Skip to main content
Log in

Long Time Average of First Order Mean Field Games and Weak KAM Theory

  • Published:
Dynamic Games and Applications Aims and scope Submit manuscript

Abstract

We show that the long time average of solutions of first order mean field game systems in finite horizon is governed by an ergodic system of mean field game type. The well-posedness of the latter system and the uniqueness of the ergodic constant rely on weak KAM theory.

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. Cardaliaguet P, Lasry J-M, Lions P-L, Porretta A (2012) Long time average of mean field games. Netw Heterog Media 7(2):279–301

    Article  MathSciNet  MATH  Google Scholar 

  2. Cardaliaguet P, Lasry J-M, Lions P-L, Porretta A Long time average of mean field games with a nonlocal coupling. SIAM J Control Optim (to appear)

  3. Fathi A (1997) Théorme KAM faible et théorie de Mather sur les systemes lagrangiens. C R Acad Sci Paris Sér I Math 324(9):1043–1046

    Article  MathSciNet  MATH  Google Scholar 

  4. Fathi A (1997) Solutions KAM faibles conjuguées et barrires de Peierls. C R Acad Sci Paris Sér I Math 325(6):649–652

    Article  MathSciNet  MATH  Google Scholar 

  5. Fathi A (2005) Weak KAM theorem and Lagrangian dynamics. Seventh preliminary version. Preprint

  6. Gomes DA, Mohr J, Souza R (2010) Discrete time, finite state space mean field games. J Math Pures Appl (9) 93(3):308–328

    Article  MathSciNet  MATH  Google Scholar 

  7. Huang M, Caines PE, Malhamé RP (2006) Large population stochastic dynamic games: closed-loop McKean–Vlasov systems and the Nash certainty equivalence principle. Commun Inf Syst 6(3):221–252

    MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  11. Lions PL (1982) Generalized solution of Hamilton–Jacobi equations. Pitman, London

    Google Scholar 

  12. Lions PL Cours au Collège de France. www.college-de-france.fr

  13. Lions P-L, Papanicolau G, Varadhan SRS (1987) Homogenization of Hamilton–Jacobi equation. Unpublished preprint

  14. Namah G, Roquejoffre J-M (1999) Remarks on the long time behaviour of the solutions of Hamilton–Jacobi equations. Commun Partial Differ Equ 24(5–6):883–893

    Article  MathSciNet  MATH  Google Scholar 

  15. Roquejoffre J-M (1998) Comportement asymptotique des solutions d’équations de Hamilton–Jacobi monodimensionnelles. C R Acad Sci Paris Sér I Math 326(12):185–189

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

We wish to thank Yves Achdou for fruitful discussions.

This work has been partially supported by the Commission of the European Communities under the 7th Framework Programme Marie Curie Initial Training Networks Project SADCO, FP7-PEOPLE-2010-ITN, No. 264735, and by the French National Research Agency ANR-10-BLAN 0112 and ANR-12-BS01-0008-01.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to P. Cardaliaguet.

Appendix: Proof of the existence and uniqueness result

Appendix: Proof of the existence and uniqueness result

The following result is stated in [10]. For convenience of the reader we recall the main ingredients of the proof.

Theorem 5.1

[10]

Let H and F satisfy conditions (4) and (7). Then (1) has a solution. If moreover the following inequality holds:

$$\int_{\mathbb {T}^d} \bigl(F(x,m_1)-F(x,m_2) \bigr)d(m_1-m_2) \geq0\quad\forall m_1,m_2 \in P\bigl(\mathbb {T}^d\bigr), $$

then the solution of (1) is unique.

Proof

The proof is based on a vanishing viscosity argument. Let ϵ>0 and (u ϵ,m ϵ) be the solution to

$$ \left \{ \begin{array}{r@{\quad}l@{\quad}l} (\mathrm{i})& -\partial_t u^\epsilon -\epsilon \Delta u^\epsilon +H\bigl(x,Du^\epsilon \bigr) =F\bigl(x,m^\epsilon (t)\bigr)&\mbox{in}\ (0,T)\times \mathbb{R}^d,\\ (\mathrm{ii}) & \partial_t m^\epsilon -\epsilon \Delta m^\epsilon -\operatorname{div} \bigl(m^\epsilon D_pH\bigl(x,Du^\epsilon \bigr)\bigr)=0& \mbox{in}\ (0,T)\times \mathbb{R}^d,\\ (\mathrm{iii})& m^T(0)=m_0, \qquad u^\epsilon (x,T)=u^f(x)& \operatorname{in}\ \mathbb{R}^d. \end{array} \right . $$
(19)

By (7) and using standard regularity results for parabolic equations and fixed point arguments, it is not difficult to check that system (19) has at least one classical solution. Following Lemma 5.2 below, the (u ϵ) are uniformly Lipschitz continuous and semi-concave in space with a semi-concavity modulus independent of ϵ. We need two other estimates: the first one is for the (m ϵ) and the second one for the time dependence of the (u ϵ).

Let us first show that the (m ϵ) are uniformly bounded. For this we note that

$$\begin{aligned} &\operatorname{div} \bigl(m^\epsilon D_pH\bigl(x,Du^\epsilon \bigr)\bigr)\\ &\quad= \big \langle Dm^\epsilon , D_pH\bigl(x,Du^\epsilon \bigr)\big \rangle + m^\epsilon \operatorname{Tr} \bigl(D^2_{xp}H\bigl(x,Du^\epsilon \bigr)+ D^2_{pp}H\bigl(x,Du^\epsilon \bigr)D^2u^\epsilon \bigr)\\ &\quad\leq\big \langle Dm^\epsilon , D_pH\bigl(x,Du^\epsilon \bigr)\big \rangle + Cm^\epsilon \end{aligned}$$

because \(D^{2}_{xp}H(x,Du^{\epsilon })\) is bounded thanks to the regularity of H and the uniform Lipschitz continuity of u ϵ, and \(\operatorname{Tr} (D^{2}_{pp}H(x,Du^{\epsilon })D^{2}u^{\epsilon })\) is bounded above because \(D^{2}_{pp}H\) is positive and u ϵ is uniformly semi-concave. So m ϵ is a subsolution of the transport equation

$$\partial_t m^\epsilon -\epsilon \Delta m^\epsilon -\big \langle Dm^\epsilon , D_pH\bigl(x,Du^\epsilon \bigr) \big \rangle -Cm^\epsilon =0\quad \mbox{in}\ (0,T)\times \mathbb{R}^d. $$

By the maximum principle we get ∥m ϵ≤∥m 0 e CT. Next we claim that the map tm ϵ(t) (as a map from [0,T] to \(\mathbb {P}(\mathbb {T}^{d})\)) is uniformly Hölder continuous: indeed, if we multiply (19)-(ii) by m ϵ and integrate in time-space, we get

$$\epsilon \int_0^T\int_{\mathbb {T}^d} \big|Dm^\epsilon \big|^2 \leq\frac{1}{2} \int_{\mathbb {T}^d} \bigl(\big|m^2(T)\big|+\big|m^2(0)\big|\bigr) + \int_0^T \int_{\mathbb {T}^d} m^\epsilon \big|Dm^\epsilon \big|\big|D_pH \bigl(x,Du^\epsilon \bigr)\big|. $$

As m ϵ and D p H(x,Du ϵ) are uniformly bounded, this implies that

$$ \epsilon \biggl(\int_0^T\int _{\mathbb {T}^d} \big|Dm^\epsilon \big|^2 \biggr)^{\frac{1}{2}} \leq C. $$
(20)

Then, for any smooth test function \(\varphi:\mathbb {T}^{d}\to \mathbb{R}\) and for any 0≤t 1t 2T, we have by (19)-(ii):

$$\begin{aligned} \int_{\mathbb {T}^d} \varphi\bigl(m^\epsilon (t_2)-m^\epsilon (t_1) \bigr) =& - \int_{t_1}^{t_2}\int_{\mathbb {T}^d} \epsilon \big \langle Dm^\epsilon , D\phi\big \rangle + m^\epsilon \big \langle D_pH(x,Du), D\phi\big \rangle \\ \leq& C(t_2-t_1)^{\frac{1}{2}}\|D\varphi \|_\infty, \end{aligned}$$

because Du ϵ and m ϵ are uniformly bounded and (20) holds. Taking the supremum over all 1-Lipschitz continuous maps φ gives \(\mathbf{d}_{1}(m(t_{1}),m(t_{2}))\leq C(t_{2}-t_{1})^{\frac{1}{2}}\).

Our last estimate is a uniform continuity in time of the (u ϵ). For this we note that, as u f(⋅) is bounded in \({\mathcal{C}}^{2}\), the maps w ±(x,t)=u f(xC 1(Tt) are sub (for ±=−) and super (for ±=+) solution of (19)-(i) for C 1 sufficiently large (but not depending of ϵ). Hence ∥u ϵ(⋅,t)−u f(⋅)∥C 1(Tt). For h>0, we consider \(u^{\epsilon }_{h}(x,t)= u(x, t-h)\). Because of the uniform Hölder regularity of the map tm ϵ(t) in \(P(\mathbb {T}^{d})\) and the uniform continuity of F, there is η(h)→0 as h→0 uniformly in ϵ, such that

$$\sup_{t\in[h,T]}\bigl \Vert F\bigl(\cdot,m^\epsilon (t-h) \bigr)-F\bigl(\cdot,m^\epsilon (t)\bigr)\bigr \Vert _\infty\leq \eta(h). $$

So u h satisfies

$$-\partial_t u^\epsilon _h -\epsilon \Delta u^\epsilon _h +H\bigl(x,Du^\epsilon _h\bigr) =F\bigl(x,m^\epsilon (t-h)\bigr)\leq F\bigl(x,m^\epsilon (t)\bigr)+ \eta(h) $$

with terminal condition \(u^{\epsilon }_{h}(x,T)= u^{\epsilon }(T-h,x)\leq u^{f}(x)+C_{1}h\). By comparison, we get \(u^{\epsilon }_{h}(x,t)-\eta(h)(T-t)-C_{1}h\leq u^{\epsilon }(t,x)\), i.e.,

$$u^\epsilon (x,t-h)\leq u^\epsilon (t,x)+\eta(h) (T-t)+C_1h. $$

In a symmetric way,

$$u^\epsilon (x,t-h)\geq u^\epsilon (t,x)-\eta(h) (T-t)-C_1h. $$

Therefore ∥u ϵ(⋅,th)−u ϵ(⋅,t)∥η(h)T+C 1 h, which proves the uniform continuity of u ϵ.

Because of the bounds on (u ϵ,m ϵ), we can assume that (up subsequences) u ϵ converges uniformly to some continuous map u. Moreover, by uniform semi-concavity, (Du ϵ) converges a.e. to Du. On another hand m ϵ converges in L -weak* and in \({\mathcal{C}}^{0}([0,T],P(\mathbb {T}^{d}))\) to some \(m\in L^{\infty}\cap{\mathcal{C}}^{0}([0,T],P(\mathbb {T}^{d}))\). In particular, m(0)=m 0. Using the continuity assumption (4), we also find that F(⋅,m ϵ(⋅)) converges uniformly to F(⋅,m(⋅)). By standard viscosity solutions arguments, we can conclude to the convergence of u ϵ to the unique solution of

$$\left \{ \begin{array}{l@{\quad}l} -\partial_t u+H(x,Du) =F\bigl(x,m(t)\bigr)& \mbox{in}\ (0,T)\times \mathbb{R}^d,\\ u(x,T)=u^f(x)& \mbox{in}\ \mathbb{R}^d. \end{array} \right . $$

Note that u is Lipschitz continuous. Next we turn to the limit of m ϵ: for a fixed test function \(\varphi\in{\mathcal{C}}^{\infty}_{c}((0,T)\times \mathbb {T}^{d})\), we have

$$\int_0^T\int_{\mathbb {T}^d} \bigl(- \partial_t \varphi-\epsilon \Delta\varphi+\big \langle D_pH(x,Du_n),D \varphi(t,x)\big \rangle \bigr) m_n(t,x)=0, $$

where D p H(x,Du n ) is bounded and converges a.e. to D p H(x,Du) while m ϵ converges weakly* to m. So we get as ϵ→+∞,

$$\int_0^T\int_{\mathbb {T}^d} \bigl(- \partial_t \varphi(t,x)+\big \langle D_pH(x,Du),D\varphi(t,x)\big \rangle \bigr) m(t,x)=0, $$

which shows that m is a solution of the continuity equation (1)-(ii). In conclusion, the pair (u,m) solves (1). The uniqueness for this system is established in full details in [10], so we omit the proof. □

Lemma 5.2

There exists a constant C>0 such that

$$\|Du^\epsilon \|_\infty\leq C\quad \mbox{\textit{and}} \quad D^2u^\epsilon \leq C\quad \forall \epsilon >0. $$

Proof

The proof uses Bernstein method. We first show the uniform Lipschitz continuity in space. Let \(z=\frac{1}{2}|Du^{\epsilon }|^{2}\). From classical computations, we have

$$\partial_t z +\frac{1}{2}\Delta z + \frac{1}{2}\big \langle D_x H \bigl(x,Du^\epsilon \bigr), Du^\epsilon \big \rangle +\frac{1}{2}\big \langle D_p H\bigl(x,Du^\epsilon \bigr), Dz\big \rangle \leq\frac{1}{2} \big \langle D_xF \bigl(x,m^\epsilon (t)\bigr), Du^\epsilon \big \rangle . $$

Using the last part of assumption (7) and the smoothing assumption (5) on F, this implies that

$$\partial_t z +\frac{1}{2}\Delta z -\frac{1}{2}C_0(1+2z) + \frac{1}{2}\big \langle D_p H\bigl(x,Du^\epsilon \bigr), Dz\big \rangle \leq\frac{1}{2} \sup_m \big\|D_xF(\cdot,m)\big\|_\infty z^{\frac{1}{2}}. $$

As \(z(T,x)= \frac{1}{2}|Du^{f}(x)|^{2}\), we obtain by comparison a uniform bound on z, i.e., on Du ϵ.

Next we prove the semi-concavity. Let us fix a direction \(v\in \mathbb{R}^{d}\) with |v|=1 and compute the derivative of equation (19)-(i) twice with respect to v:

$$-\partial_t u^\epsilon _{vv} -\epsilon \Delta u^\epsilon _{vv} + \bigl[H\bigl(\cdot,Du^\epsilon \bigr) \bigr]_{vv} =F_{vv}\bigl(x,m^\epsilon (t)\bigr), $$

where \([H(\cdot,Du^{\epsilon }) ]_{v} = H_{v}+\sum_{j=1}^{d} H_{p_{j}} u^{\epsilon }_{vx_{j}}\), so that

$$\begin{aligned} \bigl[H\bigl(\cdot,Du^\epsilon \bigr) \bigr]_{vv} = & H_{vv}+ \sum_{j=1}^d \bigl(2 H_{vp_j}u^\epsilon _{vx_j}+ H_{vp_j}u^\epsilon _{vvx_j} \bigr)+ \sum_{j,k=1}^d H_{p_kp_j}u^\epsilon _{vx_k}u_{vx_j} \\ \geq& -C\bigl(1+\big|D^2u^\epsilon \big|\bigr) +\big \langle D_pH_{v}, D u^\epsilon _{vv}\big \rangle +\frac {1}{\bar{C}} \big|D^2u^\epsilon \big|^2. \end{aligned}$$

In the last inequality, we have used the fact that H is of class \({\mathcal{C}}^{2}\) and that Du ϵ is uniformly bounded, and the strict convexity of H given in assumption (7). As F vv (⋅,m ϵ) is bounded by assumption on F, and \(-C(1+|D^{2}u^{\epsilon }|) +\frac{1}{\bar{C}} |D^{2}u^{\epsilon }|^{2}\) is bounded below by a constant independent of ϵ, we obtain

$$-\partial_t u^\epsilon _{vv} -\epsilon \Delta u^\epsilon _{vv} +\big \langle D_pH_{v}, D u^\epsilon _{vv}\big \rangle \leq C. $$

Since \(u^{\epsilon }_{vv}(x,T)=u^{f}_{vv}(x)\) is bounded, we obtain by comparison a bound from above on \(u^{\epsilon }_{vv}\) independent of ϵ and for any direction v. □

We complete the paper by a standard estimate on the continuity equation:

$$ \partial_t m +\operatorname{div}(m b)=0\quad \mbox{in}\ (0,T) \times \mathbb {T}^d. $$
(21)

Lemma 5.3

Assume that \(b:(0,T)\times \mathbb {T}^{d}\to \mathbb{R}^{d}\) is a Borel vector field withb<+∞. If m satisfies (21), then m is Lipschitz continuous as a map from [0,T] to \(P(\mathbb {T}^{d})\), with a Lipschitz constant bounded above byb.

Proof

Fix 0<t 1<t 2<T and let \(h\in{\mathcal{C}}^{\infty}_{c}(\mathbb {T}^{d})\) be 1-Lipschitz continuous. Let ϵ>0 small and

$$\varphi_\epsilon (t,x)=\left \{ \begin{array}{l@{\quad}l} (t-t_1)h(x)/\epsilon & \mbox{if}\ t\in[t_1,t_1+\epsilon ],\\ h(x) & \mbox{if}\ t\in[t_1+\epsilon ,t_2-\epsilon ],\\ (t_2-t)h(x)/\epsilon & \mbox{if}\ t\in[t_2-\epsilon ,t_2],\\ 0 & \mbox{otherwise}. \end{array} \right . $$

As

$$\int_0^T \int_{\mathbb{R}^d} \bigl(- \partial_t\varphi_\epsilon +\langle m, D\varphi_\epsilon \rangle \bigr)m = 0, $$

we have

$$\int_{t_1}^{t_1+\epsilon } \int_{\mathbb{R}^d} \frac{hm }{\epsilon } +\int_{t_1+\epsilon }^{t_2-\epsilon } \int _{\mathbb{R}^d} \langle b, Dh\rangle m + \int_{t_2-\epsilon }^{t_2} \int_{\mathbb{R}^d} -\frac{hm}{\epsilon } = o(1). $$

Letting ϵ→0 gives for a.e. 0<t 1<t 2<T:

$$\int_{\mathbb{R}^d} h\, d\bigl(m(t_1)-m(t_2) \bigr) +\int_{t_1}^{t_2} \int_{\mathbb{R}^d} \langle b, Dh\rangle m=0 . $$

So

$$\int_{\mathbb{R}^d} h\, d\bigl(m(t_1)-m(t_2) \bigr) \leq\|b\|_\infty\| Dh\|_\infty |t_2-t_1| . $$

Taking the sup over h gives then

$$\mathbf{d}_1\bigl(m(t_1),m(t_2)\bigr) \leq \|b \|_\infty|t_2-t_1| . $$

 □

Rights and permissions

Reprints and permissions

About this article

Cite this article

Cardaliaguet, P. Long Time Average of First Order Mean Field Games and Weak KAM Theory. Dyn Games Appl 3, 473–488 (2013). https://doi.org/10.1007/s13235-013-0091-x

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s13235-013-0091-x

Keywords

Navigation