Skip to main content
Log in

Lattice approximations of the first-order mean field type differential games

  • Published:
Nonlinear Differential Equations and Applications NoDEA Aims and scope Submit manuscript

Abstract

The theory of first-order mean field type differential games examines the systems of infinitely many identical agents interacting via some external media under assumption that each agent is controlled by two players. We study the approximations of the value function of the first-order mean field type differential game using solutions of model finite-dimensional differential games. The model game appears as a mean field type continuous-time Markov game, i.e., the game theoretical problem with the infinitely many agents and dynamics of each agent determined by a controlled finite state nonlinear Markov chain. Given a supersolution (resp. subsolution) of the Hamilton–Jacobi equation for the model game, we construct a suboptimal strategy of the first (resp. second) player and evaluate the approximation accuracy using the modulus of continuity of the reward function and the distance between the original and model games. This gives the approximations of the value function of the mean field type differential game by values of the finite-dimensional differential games. Furthermore, we present the way to build a finite-dimensional differential game that approximates the original game with a given accuracy.

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.

Similar content being viewed by others

References

  1. Ahmed, N., Ding, X.: Controlled McKean–Vlasov equation. Commun. Appl. Anal. 5, 183–206 (2001)

    MathSciNet  MATH  Google Scholar 

  2. Aliprantis, C.D., Border, K.C.: Infinite Dimensional Analysis: A Hitchhiker’s Guide. Springer, Berlin (2006)

  3. Ambrosio, L., Gigli, N., Savaré, G.: Gradient Flows: in Metric Spaces and in the Space of Probability Measures. Lectures in Mathematics. ETH Zurich, Birkhäuser, Basel (2005)

    MATH  Google Scholar 

  4. Andersson, D., Djehiche, B.: A maximum principle for SDEs of mean-field type. Appl. Math. Optim. 63(3), 341–356 (2011)

    Article  MathSciNet  Google Scholar 

  5. Averboukh, Y.: Approximate solutions of continuous-time stochastic games. SIAM J. Control Optim. 54, 2629–2649 (2016). https://doi.org/10.1137/16M1062247

    Article  MathSciNet  MATH  Google Scholar 

  6. Averboukh, Y.: Krasovskii–Subbotin approach to mean field type differential games. Dyn. Games Appl. 9, 573–593 (2019). https://doi.org/10.1007/s13235-018-0282-6

    Article  MathSciNet  MATH  Google Scholar 

  7. Averboukh, Y.: A stability property in mean field type differential games. J. Math. Anal. Appl. 498(1), 124940 (2021)

  8. Bardi, M., Capuzzo-Dolcetta, I.: Optimal Control and Viscosity Solutions of Hamilton–Jacobi–Bellman Equations. Birkhäuser, Basel (1996)

    MATH  Google Scholar 

  9. Bayraktar, E., Cosso, A., Pham, H.: Randomized dynamic programming principle and Feynman–Kac representation for optimal control of McKean–Vlasov dynamics. Trans. Am. Math. Soc. 370, 2115–2160 (2018)

    Article  MathSciNet  Google Scholar 

  10. Bensoussan, A., Frehse, J., Yam, P.: Mean Field Games and Mean Field Type Control Theory. Springer, New York (2013)

    Book  Google Scholar 

  11. Buckdahn, R., Djehiche, B., Li, J.: A general stochastic maximum principle for SDEs of mean-field type. Appl. Math. Optim. 64(2), 197–216 (2011)

    Article  MathSciNet  Google Scholar 

  12. Carmona, R., Delarue, F.: Forward-backward stochastic differential equations and controlled McKean–Vlasov dynamics. Ann. Probab. 43(5), 2647–2700 (2015)

    Article  MathSciNet  Google Scholar 

  13. Cavagnari, G., Marigonda, A.: Time-optimal control problem in the space of probability measures. In: Large-Scale Scientific Computing, Lecture Notes in Computer Science, vol. 9374, pp. 109–116 (2015)

  14. Cavagnari, G., Marigonda, A., Nguyen, K., Priuli, F.: Generalized control systems in the space of probability measures. Set Valued Var. Anal. 26(3), 663–691 (2018)

    Article  MathSciNet  Google Scholar 

  15. Cosso, A., Pham, H.: Zero-sum stochastic differential games of generalized McKean–Vlasov type. J. Math. Pures Appl. 129, 180–212 (2019)

    Article  MathSciNet  Google Scholar 

  16. Djehiche, B., Hamadène, S.: Optimal control and zero-sum stochastic differential game problems of mean-field type. Appl. Math. Optim. 81, 933–960 (2020)

    Article  MathSciNet  Google Scholar 

  17. Falcone, M.: Numerical methods for differential games based on partial differential equations. Int. Game Theory Rev. 8, 231–272 (2006)

    Article  MathSciNet  Google Scholar 

  18. Falcone, M., Ferretti, R.: Semi-Lagrangian Approximation Schemes for Linear and Hamilton–Jacobi Equations. SIAM, Philadelphia (2013)

    Book  Google Scholar 

  19. Fleming, W.H., Soner, H.M.: Controlled Markov Processes and Viscosity Solutions. Springer, New York (2006)

    MATH  Google Scholar 

  20. Gangbo, W., Mayorga, S., Świȩch, A.: Finite dimensional approximations of Hamilton–Jacobi–Bellman equations in spaces of probability measures. SIAM J. Math. Anal. 53(2), 1320–1356 (2021)

    Article  MathSciNet  Google Scholar 

  21. Huang, M., Malhamé, R., Caines, P.: Nash equilibria for large population linear stochastic systems with weakly coupled agents. In: Boukas, E., Malhamé, R.P. (eds.) Analysis, Control and Optimization of Complex Dynamic Systems, pp. 215–252. Springer, Berlin (2005)

    Chapter  Google Scholar 

  22. Jimenez, C., Marigonda, A., Quincampoix, M.: Optimal control of multiagent systems in the wasserstein space. Calc. Var. Partial Differ. Equ. 59, 58 (2020)

  23. Kolokoltsov, V.N.: Nonlinear Markov Process and Kinetic Equations. Cambridge University Press, Cambridge (2010)

    Book  Google Scholar 

  24. Kolokoltsov, V.N.: Markov Processes, Semigroups and Generators, De Gruyter Studies in Mathematics, vol. 38. De Gryuter, Berlin (2011)

    Google Scholar 

  25. Krasovskii, N.N., Subbotin, A.I.: Game-Theoretical Control Problems. Springer, New York (1988)

    Book  Google Scholar 

  26. Kushner, H.J.: Probability Methods for Approximations in Stochastic Control and for Elliptic Equations. Academic Press, New York (1977)

    MATH  Google Scholar 

  27. Kushner, H.J., Dupuis, P.G.: Numerical Methods for Stochastic Control Problems in Continuous Time. Springer, New York (2001)

    Book  Google Scholar 

  28. Lacker, D.: Limit theory for controlled McKean–Vlasov dynamics. SIAM J. Control Optim. 55, 1641–1672 (2017)

    Article  MathSciNet  Google Scholar 

  29. Lasry, J.M., Lions, P.L.: Jeux à champ moyen. I. Le cas stationnaire (French) [Mean field games. I. The stationary case]. C. R. Math. Acad. Sci. Paris 343, 619–625 (2006)

    Article  MathSciNet  Google Scholar 

  30. Lasry, J.M., Lions, P.L.: Jeux à champ moyen. II. Horizon fini et contrôle optimal (French) [Mean field games. II. Finite horizon and optimal control]. C. R. Math. Acad. Sci. Paris 343, 679–684 (2006)

    Article  MathSciNet  Google Scholar 

  31. Laurière, M., Pironneau, O.: Dynamic programming for mean-field type control. C. R. Math. Acad. Sci. Paris 352(9), 707–713 (2014)

    Article  MathSciNet  Google Scholar 

  32. Marigonda, A., Quincampoix, M.: Mayer control problem with probabilistic uncertainty on initial positions. J. Differ. Equ. 264(5), 3212–3252 (2018)

    Article  MathSciNet  Google Scholar 

  33. Moon, J., Başar, T.: Zero-sum differential games on the Wasserstein space. Commun. Assoc. Inf. Syst. 21(2), 219–251 (2021)

    MathSciNet  MATH  Google Scholar 

  34. Pham, H., Wei, X.: Dynamic programming for optimal control of stochastic McKean–Vlasov dynamics. SIAM J. Control Optim. 55, 1069–1101 (2017)

    Article  MathSciNet  Google Scholar 

  35. Pham, H., Wei, X.: Bellman equation and viscosity solutions for mean-field stochastic control problem. ESAIM Control Optim. Calc. Var. 24(1), 437–461 (2018)

    Article  MathSciNet  Google Scholar 

  36. Sion, M.: On general minimax theorems. Pac. J. Math. 8, 171–176 (1958)

    Article  MathSciNet  Google Scholar 

  37. Subbotin, A.I.: Generalized Solutions of First-Order PDEs. The Dynamical Perspective. Birkhäuser, Boston (1995)

    Book  Google Scholar 

  38. Sznitman, A.: Topics in Propagation of Chaos: Lecture Notes in Mathematics, vol. 1464, pp. 165–251. Springer, Berlin (1991)

    Google Scholar 

  39. Villani, C.: Optimal Transport Old and New. Springer, New York (2009)

    Book  Google Scholar 

Download references

Acknowledgements

The author would like the anonymous referees for their valuable comments and suggestions.

Author information

Authors and Affiliations

Authors

Additional information

Publisher's Note

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

This work was funded by the Russian Science Foundation (Project No. 17-11-01093)

Appendices

Appendix A: Existence and uniqueness result for the flow of probabilities generated by the distribution of players’ controls

Here we give the proof of Proposition 2.2 that is the existence and uniqueness theorem for the flow of probabilities defined by Definition 2.1.

Proof of Proposition 2.2

To derive existence of the flow of probabilities, let us introduce the set of trajectories on [0, T] those are Lipschitz continuous with the constant R that is the upper bound for the norm of f (see (2.4)):

$$\begin{aligned} {\text {Lip}}_R\triangleq \{x(\cdot )\in \mathcal {C}:\Vert x(t')-x(t'')\Vert \le R|t'-t''|,\ \ t',t''\in [0,T]\}. \end{aligned}$$

Obviously, \(x(\cdot ,s,y,\xi ,\zeta )\) lies in \({\text {Lip}}_R\) for every \(y\in \mathbb {T}^d\), \(\xi \in \mathcal {U}\), \(\zeta \in \mathcal {V}\). Moreover, \({\text {Lip}}_R\) is compact. Thanks to the Prokhorov’s theorem, \(\mathcal {P}({\text {Lip}}_R)\) is also compact in the topology of narrow convergence. Consider the mapping \(\Phi \) from \(\{\chi \in \mathcal {P}({\text {Lip}}_R):e_s\sharp \chi = m_*\}\) into itself acting by the rule: if \(\chi \in \mathcal {P}({\text {Lip}}_R)\), then set \(m(t)\triangleq e_t\sharp \chi \) and put

$$\begin{aligned} \Phi [\chi ]\triangleq {\text {traj}}^{s}_{m(\cdot )}\sharp \varkappa . \end{aligned}$$

Since the narrow convergence on \(\mathcal {P}({\text {Lip}}_R)\) is equivalent to the convergence in the Wasserstein metric \(W_2\) [3, Proposition 7.1.5] and \(W_2(e_t\sharp \chi ',e_t\sharp \chi '')\le W_2(\chi ',\chi '')\) for every \(t\in [0,T]\) and \(\chi ',\chi ''\in \mathcal {P}^2(\mathcal {C})\), we have that, if \(\{\chi _n\}_{n=1}^\infty \subset \mathcal {P}({\text {Lip}}_R)\) converges to some \(\chi \in \mathcal {P}({\text {Lip}}_R)\), then the sequence of probabilities \(\{m_n(t)\}_{n=1}^\infty \) defined by the rule \(m_n(t)\triangleq e_t\sharp \chi _n\) converges to \(m(t)\triangleq e_t\sharp \chi \) uniformly in the Wasserstein metric \(W_2\). Using the continuous dependence of the solution of the differential equation on parameter, we obtain that the limit of sequence of operators \(\{{\text {traj}}_{m_n(\cdot )}^{s}\}\) is \({\text {traj}}_{m(\cdot )}^{s}\). Lemma 5.2.1 of [3] implies that the sequence of probabilities \(\{{\text {traj}}_{m_n(\cdot )}^{s}\sharp \varkappa \}\) converges to the distribution \({\text {traj}}_{m(\cdot )}^{s}\sharp \varkappa \). This means that \(\Phi \) is continuous operator acting on compact \(\mathcal {P}({\text {Lip}}_R)\) and, thus, admits a fixed point. It determines \(m(\cdot ,s,m_*,\varkappa )\) through the evaluation operator \(e_t\).

Now let us prove the uniqueness. Let \(m^1(\cdot )\), \(m^2(\cdot )\) be flows of probabilities generated by s, \(m_*\) and \(\varkappa \). Thanks to the Lipschitz continuity of the function f (see (2.3)), we have that

$$\begin{aligned} \begin{aligned}&\Vert x(t,s,y,m^1(\cdot ),\xi ,\zeta )-x(t,s,y,m^2(\cdot ),\xi ,\zeta )\Vert ^2 \\&\quad \le 2L^2\int _s^t \Vert x(t',s,y,m^1(\cdot ),\xi ,\zeta )-x(t',s,y,m^2(\cdot ),\xi ,\zeta )\Vert ^2dt' \\ {}&\qquad +2L^2\int _s^tW_2^2(m^1(t'),m^2(t'))dt'. \end{aligned} \end{aligned}$$

The Gronwall’s inequality gives that

$$\begin{aligned}\begin{aligned}&\Vert x(t,s,y,m^1(\cdot ),\xi ,\zeta )-x(t,s,y,m^2(\cdot ),\xi ,\zeta )\Vert ^2\\ {}&\quad \le C_6\int _s^tW_2^2(m^1(t'),m^2(t'))dt',\end{aligned} \end{aligned}$$

where

$$\begin{aligned} C_6\triangleq 2L^2e^{2L^2T}. \end{aligned}$$

Further, choosing the transportation plan \(\pi ^{1,2}\) between \(m^1(t)\) and \(m^2(t)\) by the rule

$$\begin{aligned} \pi ^{1,2}\triangleq (e_t\circ {\text {traj}}^{s}_{m^1(t)},e_t\circ {\text {traj}}^{s}_{m^2(t)})\sharp \varkappa , \end{aligned}$$

we deduce that

$$\begin{aligned} \begin{aligned}&W_2^2(m^1(t),m^2(t))\le \int _{\mathbb {T}^d\times \mathbb {T}^d}\Vert x^1-x^2\Vert ^2\pi ^{1,2}(d(x^1,x^2))\\&\quad = \int _{\mathbb {T}^d\times \mathcal {U}\times \mathcal {V}}\Vert x(t,s,y,m^1(\cdot ),\xi ,\zeta )-x(t,s,y,m^2(\cdot ),\xi ,\zeta )\Vert ^2\varkappa (d(y,\xi ,\zeta )) \\ {}&\quad \le C_6\int _s^tW_2^2(m^1(t'),m^2(t'))dt'. \end{aligned} \end{aligned}$$

Using the Gronwall’s inequality once more time, we conclude that \(W_2^2(m^1(t), m^2(t))=0\) for every \(t\in [0,T]\). This implies the uniqueness of the flow of probabilities generated by s, \(m_*\) and \(\varkappa \). \(\square \)

Appendix B: A property of Wasserstein metric on a simplex

The purpose of Appendix B is to prove the inequalities between the Wasserstein metric on \(\mathcal {P}(\mathcal {S})\) and the p-th metric on \(\Sigma \) formulated in Proposition 3.1.

Proof of Proposition 3.1

Since the set \(\mathcal {S}\) is finite we have that, for \(\pi \in \Pi (\widetilde{\mu ^1},\widetilde{\mu ^2})\), there exist nonnegative numbers \(b_{{\bar{x}},{\bar{y}}}[\pi ]\), \({\bar{x}},{\bar{y}}\in \mathcal {S}\), such that

$$\begin{aligned} \pi =\sum _{{\bar{x}},{\bar{y}}\in \mathcal {S}}b_{{\bar{x}},{\bar{y}}}[\pi ]\delta _{{\bar{x}},{\bar{y}}}. \end{aligned}$$
(B.1)

Notice that each number \(b_{{\bar{x}},{\bar{y}}}[\pi ]\) is equal to the mass transported from \({\bar{x}}\) to \({\bar{y}}\) according to the transportation plan \(\pi \).

Now let us prove inequality (3.4). Let \(\pi _0\in \Pi (\widetilde{\mu ^1},\widetilde{\mu ^2})\) be such that

$$\begin{aligned} W_p^p\Bigl (\widetilde{\mu ^1},\widetilde{\mu ^2}\Bigr )=\int _{\mathcal {S}\times \mathcal {S}}\Vert x-y\Vert ^p\pi _0(d(x,y)). \end{aligned}$$

From (B.1) it follows that

$$\begin{aligned} W_p^p\Bigl (\widetilde{\mu ^1},\widetilde{\mu ^2}\Bigr )=\sum _{{\bar{x}},{\bar{y}}\in \mathcal {S}}b_{{\bar{x}},{\bar{y}}}[\pi _0]\Vert {\bar{x}}-{\bar{y}}\Vert ^p. \end{aligned}$$

The fact that \(b_{{\bar{x}},{\bar{y}}}[\pi _0]\) is the mass transported from \({\bar{x}}\) to \({\bar{y}}\) according to the plan \(\pi _0\) yields the following inequality:

$$\begin{aligned} \sum _{{\bar{y}}\in \mathcal {S},{\bar{y}}\ne {\bar{x}}}b_{{\bar{x}},{\bar{y}}}[\pi _0]\ge (\mu _{{\bar{x}}}^1-\mu _{{\bar{x}}}^2)^+, \end{aligned}$$

where \(a^+\) is equal to a provided that a is positive, and 0 otherwise. Using this estimate and the definition of \(d(\mathcal {S})\) (see (3.1)), we conclude that

$$\begin{aligned} W_p^p\Bigl (\widetilde{\mu ^1},\widetilde{\mu ^2}\Bigr )\ge \sum _{{\bar{x}}\in \mathcal {S}}\sum _{{\bar{y}}\in \mathcal {S},{\bar{y}}\ne {\bar{x}}}b_{{\bar{x}},{\bar{y}}}[\pi _0](d(\mathcal {S}))^p\ge (d(\mathcal {S}))^p\sum _{{\bar{x}}\in \mathcal {S}}(\mu _{{\bar{x}}}^1-\mu _{{\bar{x}}}^2)^+. \end{aligned}$$

Since \(\mu ^1,\mu ^2\in \Sigma \), we have that

$$\begin{aligned} \sum _{{\bar{x}}\in \mathcal {S}}(\mu _{{\bar{x}}}^1-\mu _{{\bar{x}}}^2)^+=\frac{1}{2}\sum _{{\bar{x}}\in \mathcal {S}}|\mu _{{\bar{x}}}^2-\mu _{{\bar{x}}}^1|. \end{aligned}$$
(B.2)

Furthermore, for every \({\bar{x}}\in \mathcal {S}\), \(|\mu _{{\bar{x}}}^2-\mu _{{\bar{x}}}^1|\le 1\). Thus, we have that

$$\begin{aligned} W_p^p\Bigl (\widetilde{\mu ^1},\widetilde{\mu ^2}\Bigr )\ge \frac{(d(\mathcal {S}))^p}{2}\sum _{{\bar{x}}\in \mathcal {S}}|\mu _{{\bar{x}}}^2-\mu _{{\bar{x}}}^1|\ge \frac{(d(\mathcal {S}))^p}{2}\sum _{{\bar{x}}\in \mathcal {S}}|\mu _{{\bar{x}}}^2-\mu _{{\bar{x}}}^1|^p. \end{aligned}$$

This proves (3.4).

To prove inequality (3.5), choose a probability \(\pi '\in \Pi (\widetilde{\mu ^1},\widetilde{\mu ^2})\) such that

$$\begin{aligned} b_{{\bar{x}},{\bar{x}}}[\pi ']\triangleq \mu ^1_{{\bar{x}}}\wedge \mu _{{\bar{x}}}^2. \end{aligned}$$

Here \(b_{{\bar{x}},{\bar{y}}}[\pi ']\) are given by representation (B.1). Therefore, for every \({\bar{x}}\in \mathcal {S}\)

$$\begin{aligned} \sum _{{\bar{y}}\in \mathcal {S},{\bar{y}}\ne {\bar{x}}}b_{{\bar{x}},{\bar{y}}}[\pi ']=(\mu _{{\bar{x}}}^1-\mu _{{\bar{x}}}^2)^+. \end{aligned}$$
(B.3)

The inclusion \(\mathcal {S}\subset \mathbb {T}^d\) yields

$$\begin{aligned} W_p^p\Bigl (\widetilde{\mu ^1},\widetilde{\mu ^2}\Bigr )\le \sum _{{\bar{x}},{\bar{y}}\in \mathcal {S}} \Vert {\bar{x}}-{\bar{y}}\Vert ^pb_{{\bar{x}},{\bar{y}}}[\pi ']\le d^{p/2}\sum _{{\bar{x}}\in \mathcal {S}}\sum _{{\bar{y}}\in \mathcal {S},{\bar{y}}\ne {\bar{x}}}b_{{\bar{x}},{\bar{y}}}[\pi ']. \end{aligned}$$

This and equalities (B.2), (B.3) imply that

$$\begin{aligned} W_p^p\Bigl (\widetilde{\mu ^1},\widetilde{\mu ^2}\Bigr )\le \frac{d^{p/2}}{2}\sum _{{\bar{x}}\in \mathcal {S}} |\mu _{{\bar{x}}}^2-\mu _{{\bar{x}}}^1|=\frac{d^{p/2}}{2}\Vert \mu ^2-\mu ^1\Vert _1. \end{aligned}$$

The Holder’s inequality gives that

$$\begin{aligned} W_p^p\Bigl (\widetilde{\mu ^1},\widetilde{\mu ^2}\Bigr )\le \frac{d^{p/2+p'}}{2}\Vert \mu ^2-\mu ^1\Vert _p, \end{aligned}$$

where \(p'\) is such that \(1/p'+1/p=1\). This proves (3.5) \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Averboukh, Y. Lattice approximations of the first-order mean field type differential games. Nonlinear Differ. Equ. Appl. 28, 65 (2021). https://doi.org/10.1007/s00030-021-00727-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00030-021-00727-2

Keywords

Mathematics Subject Classification

Navigation