Skip to main content
Log in

Continuous Time Finite State Mean Field Games

  • Published:
Applied Mathematics & Optimization Submit manuscript

Abstract

In this paper we consider symmetric games where a large number of players can be in any one of d states. We derive a limiting mean field model and characterize its main properties. This mean field limit is a system of coupled ordinary differential equations with initial-terminal data. For this mean field problem we prove a trend to equilibrium theorem, that is convergence, in an appropriate limit, to stationary solutions. Then we study an N+1-player problem, which the mean field model attempts to approximate. Our main result is the convergence as N→∞ of the mean field model and an estimate of the rate of convergence. We end the paper with some further examples for potential mean field games.

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.

Institutional subscriptions

Similar content being viewed by others

References

  1. Achdou, Y., Dolcetta, I.C.: Mean field games: numerical methods. Preprint (2010)

  2. Adlakha, S., Johari, R., Weintraub, G.Y.: Equilibria of dynamic games with many players: existence, approximation, and market structure. Submitted for publication (2011)

  3. Salomon, J., Lachapelle, A., Turinici, G.: Computation of mean field equilibria in economics

  4. Cardaliaguet, P.: Notes on Mean Field Games (2010)

    Google Scholar 

  5. Gomes, D., Mohr, J., Souza, R.R.: Discrete time, finite state space mean field games. J. Math. Pures Appl. 93(2), 308–328 (2010)

    MathSciNet  MATH  Google Scholar 

  6. Gomes, D., Pires, G.E., Sanchez-Morgado, H.: A-priori estimates for stationary mean-field games. Preprint (2011)

  7. Gomes, D., Sanchez-Morgado, H.: A stochastic Evans-Aronsson problem. Preprint (2011)

  8. Gueant, O.: Mean field games and applications to economics. Ph.D. thesis, Université Paris Dauphine, Paris (2009)

  9. Gueant, O.: An existence and uniqueness result for mean field games with congestion effect on graphs. Preprint (2011)

  10. Gueant, O.: From infinity to one: the reduction of some mean field games to a global control problem. Preprint (2011)

  11. Huang, M., Caines, P.E., Malhamé, R.P.: Large-population cost-coupled LQG problems with nonuniform agents: individual-mass behavior and decentralized ϵ-Nash equilibria. IEEE Trans. Autom. Control 52(9), 1560–1571 (2007)

    Article  Google Scholar 

  12. Huang, M., Malhamé, R.P., Caines, P.E.: Large population stochastic dynamic games: closed-loop McKean-Vlasov systems and the Nash certainty equivalence principle. Commun. Inf. Syst. 6(3), 221–251 (2006)

    MathSciNet  MATH  Google Scholar 

  13. Hayel, Y., Tembine, H., Altman, E., El-Azouzi, R.: A Markov decision evolutionary game for individual energy management. In: Advances in Dynamic Games. Ann. Internat. Soc. Dynam. Games, vol. 11, pp. 313–335. Birkhäuser/Springer, New York (2011)

    Chapter  Google Scholar 

  14. Kapur, S.: Markov perfect equilibria in an N-player war of attrition. Econ. Lett. 47(2), 149–154 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  15. Kolokoltsov, V.N., Li, J., Yang, W.: Mean field games and nonlinear Markov processes. Preprint (2011)

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

    MATH  Google Scholar 

  17. Livshits, I.: On non-existence of pure strategy Markov perfect equilibrium. Econ. Lett. 76(3), 393–396 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  18. Lasry, J.-M., Lions, P.-L.: Jeux à champ moyen. I. Le cas stationnaire. C. R. Math. Acad. Sci. Paris 343(9), 619–625 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  19. Lasry, J.-M., Lions, P.-L.: Jeux à champ moyen. II. Horizon fini et contrôle optimal. C. R. Math. Acad. Sci. Paris 343(10), 679–684 (2006)

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  21. Lasry, J.-M., Lions, P.-L.: Mean Field Games. Cahiers de la Chaire Finance et Développement Durable (2007)

    Google Scholar 

  22. Lasry, J.-M., Lions, P.-L., Guéant, O.: Application of mean field games to growth theory. Preprint (2010)

  23. Lasry, J.-M., Lions, P.-L., Guéant, O.: Mean Field Games and Applications. Paris-Princeton Lectures on Mathematical Finance (2010)

    Google Scholar 

  24. Moll, B., Lucas, R.: Knowledge growth and the allocation of time. Preprint (2011)

  25. Maskin, E., Tirole, J.: Markov perfect equilibrium. I. Observable actions. J. Econ. Theory 100(2), 191–219 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  26. Pakes, A., McGuire, P.: Stochastic algorithms, symmetric Markov perfect equilibrium, and the “curse” of dimensionality. Econometrica 69(5), 1261–1281 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  27. Pichler, P., Sorger, G.: Wealth distribution and aggregate time-preference: Markov-perfect equilibria in a Ramsey economy. J. Econ. Dyn. Control 33(1), 1–14 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  28. Sleet, C.: Markov perfect equilibria in industries with complementarities. Econom. Theory 17(2), 371–397 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  29. Streufert, P.A.: Markov-perfect equilibria in intergenerational games with consistent preferences. J. Econ. Dyn. Control 17(5–6), 929–951 (1993)

    Article  MathSciNet  MATH  Google Scholar 

  30. Tembine, H., Le Boudec, J.Y., ElAzouzi, R., Altman, E.: In: Proceedings of GameNets, Istanbul, Turkey (2009)

    Google Scholar 

  31. Weintraub, G.Y., Lanier Benkard, C., Van Roy, B.: Markov perfect industry dynamics with many firms. Econometrica 76(6), 1375–1411 (2008)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Diogo A. Gomes.

Additional information

D. Gomes was partially supported by CAMGSD-LARSys through FCT-Portugal and by grants PTDC/MAT-CAL/0749/2012, UTA-CMU/MAT/0007/2009 PTDC/MAT/114397/2009, UTAustin-MAT/0057/2008, and by the bilateral agreement Brazil-Portugal (CAPES-FCT) 248/09.

R.R.S. was partially supported by the bilateral agreement Brazil-Portugal (CAPES-FCT) 248/09.

J.M. was partially supported by the bilateral agreement Brazil-Portugal (CAPES-FCT) 248/09.

Appendix: Auxiliary Results

Appendix: Auxiliary Results

Proof of Proposition 1

To prove the first item we use the definition of h and α and also that \(v^{i}\sum_{j} \alpha^{*}_{j}(z,\theta,i)=0\) to get

$$h(z,\theta,i)+\sum_j \alpha^*_j(z, \theta,i)v^j= c\bigl(i,\theta ,\alpha ^*(z,\theta,i)\bigr)+\sum _j \alpha^*_j(z,\theta,i) \bigl(z^j+v^j-z^i-v^i\bigr). $$

Hence by the definition of h(z+v,θ,i) we have

$$h(z,\theta,i)+\sum_j \alpha^*_j(z, \theta,i)v^j\geq h(z+v,\theta,i). $$

From this (7) holds and we deduct that if h is differentiable

$$\alpha^*_j=\frac{\partial h(\Delta_iz,\theta,i) }{\partial z^j}. $$

Note that item (c) is a direct corollary of item (b), since

$$ h(p,\theta,i)=c\bigl(i,\theta,\alpha^*(p, \theta,i)\bigr)+\alpha^*(p,\theta ,i)\cdot p $$
(63)

and the function c is Lipschitz in θ and differentiable in α.

From this point on in this proof we will omit the index i as it is not relevant and simplifies the notation. To prove item (b) we will use the following inequalities, which are consequence of the uniform convexity of c: for all \(\theta, \theta '\in\mathcal{S}^{d}\), \(\alpha', \alpha\in(\mathbb{R}_{0}^{+})^{d}\), \(\sum_{k} \alpha_{k}=\sum_{k} \alpha_{k}'=0\), and \(p,p'\in\mathbb{R}^{d}\), we have

$$ c\bigl(\theta,\alpha'\bigr)+\alpha' \cdot p '\geq c(\theta,\alpha)+\alpha\cdot p'+\bigl( \nabla_\alpha c(\theta,\alpha )+p '\bigr)\cdot\bigl( \alpha'-\alpha\bigr)+\gamma\|\alpha'-\alpha \|^2, $$
(64)

and because α (p,θ) is a minimizer,

$$ \bigl(\nabla_\alpha c\bigl(\theta,\alpha ^*(p,\theta) \bigr)+p\bigr)\cdot\bigl(\alpha'-\alpha^*(p,\theta)\bigr)\geq0. $$
(65)

We will first prove that α is uniformly Lipschitz in p: for that, we suppose that θ is fixed. By the definition of α and (64) we have

hence

Now using (65) we obtain

$$0\geq\bigl(p'-p\bigr)\cdot\bigl(\alpha^*\bigl(p' \bigr)-\alpha^*(p)\bigr)+\gamma\bigl\Vert\alpha ^*\bigl(p '\bigr)- \alpha^*(p)\bigr\Vert^2. $$

Therefore

$$\|p-p'\|\bigl\Vert\alpha^*\bigl(p'\bigr)-\alpha^*(p)\bigr\Vert\geq \gamma\bigl\Vert\alpha ^*\bigl(p '\bigr)-\alpha^*(p)\bigr\Vert^2, $$

which implies

$$\bigl\Vert\alpha^*\bigl(p'\bigr)-\alpha^*(p)\bigr\Vert\leq\frac{1}{\gamma} \|p'-p \|. $$

This shows that α is uniformly Lipschitz in p.

Now we prove that α is Lipschitz in θ: for that, we suppose that p is fixed. Again by the definition of α and by (64) we have

and then

$$0\geq\bigl(\nabla_\alpha c\bigl(\theta',\alpha^*(\theta) \bigr)+p\bigr)\cdot\bigl(\alpha ^*\bigl(\theta'\bigr)-\alpha^*( \theta)\bigr)+\gamma\bigl\Vert\alpha^*\bigl(\theta '\bigr)-\alpha ^*(\theta) \bigr\Vert^2. $$

Using (65) we get

$$0\geq \bigl[\nabla_\alpha c\bigl(\theta',\alpha^*(\theta) \bigr)-\nabla _\alpha c\bigl(\theta,\alpha^*(\theta)\bigr) \bigr]\cdot \bigl(\alpha^*\bigl(\theta'\bigr)-\alpha ^*(\theta)\bigr) +\gamma\bigl\Vert \alpha^*\bigl(\theta'\bigr)-\alpha^*(\theta) \bigr\Vert^2. $$

As ∇ α c(θ,α) is Lipschitz in the variable θ we have

$$K_c\|\theta'-\theta\|\bigl\Vert\alpha^*(\theta)-\alpha^* \bigl(\theta'\bigr)\bigr\Vert \geq \gamma\bigl\Vert\alpha^*\bigl( \theta'\bigr)-\alpha^*(\theta) \bigl\Vert^2. $$

Therefore

$$\bigl\Vert\alpha^*(\theta)-\alpha^*\bigl(\theta'\bigr)\bigl\Vert\leq \frac{K_c}{\gamma } \| \theta-\theta' \|, $$

which implies that α is Lipschitz in θ. □

Proof of Theorem 5

The main tool for proving Theorem 5 is once again the Dynkin Formula, now adapted to the present situation: suppose α and β are two admissible controls.

We recall the infinitesimal generator of the process (i,n) defined in (50). We have

(66)

where γ β is defined by (30).

We have that, for any function \(\varphi:I_{d} \times\mathcal{S}^{d}_{N} \times [0,+\infty) \rightarrow{\mathbb{R}}\), C 1 in the last variable, and any t<T,

$$ \mathbb{E}^{\beta,\alpha}_{A_t(i, n)} \bigl[ \varphi^{{\mathbf{i}}_T}_{{\mathbf{n}}_T}(T)-\varphi ^i_n(t) \bigr] = \mathbb{E}^{\beta,\alpha}_{A_t(i, n)} \biggl[ \int _t^T \frac{d\varphi^{{\mathbf{i}}_s}_{{\mathbf{n}}_s}}{dt} (s) + \bigl(A^{\beta,\alpha }\varphi\bigr)^{{\mathbf{i}} _s}_{{\mathbf{n}}_s}(s) ds \biggr], $$
(67)

where A t (i,n) denotes the event i t =i and n t =n.

Now we prove the theorem. In the Dinkyn formula (67) let φ=v. Using the terminal condition \(v^{{\mathbf{i}}_{T}}_{{\mathbf {n}}_{T}}(T)=\psi^{{\mathbf{i}} _{T}} (\frac{{\mathbf{n}}_{T}}{N} )\) we have that, for any admissible control α,

$$ \mathbb{E}^{\beta,\alpha}_{A_t(i,n)} \biggl[ \psi^{{\mathbf {i}}_T} \biggl(\frac{{\mathbf{n}} _T}{N} \biggr)-v^{i}_{n}(t) \biggr] = \mathbb{E}^{\beta,\alpha}_{A_t(i,n)} \biggl[ \int _t^T \frac{d v^{{\mathbf{i}}_s}_{{\mathbf{n}}_s} }{dt}(s) + \bigl(A^{\beta ,\alpha}v\bigr)^{{\mathbf{i}} _s}_{{\mathbf{n}} _s}(s) ds \biggr]. $$
(68)

In the next steps we will use the definition of u, given in (31), and then (68), (66), (4) to have

where the last equation holds because v is a solution to the Hamilton-Jacobi equation (32). Note that in this last calculation we are also proving that, for the specific control \(\tilde{\alpha}\) given by (34), we have \(u^{i}_{n}(t,\beta,\tilde{\alpha})=v^{i}_{n}(t)\) which show us that \(\tilde{\alpha}\) is the optimal control and that the objective function \(u^{i}_{n}(t,\beta)\) is given by \(v_{n}^{i}(t)\). □

Proof of Proposition 6

Let u be a solution to (32). Let \(\tilde{u}=u+\rho(T-t)\). Then

$$- \frac{d\tilde{u}_n^i}{dt} =\rho+ \sum_{k,j} \gamma_{\beta,kj}^{n,i} \bigl(\tilde{u}^i_{n+e_{jk}}- \tilde{u}^i_n\bigr) + h \biggl( \Delta_i \tilde{u}_n,\frac{n}{N},i \biggr). $$

Let (i,n,t) be a minimum point of \(\tilde{u}\) on \(I_{d}\times\mathcal {S}^{d}_{N} \times[0,T]\). We have \(\tilde{u}^{i}_{n+e_{jk}}\geq\tilde{u}^{i}_{n}\). This implies \(\gamma_{\beta,kj}^{n,i} (\tilde{u}^{i}_{n+e_{jk}}-\tilde{u}^{i}_{n})\geq0\). We also have \(\tilde{u}^{j}_{n}(t)-\tilde{u}^{i}_{n}(t)\geq0\) hence \(\Delta_{i} \tilde{u}_{n}= (\tilde{u}^{1}_{n}(t)-\tilde{u}^{i}_{n}(t),\ldots,\tilde{u}^{d}_{n}(t)-\tilde{u}^{i}_{n}(t))\geq0\). Hence

$$-\frac{d\tilde{u}_n^i}{dt}(t) \geq h \biggl( \Delta_i \tilde{u}_n, \frac{n}{N}, i \biggr)+\rho\geq h \biggl(0,\frac{n}{N}, i \biggr)+ \rho, $$

because the definition of h i p,θ,i), with Δ i p≥0. Furthermore, if we take M<ρ<2M we get

$$-\frac{d\tilde{u}_n^i}{dt}(t)>0. $$

This shows that the minimum of \(\tilde{u}\) is achieved at T hence

$$u_n^i(t) \geq-\|u(T)\|_{\infty}-2M (T-t). $$

Similarly, let (i,n,t) be a maximum point of \(\tilde{u}\) on \(I_{d}\times\mathcal {S}^{d}_{N}\times[0,T]\). We have \(\tilde{u}^{i}_{n+e_{jk}}\leq\tilde{u}^{i}_{n}\). This implies \(\gamma_{\beta,kj}^{n,i} (\tilde{u}^{i}_{n+e_{jk}}-\tilde{u}^{i}_{n})\leq0\). We also have \(\Delta_{i} \tilde{u}_{n}\leq0\). Hence

$$-\frac{d\tilde{u}_n^i}{dt}(t) \leq h \biggl(\Delta_i \tilde{u}_n, \frac{n}{N}, i \biggr)+\rho\leq h \biggl(0,\frac{n}{N}, i \biggr)+ \rho. $$

Furthermore, if we take −2M<ρ<−M we get

$$-\frac{d\tilde{u}_n^i}{dt}(t)<0. $$

This shows that the maximum of \(\tilde{u}\) is achieved at T hence

$$u_n^i(t)\leq\|u(T)\|_{\infty}+2M (T-t). $$

 □

Proof Lemma 3

Recall that α (p,θ,i) is Lipschitz in (p,θ). Let K be the corresponding Lipschitz constant. Since ∥p∥ bounded, we have |α (p,.,.)|≤C. Then

 □

Proof of Lemma 4

Let z be a solution of (40), and fix ϵ>0. We define \(\tilde{z}= z+\epsilon(t-\nobreak T)\). Hence \(\tilde{z}\) satisfies

$$-\dot{\tilde{z}}_m=-\epsilon+\sum_{m'\in I_d\times{\mathcal {S}^d_N}}a_{mm'}(t) ({\tilde{z}}_{m'}-{\tilde{z}}_m). $$

Let (m,t) be a maximum point of \(\tilde{z}\) on \(I_{d}\times{\mathcal {S}^{d}_{N}}\times[0,T]\). We have \({\tilde{z}}_{m}(t)\geq{\tilde{z}}_{m'}(t)\) and this implies \(a_{mm'}(t) ({\tilde{z}}_{m'}-{\tilde{z}}_{m})\leq0\)m′. Hence

$$-\frac{d\tilde{z}_m}{dt}(t) \leq-\epsilon. $$

This shows that the maximum of \(\tilde{z}\) is achieved at T. Therefore, for all (m′,t),

$$z_{m'}(t)+\epsilon(t-T)=\tilde{z}_{m'}(t)\leq\tilde{z}_m(T)=z_m(T). $$

Letting ϵ→0, we get

$$z_{m'}(t)\leq\max_{m} z_m(T),\quad \forall\bigl(m',t\bigr). $$

From this inequality we have the following conclusions:

  1. 1.

    if z(T)≤0, we then have z m(t)≤0, for all (m′,t), and so z(t)≤0;

  2. 2.

    for all (m′,t),

    $$z_{m'}(t)\leq\|z(T)\|_{\infty}. $$

Now we define \(\tilde{z}= z+\epsilon(T-t)\). Hence \(\tilde{z}\) satisfies

$$-\dot{\tilde{z}}_m=\epsilon+\sum_{m'\in I_d\times{\mathcal {S}^d_N}}a_{mm'}(t) ({\tilde{z}}_{m'}-{\tilde{z}}_m). $$

Let (m,t) be a minimum point of \(\tilde{z}\) on \(I_{d}\times{\mathcal {S}^{d}_{N}}\times[0,T]\). We have \(a_{mm'}(t) ({\tilde{z}}_{m'}-{\tilde{z}}_{m})\geq0\). Therefore we have

$$-\frac{d\tilde{z}_m}{dt}(t) \geq\epsilon. $$

This shows that the minimum of \(\tilde{z}\) is also achieved at T, hence for all (m′,t) we have

$$z_{m'}(t)+\epsilon(T-t)=\tilde{z}_{m'}(t)\geq\tilde{z}_m(T)=z_m(T). $$

Letting ϵ→0, we get z m(t)≥min m z m (T). Hence

$$z_{m'}(t)\geq- \|z(T)\|_{\infty}, $$

and therefore we have ∥z(t)∥≤∥z(T)∥. □

Proof of Lemma 6

We note that if tsT we have K(t,s)K(s,T)=K(t,T), which implies

$$\frac{d}{ds} \bigl( K(t,s)K(s, T) \bigr)=0. $$

Hence, using (42) we get

$$-K(t,s)M(s)K(s,T)+ \biggl(\frac{d}{ds} K(t,s) \biggr) K(s,T)=0, $$

and therefore, by taking T=s we conclude that

$$ \frac{d}{ds} K(t,s)=K(t, s)M(s). $$
(69)

Multiplying (43) by K(t,s) and using Lemma 5, we have

$$-K(t,s)\dot{z}(s)\leq K(t,s)M(s) z(s) +K(t,s)f\bigl(z(s)\bigr). $$

Using the identity

$$\frac{d}{ds} K(t,s)z(s)=K(t,s) \dot{z}(s)+ K(t, s) M(s) z(s), $$

which follows from (69), we get

$$-\frac{d}{ds} \bigl(K(t,s)z(s) \bigr) \leq K(t, s)f\bigl(z(s)\bigr). $$

Thus, integrating between t and T, we have

$$z(t)-K(t, T)z(T)\leq\int_t^T K(t, s)f \bigl(z(s)\bigr) ds. $$

Note that if z(t)=K(t,T)z(T) is a solution of (40) with terminal data z(T)=b, then Lemma 4 implies that ∥z(t)∥≤∥z(T)∥, hence ∥K(t,T)z(T)∥≤∥z(T)∥.

Therefore for all \(m\in I_{d}\times{\mathcal{S}}^{d}_{N}\) we have

$$z_m(t)\leq\|z(T)\|_{\infty}+\int_t^T \bigl\Vert f\bigl(z(s)\bigr)\bigr\Vert_{\infty}ds. $$

 □

Proof of Lemma 7

Note that (44) implies that v is a monotone decreasing function of s and is equivalent to

$$\everymath{\displaystyle} \begin{cases} \frac{ds}{d v}= \frac{-1}{Cv+C N v^2 + \frac{C}{N}},\\ s\biggl(\frac{C}{N}\biggr)\leq T. \end{cases} $$

This implies by direct integration that

$$s \biggl(\frac{2C}{N} \biggr) \leq T - \int _{\frac{C}{N}}^{\frac{2C}{N}} \frac{dv}{Cv+C N v^2 + \frac {C}{N}}. $$

Now

$$\int_{\frac{C}{N}}^{\frac{2C}{N}} \frac{dv}{Cv+C N v^2 + \frac{C}{N}} \geq \int _{\frac{C}{N}}^{\frac{2C}{N}} \frac{N}{2C^2+4C^3+C}dv = \frac {1}{2C+4C^2+1}. $$

Therefore if we define \(T^{\star}=\frac{1}{2C+4C^{2}+1} \), we have that \(s (\frac{2C}{N} ) \leq0\) if TT . Hence this implies \(v(0) \leq\frac{2C}{N}\), which yields the desired result when we take into account that v is a decreasing function of s. □

Rights and permissions

Reprints and permissions

About this article

Cite this article

Gomes, D.A., Mohr, J. & Souza, R.R. Continuous Time Finite State Mean Field Games. Appl Math Optim 68, 99–143 (2013). https://doi.org/10.1007/s00245-013-9202-8

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00245-013-9202-8

Keywords

Navigation