Skip to main content
Log in

Population Games and Discrete Optimal Transport

  • Published:
Journal of Nonlinear Science Aims and scope Submit manuscript

Abstract

We propose an evolutionary dynamics for population games with discrete strategy sets, inspired by optimal transport theory and mean field games. The proposed dynamics is the Smith dynamics with strategy graph structure, in which payoffs are modified by logarithmic terms. The dynamics can be described as a Fokker–Planck equation on a discrete strategy set. For potential games, the dynamics is a gradient flow system under a Riemannian metric from optimal transport theory. The stability of the dynamics is studied through optimal transport metric tensor, free energy and Fisher information.

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
Fig. 9
Fig. 10

Similar content being viewed by others

Notes

  1. \(\beta \) represents the inverse of temperature.

References

  • Akin, E.: The Geometry of Population Genetics, vol. 280. Springer, Berlin (1979)

    Book  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  • Blanchet, A., Carlier, G.: Optimal transport and Cournot-Nash equilibria. arXiv preprint arXiv:1206.6571 (2012)

  • Blanchet, A., Carlier, G.: From Nash to Cournot–Nash equilibria via the Monge–Kantorovich problem. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 372(2028), 20130398 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  • Cardaliaguet, P.: Notes on mean field games, Technical report (2010)

  • Carrillo, J.A., McCann, R.J., Villani, C.: Kinetic equilibration rates for granular media and related equations: entropy dissipation and mass transportation estimates. Revista Matematica Iberoamericana 3(19), 971–1018 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  • Chow, S.-N., Huang, W., Li, Y., Zhou, H.: Fokker–Planck equations for a free energy functional or Markov process on a graph. Arch. Ration. Mech. Anal. 203(3), 969–1008 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  • Chow, S.-N., Li, W., Zhou, H.: Entropy dissipation of Fokker–Planck equations on graphs. arXiv:1701.04841 (2017a)

  • Chow, S.-N., Li, W., Lu, J., Zhou, H.: Game theory and discrete optimal transport. arXiv:1703.08442 (2017b)

  • Coucheney, P., Gaujal, B., Mertikopoulos, P.: Penalty-regulated dynamics and robust learning procedures in games. Math. Oper. Res. 40(3), 513–796 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  • Degond, P., Liu, J.-G., Ringhofer, C.: Large-scale dynamics of mean-field games driven by local Nash equilibria. J. Nonlinear Sci. 24(1), 93–115 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  • Erbar, M., Maas, J.: Ricci curvature of finite Markov chains via convexity of the entropy. Arch. Ration. Mech. Anal. 206(3), 997–1038 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  • Frieden, B.R.: Science from Fisher Information: A Unification. Cambridge University Press, Cambridge (2004)

    Book  MATH  Google Scholar 

  • Fudenberg, D., Levine, D.K.: The Theory of Learning in Games. MIT Press, Cambridge (1998)

    MATH  Google Scholar 

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

    MATH  Google Scholar 

  • Hofbauer, J., Sigmund, K.: Evolutionary game dynamics. Bull. Am. Math. Soc. 40(4), 479–519 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  • Hofbauer, J., Sandholm, W.H.: On the global convergence of stochastic fictitious play. Econometrica 70, 2265–2294 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  • Hofbauer, J., Sandholm, W.H.: Evolution in games with randomly disturbed payoffs. J. Econ. Theory 132, 47–69 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  • Huang, Y., Hao, Y., Wang, M., Zhou, W., Wu, Z.: Optimality and stability of symmetric evolutionary games with applications in genetic selection. Math. Biosci. Eng. 12(3), 503–523 (2015)

  • Lasry, J.-M., Lions, P.-L.: Mean field games. Jpn. J. Math. 2(1), 229–260 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  • Li, W.: A study of stochastic differential equations and Fokker–Planck equations with applications. Ph.D thesis (2016)

  • Li, W.: Geometry of probability simplex via optimal transport. arXiv:1803.06360 (2018)

  • Lieberman, E., Hauert, C., Nowak, M.A.: Evolutionary dynamics on graphs. Nature 433(7023), 312–316 (2005)

    Article  Google Scholar 

  • David, S.L., Collins, E.J.: Individual Q-learning in normal form games. SIAM J. Control Optim. 44(2), 495–514 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  • Maas, J.: Gradient flows of the entropy for finite Markov chains. J. Funct. Anal. 261(8), 2250–2292 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  • Matsui, A.: Best response dynamics and socially stable strategies. J. Econ. Theory 57(2), 343–362 (1992)

    Article  MathSciNet  MATH  Google Scholar 

  • Mertikopoulos, P., Sandholm, W.H.: Riemannian game dynamics. arXiv preprint arXiv:1603.09173 (2016)

  • Mertikopoulos, P., Sandholm, W.H.: Learning in games via reinforcement and regularization. Math. Oper. Res. 41(4), 1297–1324 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  • Monderer, D., Shapley, L.S.: Potential games. Games Econ. Behav. 14(1), 124–143 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  • Nash, J.F.: Equilibrium points in n-person games. Proc. Natl. Acad. Sci. 36(1), 48–49 (1950)

    Article  MathSciNet  MATH  Google Scholar 

  • Nowak, M.A.: Evolutionary Dynamics. Harvard University Press, Harvard (2006)

    Book  MATH  Google Scholar 

  • Sandholm, W.H.: Population Games and Evolutionary Dynamics. MIT Press, Cambridge (2010)

    MATH  Google Scholar 

  • Sandholm, W.H.: Evolutionary Game Theory. In: Meyers, R. (ed) Encyclopedia of Complexity and Systems Science, pp. 3176–3205. Springer, New York, NY (2012a)

  • Sandholm, W.H.: Decompositions and potentials for normal form games. Games Econ. Behav. 70(2), 446–456 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  • Sandholm, W.H.: Local stability of strict equilibria under evolutionary game dynamics. J. Dyn. Games 5(1), 27–50 (2012b)

    Google Scholar 

  • Shah, D., Shin, J.: Dynamics in congestion games. In: ACM SIGMETRICS Performance Evaluation Review, vol. 38, pp. 107–118. ACM (2010)

  • Shahshahani, S.: A new mathematical framework for the study of linkage and selection. Mem. Am. Math. Soc. 211 (1979). https://doi.org/10.1090/memo/0211

  • Sigmund, K., Nowak, M.A.: Evolutionary game theory. Curr. Biol. 9(14), R503–R505 (1999)

    Article  Google Scholar 

  • Michael, J.S.: The stability of a dynamic model of traffic assignment-an application of a method of Lyapunov. Transp. Sci. 18(3), 245–252 (1984)

    Article  MathSciNet  Google Scholar 

  • Szabo, G., Fath, G.: Evolutionary games on graphs. Phys. Rep. 446(4), 97–216 (2007)

    Article  MathSciNet  Google Scholar 

  • Villani, C.: Topics in Optimal Transportation, vol. 58. American Mathematical Society, Providence (2003)

    MATH  Google Scholar 

  • Villani, C.: Optimal Transport: Old and New, vol. 338. Springer, Berlin (2008)

    MATH  Google Scholar 

  • Von Neumann, J., Morgenstern, O.: Theory of Games and Economic Behavior (60 Anniversary Commemorative Edition). Princeton University Press, Princeton (2007)

    Book  Google Scholar 

  • Wu, A., Liao, D., Tlsty, T.D., Sturm, J.C., Austin, R.H.: Game theory in the death galaxy: interaction of cancer and stromal cells in tumour microenvironment. Interface Focus 4(4), 20140028 (2014)

    Article  Google Scholar 

Download references

Acknowledgements

This paper is based on Wuchen Li’s thesis Li (2016).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Wuchen Li.

Additional information

Communicated by Paul Newton.

This work is partially supported by NSF Awards DMS–1419027, DMS-1620345, and ONR Award N000141310408.

Appendix

Appendix

In this section, we briefly review the Best-reply dynamics and its connection with optimal transport theory. These serve the motivations of the dynamics considered in this paper. For more details see Degond et al. (2014), Villani (2008).

Best-reply dynamics and Fokker–Planck equations We first consider a game consisting N players \(i\in \{1,\ldots ,N\}\). Each player i chooses a strategy \(x_i\) from a same Borel strategy set S. For concreteness, we consider \(S=\mathbb {T}^d\), which is a d dimensional torus. Suppose each player receives a payoff function \(F_i\in C^{\infty }(S)\). For notational connivence, we denote \(F_i(x_i,x_{-i})=F_i(x_1,\ldots , x_N)\), where we abuse the notation by

$$\begin{aligned} x_{-i}=\{x_1,\ldots , x_{i-1}, x_{i+1},\ldots , x_N\}\ . \end{aligned}$$

We model players’ decision-making processes in a game by stochastic process \(x_{i}(t),~t\in [0,+\infty )\). Here, t is an artificial time variable, at which player i selects his/her decision based on the current strategies of all other players \(x_{-i}(t)\). We note that all players make their decisions simultaneously and without knowing others’ decisions. Each player selects his or her strategy that increases the player’s payoff most rapidly. In other words, we model the game by the following stochastic differential equations (SDEs)

$$\begin{aligned} \hbox {d} x_i= \nabla _{x_i}F_i (x_i, x_{-i})\hbox {d}t + \sqrt{2\beta } dB_{t}^{i}\ , \end{aligned}$$
(16)

where the independent Brownian motion \((B_t^i)_{i=1}^N\) is added to model the uncertainty of each player and \(\beta >0\) controls the magnitude of the noise.

Under the standard assumptions in population games, i.e., the game is autonomous and the players are symmetric, one can simply encode all the information of players into one probability density \(\rho \in \mathcal {P}(S)\) by taking \(N\rightarrow \infty \). In this limiting processes, each player’s cost function is rewritten as \(F:S\times \mathcal {P}(S)\rightarrow \mathbb {R}\), and the limiting stochastic process forms the following mean field SDE

$$\begin{aligned} \hbox {d}X_t= \nabla _{X_t}F (X_t, \rho )\hbox {d}t + \sqrt{2\beta } dB_{t}^{i}\ , \end{aligned}$$
(17)

where \(X_t\) has probability law \(\rho (t,x)\).

In Degond et al. (2014), SDE (17) is called the Best-reply dynamics, and \(X_t\) is the Best-reply decision process. Here, the transition density function \(\rho (t,x)\) of the stochastic process X(t) satisfies the FPE

$$\begin{aligned} \frac{\partial \rho (t,x)}{\partial t}=-\nabla \cdot (\rho (t,x)F(x,\rho ))+\beta \Delta \rho (t,x)\ . \end{aligned}$$
(18)

The game is called a potential game if there exists a potential function \(\mathcal {F}:\mathcal {P}(S)\rightarrow \mathbb {R}\), such that \(\frac{\delta }{\delta \rho (x)}\mathcal {F}(\rho )=F(x,\rho )\). For potential games, the Best-reply SDE (17) becomes

$$\begin{aligned} \hbox {d}X_t=\nabla \frac{\delta }{\delta \rho (t, X_t)}\mathcal {F}(\rho ) \hbox {d}t+\sqrt{2\beta } dB_t\ , \end{aligned}$$

which is a perturbed gradient flow and whose transition equation (FPE) forms

$$\begin{aligned} \frac{\partial \rho (t,x)}{\partial t}=-\nabla \cdot (\rho (t,x)\nabla \frac{\delta }{\delta \rho (t,x)}\mathcal {F}(\rho ))+\beta \Delta \rho (t,x)\ . \end{aligned}$$
(19)

From the theory of optimal transport, Equation (19) can be interpreted as a gradient ascend flow of the free energy

$$\begin{aligned} \mathcal {\bar{F}}(\rho )=\mathcal {F}(\rho )-\beta \int _{S}\rho (x)\log \rho (x)\hbox {d}x\ . \end{aligned}$$
(20)

Optimal transport and density manifold We next review the geometry of optimal transport on the continuous strategy set S.

Consider the set \(\mathcal {P}_2(S)\) of Borel measurable probability density functions on S with finite second moment. Given \(\rho ^0, \rho ^1\in \mathcal {P}_2(S)\), the \(L^2\)-Wasserstein distance between \(\rho ^0\) and \(\rho ^1\) is denoted by \(W:\mathcal {P}_2(S)\times \mathcal {P}_2(S)\rightarrow \mathbb {R}_+\). There are two equivalent ways of defining this distance.

The first definition is the following linear programming formulation:

$$\begin{aligned} W(\rho ^0, \rho ^1)^2=\inf _{\pi \in \Pi (\rho ^0, \rho ^1)}\int _{\Omega \times \Omega }d_\Omega (x,y)^2\pi (\hbox {d}x,\hbox {d}y)\ , \end{aligned}$$
(21)

where the infimum is taken over the set \(\Pi \) of joint probability measures on \(\Omega \times \Omega \) that have marginals \(\rho ^0\), \(\rho ^1\).

The second definition considers a probability path \(\rho :[0,1]\rightarrow \mathcal {P}_2(S)\) connecting \(\rho ^0\), \(\rho ^1\). And the distance is defined by a variational problem known as the Benamou–Brenier formula:

$$\begin{aligned} W(\rho ^0, \rho ^1)^2=\inf _{\Phi }~\int _0^1\int _{\Omega } (\nabla \Phi (t,x), \nabla \Phi (t,x))\rho (t,x) \hbox {d}x \hbox {d}t\ , \end{aligned}$$
(22a)

where the infimum is taken over the set of Borel potential functions \([0,1]\times S \rightarrow \mathbb {R}\). Each potential function \(\Phi \) determines a corresponding density path \(\rho \) as the solution of the continuity equation

$$\begin{aligned} \frac{\partial \rho (t,x)}{\partial t}+\text {div} (\rho (t,x)\nabla \Phi (t,x))=0\ ,\quad \rho (0,x)=\rho ^0(x)\ ,\quad \rho (1,x)=\rho ^1(x)\ . \end{aligned}$$
(22b)

Here, \(\text {div}\) and \(\nabla \) are the divergence and gradient operators in \(\Omega \). The continuity equation is known as the probability density transition equation according to the given vector field.

The equivalence between the static (21) and dynamical (22) formulations is well known. Moreover, the variational formulation (22) entails a similar Riemannian structure used in this paper. For simplicity, we only consider the set of smooth and strictly positive probability densities

$$\begin{aligned} \mathcal {P}_+(S)=\Big \{\rho \in C^{\infty }(\Omega ):\rho (x)>0\ ,~\int _{\Omega }\rho (x)\hbox {d}x=1\Big \} \subset \mathcal {P}_2(S)\ . \end{aligned}$$

Denote \(\mathcal {F}(S):=C^{\infty }(S)\) the set of smooth real valued functions on S. The tangent space of \(\mathcal {P}_+(S)\) is given by

$$\begin{aligned} T_\rho \mathcal {P}_+(S) = \Big \{\sigma \in \mathcal {F}(S):\int _{S}\sigma (x) \hbox {d}x=0 \Big \}\ . \end{aligned}$$

Given \(\Phi \in \mathcal {F}(S)\) and \(\rho \in \mathcal {P}_+(S)\), define

$$\begin{aligned} V_{\Phi }(x):=-\text {div} (\rho (x) \nabla \Phi (x))\ . \end{aligned}$$

Thus, \(V_\Phi \in T_{\rho }\mathcal {P}_+(S)\). The elliptic operator \(\nabla \cdot (\rho \nabla )\) identifies the function \(\Phi \) on S modulo additive constants with the tangent vector \(V_{\Phi }\) of the space of densities. This gives an isomorphism

$$\begin{aligned} \mathcal {F}(S)/\mathbb {R}\rightarrow T_{\rho }\mathcal {P}_+(S); \quad \Phi \mapsto V_\Phi \ . \end{aligned}$$

Define the Riemannian metric (inner product) on the tangent space of positive densities \(g^W:{T_\rho }\mathcal {P}_+(S)\times {T_\rho }\mathcal {P}_+(S)\rightarrow \mathbb {R}\) by

$$\begin{aligned} g^W_\rho (V_{\Phi }, V_{\tilde{\Phi }})=\int _{S}(\nabla \Phi (x), \nabla \tilde{\Phi }(x))\rho (x) \hbox {d}x\ , \end{aligned}$$

where \(\Phi (x)\), \(\tilde{\Phi }(x)\in \mathcal {F}(S)/\mathbb {R}\). The inner product endows \(\mathcal {P}_+(S)\) with an infinite-dimensional Riemannian metric tensor. In other words, the variational problem (22) is a geometric action energy in \((\mathcal {P}_+(S), g^W)\).

We are now ready to present the gradient operator of free energy w.r.t. \(L^2\)-Wasserstein metric tensor. Following

$$\begin{aligned} g^W(\text {grad}_W\mathcal {\bar{F}}(\rho ), V_{\Phi })=\int _{S}\frac{\delta }{\delta \rho (x)}\mathcal {\bar{F}}(\rho )V_{\Phi }\hbox {d}x\ \end{aligned}$$

and \(\frac{\delta }{\delta \rho (x)}\mathcal {F}(\rho )=F(x,\rho )\), and noticing \(\frac{\delta }{\delta \rho (x)}\int _{S}\rho (x)\log \rho (x)\hbox {d}x=\log \rho (x)+1\), we obtain

$$\begin{aligned} \text {grad}_W\mathcal {\bar{F}}(\rho )=-\nabla \cdot (\rho \nabla (F(x,\rho )-\beta \log \rho (x)))\ . \end{aligned}$$

From the fact that \(\nabla \cdot (\rho \nabla \log \rho )=\nabla \cdot (\nabla \rho )=\Delta \rho \), we derive FPE (19) by the gradient flow of the free energy

$$\begin{aligned} \frac{\partial \rho }{\partial t}=\text {grad}_W\mathcal {F}(\rho )=-\nabla \cdot (\rho \nabla F(x,\rho ))+\beta \Delta \rho . \end{aligned}$$

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chow, SN., Li, W., Lu, J. et al. Population Games and Discrete Optimal Transport. J Nonlinear Sci 29, 871–896 (2019). https://doi.org/10.1007/s00332-018-9507-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00332-018-9507-5

Keywords

Navigation