbf{z},T<\infty ) \end{aligned}$$

      for all \(\textbf{y}\in C\) and \(\textbf{z}\in \Delta _{M,N}.\) In particular, \(\pi (\textbf{y})=\lim _{k\rightarrow \infty }P_N^k(\textbf{z},\textbf{y})\) whenever \(\textbf{y},\textbf{z}\in C.\)

  2. (ii)

    A stationary distribution \(\pi \) of \(\textbf{X}^{(N)}\) such that \(\pi (\textbf{x})>0\) for some interior point \(\textbf{x}\in \Delta _{M,N}^\circ \) exists if and only if \(\textbf{X}^{(N)}\) is irreducible and aperiodic, in which case the stationary distribution is unique and \(\pi (\textbf{y})=\lim _{k\rightarrow \infty } P^k_N(\textbf{x},\textbf{y})>0\) for any \(\textbf{x},\textbf{y}\in \Delta _{M,N}.\)

  3. (iii)

    Let \(B=\big \{\exists ~k\in {{\mathbb {N}}}: \textbf{X}^{(N)}_k\in \partial \big (\Delta _{M,N}\big )~\text{ for } \text{ all }~m\ge k\big \}\) be the event that the trajectory of \(\textbf{X}^{(N)}\) reach eventually the boundary of the simplex and stays there forever. Then either the Markov chain \(\textbf{X}^{(N)}\) is irreducible and aperiodic or \(P(B)=1\) for any state \(\textbf{X}^{(N)}_0.\)

  4. (iv)

    If \({\varvec{\Gamma }}(\textbf{e}_j)=\textbf{e}_j\) for some \(j\in S_M,\) then \(\lim _{k\rightarrow \infty } X^{(N)}_k(j)\in \{0,1\},\) a. s., for any initial distribution of \(\textbf{X}^{(N)}.\)

We will only prove the claim in (a) of part (i), since the rest of the corollary is an immediate result of a direct combination of Theorem 7.3 and general properties of finite-state Markov chains reviewed in the beginning of the section.

Proof of Corollary 7.4-(i)-a

Suppose that \({\varvec{\Gamma }}(\textbf{y})\cdot \textbf{y}>0\) for some \(\textbf{y}\in \Delta _{[J],N}.\) Let \(\textbf{z}\) be the projection of \(\textbf{y}\) into \(\Delta _{[I],M},\) where \(I:=\{j\in J:y(j)\Gamma _j(\textbf{y})>0\}.\) That is,

$$\begin{aligned} z(i)= \left\{ \begin{array}{ll} y(i)&{}\text{ if }~y(i)\Gamma _i(\textbf{y})>0, \\ 0&{}\text{ otherwise }. \end{array} \right. \end{aligned}$$

Then \(\textbf{z}\in \Delta _{[I],M}^\circ \) and \(P_N(\textbf{z},\textbf{z})>0.\) Since \(\textbf{z}\in \Delta _{[J],N},\) by Theorem 7.3\(\textbf{z}\) belongs to the same closed communication class as \(\textbf{x}.\) The asserted aperiodicity of \(\textbf{X}^{(N)}\) restricted to C is a direct consequence of the existence of the 1-step communication loop at \(\textbf{z}\) (Lawler 2006). \(\square \)

Example 6

(Harper and Fryer 2016) If \(\Gamma _j(\textbf{x})=x(j)^\alpha f(\textbf{x})\) for some \(j\in S_M,\) function \(f:\Delta _M\rightarrow {{\mathbb {R}}}_+\) such that \(f(\textbf{e}_j)=1,\) and a constant \(\alpha >0,\) then \({\varvec{\Gamma }}(\textbf{e}_j)=\textbf{e}_j.\) By part (iv) of Corollary 7.4, this fact alone suffices to conclude that \(\lim _{k\rightarrow \infty } X^{(N)}_k(j)\in \{0,1\},\) a. s., for this specific type j.

A simple sufficient condition for the Markov chain \(\textbf{X}^{(N)}\) to be irreducible and aperiodic is that \({\varvec{\Gamma }}(\Delta _M)\subset \Delta _M^\circ ,\) in which case \(P_N(\textbf{x},\textbf{y})>0\) for any \(\textbf{x},\textbf{y}\in \Delta _{M,N}.\) If \({\varvec{\Gamma }}\big (\Delta _{M,N}^\circ \big )\subset \Delta _{M,N}^\circ \) but \({\varvec{\Gamma }}(\textbf{e}_j)=\textbf{e}_j\) for all \(j\in S_M,\) the model can be viewed as a multivariate analog of what is termed in Chalub and Souza (2017) as a Kimura class of Markov chains. The latter describes evolution dynamics of a population that initially consists of two phenotypes, each one eventually either vanishes or takes over the entire population.

1.2 Appendix B. Permanence of the mean-field dynamics

In this appendix, we focus on Wright–Fisher models with a permanent mean-field dynamics \({\varvec{\Gamma }}.\) The permanence means that the boundary of the state space \(\partial (\Delta _M)\) is a repeller for the mean-field dynamics, implying that all types under the mean-field dynamics will ultimately survive (Garay and Hofbauer 2003; Hofbauer and Sigmund 1998; Hutson and Schmitt 1992; Kang and Chesson 2010). This condition is particularly relevant to the stochastic Wright–Fisher model because, as stated in Appendix A, all absorbing states (possibly an empty set) of the Markov chain \(\textbf{X}^{(N)}\) lie within the boundary set \(\partial (\Delta _M).\) Of course, in the absence of mutations, any vertex of \(\Delta _M\) is an absorbing state for the Markov chain \(\textbf{X}^{(N)},\) and the latter is going to eventually become fixed on one of the absorbing states with probability one (see Appendix A for formal details). However, we showed in Sect. 4 that a permanent mean-field dynamics induces metastability and a prolonged coexistence of types.

Definition 4

(Hofbauer and Sigmund 1998) \({\varvec{\Gamma }}\) is permanent if there exists \(\delta >0\) such that

$$\begin{aligned} \liminf _{k\rightarrow \infty } \psi _k(i)>\delta \qquad \forall \,i\in S_M,{{\varvec{\psi }}}_0\in \Delta _M^\circ . \end{aligned}$$

See Garay and Hofbauer (2003), Schreiber et al. (2018) and references therein for sufficient conditions for the permanence. It turns out that under reasonable additional assumptions, \({\varvec{\Gamma }}\) is permanent for the fitness defined in either (11) or (12). More precisely, we have (for a more general version of the theorem, see Garay and Hofbauer (2003, Theorem 11.4) and Hofbauer and Sigmund (2003, Theorem 5):

Theorem 3

(Garay and Hofbauer 2003) Let \(\textbf{A}\) be an \(M\times M\) matrix which satisfies the following condition: There exists \(\textbf{y}\in \Delta _M^\circ \) such that

$$\begin{aligned} \textbf{y}^T\textbf{A} \textbf{z}> \textbf{z}^T\textbf{A}{} \textbf{z} \end{aligned}$$

for all \(\textbf{z}\in \partial (\Delta _M)\) solving the fixed-point equation \(\textbf{z}=\frac{1}{\textbf{z}^T\textbf{A}{} \textbf{z}}\big (\textbf{z}\circ \textbf{A}{} \textbf{z}\big ).\)

Then:

  1. (i)

    There exists \(\omega _0>0\) such that \({\varvec{\Gamma }}\) given by (10) and (11) is permanent for all \(\omega \in (0,\omega _0).\)

  2. (ii)

    \({\varvec{\Gamma }}\) given by (10) and (12) is permanent for all \(\eta >0.\)

Recall Definitions 1 and 2. The setting of the theorem provides a natural within our context example of a non-constant complete Lyapunov function.

Example 7

Suppose that \({\varvec{\Gamma }}\) is a replicator function defined in (10). If \(\textbf{A}\) is an \(M\times M\) matrix that satisfies the conditions of Theorem 3 and \({{\varvec{\varphi }}}\) in (10) is defined by either (11) with small enough \(\omega >0\) or by (12), then \({{\mathcal {R}}}({\varvec{\Gamma }}) \subset K\bigcup \partial (\Delta _M),\) where K is a compact subset of \(\Delta _M^\circ .\) In particular, the complete Lyapunov function h constructed in Theorem 2 (p. 14) is strictly increasing on the non-empty open set \(\Delta _M^\circ \backslash K.\) Notice that while the average \(\textbf{x}\cdot {{\varvec{\varphi }}}(\textbf{x})\) may not be a complete Lyapunov function for \({\varvec{\Gamma }},\) the average \(h(\textbf{x})\) of the reproductive fitness \(\widetilde{{\varvec{\varphi }}}(\textbf{x}):=h(\textbf{x}){{\varvec{\varphi }}}(\textbf{x})\) is.

In Sect. 4.3 we considered a stochastic Wright–Fisher model with a permanent mean-field map \({\varvec{\Gamma }}.\) This model has a global attractor, consisting of a unique equilibrium point. A sufficient condition for such a dynamics is given by the following theorem (Losert and Akin 1983). Recall (20).

Theorem 4

(Losert and Akin 1983) Let Assumption 4.4 hold. Then the following holds true:

  1. (a)

    Recall \({{\varvec{\psi }}}_k\) from (3). If \({{\varvec{\psi }}}_0\in \Delta _M^\circ ,\) then \(\lim _{k\rightarrow \infty } {{\varvec{\psi }}}_k={{\varvec{\chi _{eq}}}}.\)

  2. (b)

    \({\varvec{\Gamma }}\) is a diffeomorphism.

Note that the weak selection condition (cf. Antal et al. 2009; Boenkost et al. 2019; Imhof and Nowak 2006) imposed in part (i) of Theorem 3 is not required in the conditions of Theorem 4.

The evolutionary games with \(\textbf{A}=\textbf{A}^T\) are called partnership games (Hofbauer 2011). We refer to Hummert et al. (2018), Huttegger et al. (2014), Traulsen and Reed (2012) for applications of partnership games in evolutionary biology. Condition (32) implies the existence of a unique evolutionary stable equilibrium for the evolutionary game defined by the payoff matrix \(\textbf{A}.\) For a symmetric reversible matrix \(\textbf{A}\) this condition holds if and only if \(\textbf{A}\) has exactly one positive eigenvalue (Kingman 1961; Mandel 1959). Condition (33) ensures that the equilibrium is an interior point. Note that since the equilibrium is unique, the conditions of Theorem 3 are automatically satisfied, that is \({{\varvec{\psi }}}_k\) is permanent. Finally, under the conditions of Theorem 4, the average payoff \(\textbf{x}^T\textbf{A}{} \textbf{x}\) is a Lyapunov function (Kingman 1961; Mandel 1959):

$$\begin{aligned} {{\varvec{\psi }}}_k^T\textbf{A}{{\varvec{\psi }}}_k<{{\varvec{\psi }}}_{k+1}^T\textbf{A}{{\varvec{\psi }}}_{k+1}<{{\varvec{\chi _{eq}}}}^T \textbf{A}{{\varvec{\chi _{eq}}}}\qquad \qquad \forall \,k\in {{\mathbb {Z}}}_+, \end{aligned}$$

as long as \({{\varvec{\psi }}}_0\in \Delta _M^\circ \) and \({{\varvec{\psi }}}_0\ne {{\varvec{\chi _{eq}}}}.\)

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

Roitershtein, A., Rastegar, R., Chapkin, R.S. et al. Extinction scenarios in evolutionary processes: a multinomial Wright–Fisher approach. J. Math. Biol. 87, 63 (2023). https://doi.org/10.1007/s00285-023-01993-7

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00285-023-01993-7

Keywords

Mathematics Subject Classification

Navigation