This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Second order local minimal-time Mean Field Games

Romain Ducasse Université de Paris and Sorbonne Université, CNRS, Laboratoire Jacques-Louis Lions (LJLL), F-75006 Paris, France. [email protected] Guilherme Mazanti Université Paris-Saclay, CNRS, CentraleSupélec, Inria, Laboratoire des signaux et systèmes, 91190, Gif-sur-Yvette, France. [email protected]  and  Filippo Santambrogio Institut Camille Jordan, Université Claude Bernard - Lyon 1; 43 boulevard du 11 novembre 1918, 69622 Villeurbanne cedex, France & Institut Universitaire de France. [email protected]
Abstract.

The paper considers a forward-backward system of parabolic PDEs arising in a Mean Field Game (MFG) model where every agent controls the drift of a trajectory subject to Brownian diffusion, trying to escape a given bounded domain Ω\Omega in minimal expected time. Agents are constrained by a bound on the drift depending on the density of other agents at their location. Existence for a finite time horizon TT is proven via a fixed point argument, but the natural setting for this problem is in infinite time horizon. Estimates are needed to treat the limit TT\to\infty, and the asymptotic behavior of the solution obtained in this way is also studied. This passes through classical parabolic arguments and specific computations for MFGs. Both the Fokker–Planck equation on the density of agents and the Hamilton–Jacobi–Bellman equation on the value function display Dirichlet boundary conditions as a consequence of the fact that agents stop as soon as they reach Ω\partial\Omega. The initial datum for the density is given, and the long-time limit of the value function is characterized as the solution of a stationary problem.

Key words and phrases:
Mean Field Games, congestion games, parabolic PDEs, MFG system, existence of solutions, asymptotic behavior
2020 Mathematics Subject Classification:
35Q89, 35K40, 35B40, 35A01, 35D30
Corresponding author

1. Introduction

Introduced around 2006 by Jean-Michel Lasry and Pierre-Louis Lions [21, 22, 23] and at the same time by Peter Caines, Minyi Huang, and Roland Malhamé [14, 15, 16], the theory of Mean Field Games (MFGs, for short) describes the interaction of a continuum of players, assumed to be rational, indistinguishable, and negligible, when each one tries to solve a dynamical control problem influenced only by the average behavior of the other players (through a mean-field type interaction, using the physicists’ terminology). The Nash equilibrium in these continuous games is described by a system of PDEs: a Hamilton–Jacobi–Bellman equation for the value function of the control problem of each player, where the distribution (density) of the players appears, coupled with a continuity equation describing the evolution of such a density, where the velocity field is the optimal one in order to solve the control problem, and is therefore related to the gradient of the value function. This system is typically forward-backward in nature: the density evolves forward in time starting from a given initial datum, and the value function backward in time, according to Bellman’s dynamical programming principle, and its final value at a given time horizon TT is usually known.

The literature about MFG theory is quickly growing and many references are available. The 6-year course given by P.-L. Lions at Collège de France, for which video-recording is available in French [26], explains well the birth of the theory, but the reader can also refer to the lecture notes by P. Cardaliaguet [6], based on the same course.

In most of the MFG models studied so far the agents consider a fixed time interval [0,T][0,T] and optimize a trajectory x:[0,T]Ωx:[0,T]\to\Omega (where Ωd\Omega\subset\mathbb{R}^{d} is the state space) trying to minimize a cost of the form 0TL(t,x(t),x(t),ρt)dt+Ψ(x(T),ρT),\int_{0}^{T}L(t,x(t),x^{\prime}(t),\rho_{t})\operatorname{d\!}t+\Psi(x(T),\rho_{T}), where ρt\rho_{t} denotes the distribution of players at time tt. The function LL is typically increasing in |x|\lvert x^{\prime}\rvert and, in some sense, in ρ\rho. This means that high velocities are costly, and passing through areas where the population is strongly concentrated is also costly. Some MFGs, called MFGs of congestion (see, for instance, [1]), consider costs which include a product of the form ρt(x(t))α|x(t)|β\rho_{t}(x(t))^{\alpha}\lvert x^{\prime}(t)\rvert^{\beta} (for some exponents α,β>0\alpha,\beta>0), which means that high velocities are costly, and that they are even more costly in the presence of high concentrations. These models present harder mathematical difficulties compared to those where the cost is decomposed into L(t,x(t),x(t))+g(t,x(t),ρt)L(t,x(t),x^{\prime}(t))+g(t,x(t),\rho_{t}). Indeed, in many cases the latter MFG admits a variational formulation: equilibria can be found by minimizing a global energy among all possible evolutions (ρt)t(\rho_{t})_{t} (hence, they are potential games). This allows to prove the existence of the equilibrium via semicontinuity methods, and we refer to [4] and [30] for a detailed discussion of this branch of MFG theory.

When the MFG has no variational interpretation, then the existence of a solution is usually obtained via fixed-point theorems, but these theorems require much more regularity. Roughly speaking, given an evolution ρ\rho one computes the corresponding value function φ\varphi as a solution to a Hamilton–Jacobi–Bellman equation and, given φ\varphi, one computes a new density evolution ρ~\tilde{\rho} by following an evolution equation. We need existence, uniqueness, and stability results for these equations in order to find a fixed point ρ~=ρ\tilde{\rho}=\rho. This usually requires regularity of the velocity field φ-\nabla\varphi, which is difficult to prove, and can be essentially only obtained in two different frameworks: either the dependence of the cost functions on the distribution ρ\rho is highly regularizing (which usually means that it is non-local, and passes through averaged quantities such as convolutions η(xy)dρ(y)\int\eta(x-y)\operatorname{d\!}\rho(y)), or diffusion of the agents is taken into account, transforming the optimal control problem into a stochastic one. In this latter case, agents minimize 𝔼[0TL(t,Xt,α(t),ρt)dt+Ψ(XT,ρT)]\mathbb{E}[\int_{0}^{T}L(t,X_{t},\alpha(t),\rho_{t})\operatorname{d\!}t+\Psi(X_{T},\rho_{T})] where the process XX follows dXt=αtdt+dBt\operatorname{d\!}X_{t}=\alpha_{t}\operatorname{d\!}t+\operatorname{d\!}B_{t} and (Bt)t0(B_{t})_{t\geq 0} represents a standard Brownian motion.

In [27], the second and third authors of the present paper introduced a different class of models, called minimal-time MFGs. The main difference is that instead of considering a cost for the players penalizing both the velocity and the density, and minimizing the integral of such a cost on a fixed time interval [0,T][0,T], the dynamics is subject to a constraint where the maximal velocity of the agents cannot exceed a quantity depending on the density ρt\rho_{t}, and the goal of each agent is to arrive to a given target as soon as possible. In the typical situation, the target of the agents is the boundary Ω\partial\Omega of the domain where the evolution occurs. This can model, for instance, an evacuation phenomenon in crowd motion. The system that one obtains is the following

(1.1) {tρ(ρk[ρ]φ|φ|)=0, in (0,T)×Ω,tφ+k[ρ]|φ|1=0, in (0,T)×Ω,ρ(0,x)=ρ0(x), in Ω,φ(t,x)=0, on (0,T)×Ω,\left\{\begin{aligned} &\partial_{t}\rho-\nabla\cdot\left(\rho k[\rho]\frac{\nabla\varphi}{\lvert\nabla\varphi\rvert}\right)=0,&\quad&\text{ in }(0,T)\times\Omega,\\ &-\partial_{t}\varphi+k[\rho]\lvert\nabla\varphi\rvert-1=0,&&\text{ in }(0,T)\times\Omega,\\ &\rho(0,x)=\rho_{0}(x),&&\text{ in }\Omega,\\ &\varphi(t,x)=0,&&\text{ on }(0,T)\times\partial\Omega,\end{aligned}\right.

where the function k[ρt](x)k[\rho_{t}](x) denotes the maximal speed that agents can have at point xx at time tt, i.e., the dynamics is constrained to satisfy |x(t)|k[ρt](x(t))\lvert x^{\prime}(t)\rvert\leq k[\rho_{t}](x(t)). Ideally, one would like to choose kk to be a non-increasing function of the density itself, such as k[ρ](x)=(1ρ(x))+k[\rho](x)=(1-\rho(x))_{+}. This choice is what is done in the well-known Hughes’ model for crowd motion [17, 18]. Indeed, this model is very similar to Hughes’, which also considers agents who aim at leaving in minimal time a bounded domain under a congestion-dependent constraint on their speeds.

The main difference between the model in [27] (from which the present paper stems) and Hughes’ is that, in the latter, at each time, an agent moves in the optimal direction to the boundary assuming that the distribution of agents remains constant, whereas in [27] and here agents take into account the future evolution of the distribution of agents in the computation of their optimal trajectories. This accounts for the time derivative in the Hamilton–Jacobi–Bellman equation from (1.1), which is the main difference between (1.1) and the equations describing the motion of agents in Hughes’ model and stands for the anticipation of future behavior of other agents.

Another crucial (and disappointing) similarity between the above MFG system and Hughes’ model is the fact that general mathematical results do not exist in the case k[ρ]=(1ρ)+k[\rho]=(1-\rho)_{+} and more generally in the local case (except few results in the Hughes case in 1D). Indeed, the lack of regularity makes the model too hard to study, and the MFG case is not variational.

In some sense the closest MFG model to this one is the one with multiplicative costs in [1] (MFG with congestion). Indeed, an LL^{\infty} constraint |x|k[ρ]\lvert x^{\prime}\rvert\leq k[\rho] can be seen as a limit as mm\to\infty of an integral penalization

||x(t)|k[ρt](x(t))|mdt.\int\left\lvert\frac{\lvert x^{\prime}(t)\rvert}{k[\rho_{t}](x(t))}\right\rvert^{m}\operatorname{d\!}t.

Note that the boundaries of the time interval have been omitted on purpose from the above integral, since the model in [1] is set on a fixed time horizon but this is not part of our setting. For MFG with congestion, [1] presents not only existence but also uniqueness results, under the assumption that the exponents appearing in the running cost satisfy a certain inequality. Unfortunately this inequality is never satisfied in the limit m=+m=+\infty as above, and it is not surprising that in our work we are not able to establish uniqueness results for our MFG system. Additional results for MFG with congestion were presented, for instance, in [12], under a smallness assumption on the time horizon, but this assumption cannot be made here, as the model is exactly meant to consider the case where a time horizon is not fixed.

Because of these difficulties, [27] studied the case of a non-local dependence of kk w.r.t. ρ\rho (say, k[ρ](x)=κ(η(xy)dρ(y))k[\rho](x)=\kappa(\int\eta(x-y)\operatorname{d\!}\rho(y)), for a non-increasing function κ\kappa and a positive convolution kernel η\eta), and proved existence of an equilibrium, characterized it as a solution of a non-local MFG system, and analyzed some examples, including numerical simulations. Instead, in the present paper we want to study the local case with diffusion.

This means that we will consider a local dependence k[ρ](x):=κ(ρ(x))k[\rho](x):=\kappa(\rho(x)), and each agent solves a stochastic control problem

inf{𝔼[τ]:X(τ)Ω,X(0)=x0,dXt=αtdt+2νdBt,|αt|κ(ρ(t,Xt))},\inf\left\{\mathbb{E}[\tau]\,:\,X(\tau)\in\partial\Omega,X(0)=x_{0},\,\operatorname{d\!}X_{t}=\alpha_{t}\operatorname{d\!}t+\sqrt{2\nu}\operatorname{d\!}B_{t},\,\lvert\alpha_{t}\rvert\leq\kappa(\rho(t,X_{t}))\right\},

where (Bt)t0(B_{t})_{t\geq 0} denotes a standard Brownian motion and the Brownian motions for all players are assumed to be mutually independent. Defining the corresponding value function φ\varphi, from classical results on stochastic optimal control (see [10, Chapter IV]), under suitable assumptions, the optimal control is given in feedback form by

αt=κ(ρ(t,Xt))φ(t,Xt)|φ(t,Xt)|,\alpha_{t}=-\kappa(\rho(t,X_{t}))\frac{\nabla\varphi(t,X_{t})}{\left\lvert\nabla\varphi(t,X_{t})\right\rvert},

(a definition which has to be carefully adapted to the case φ=0\nabla\varphi=0); moreover, the value function solves the Hamilton–Jacobi–Bellman equation

tφ(t,x)νΔφ(t,x)+K(t,x)|φ(t,x)|1=0,(t,x)[0,T)×Ω,-\partial_{t}\varphi(t,x)-\nu\Delta\varphi(t,x)+K(t,x)\left\lvert\nabla\varphi(t,x)\right\rvert-1=0,\quad(t,x)\in[0,T)\times\Omega,

for K=κ(ρ)K=\kappa(\rho). Hence, we know the drift of the optimal stochastic processes followed by each agent, and this allows to write the Fokker–Planck equation solved by the law of this process. Putting together all this information, we obtain the following MFG system

(1.2) {tρνΔρ(ρκ(ρ)φ|φ|)=0, in +×Ω,tφνΔφ+κ(ρ)|φ|1=0, in +×Ω,ρ(0,x)=ρ0(x),ρ(t,x)=0,φ(t,x)=0, in Ω, on +×Ω,\left\{\begin{aligned} &\partial_{t}\rho-\nu\Delta\rho-\nabla\cdot\left(\rho\kappa(\rho)\frac{\nabla\varphi}{\lvert\nabla\varphi\rvert}\right)=0,&\quad&\text{ in }\mathbb{R}_{+}\times\Omega,\\ &-\partial_{t}\varphi-\nu\Delta\varphi+\kappa(\rho)\lvert\nabla\varphi\rvert-1=0,&&\text{ in }\mathbb{R}_{+}\times\Omega,\\ &\begin{aligned} \rho(0,x)&=\rho_{0}(x),\\ \rho(t,x)&=0,\quad\varphi(t,x)=0,\end{aligned}&&\begin{aligned} &\text{ in }\Omega,\\ &\text{ on }\mathbb{R}_{+}\times\partial\Omega,\end{aligned}\end{aligned}\right.

where Ωd\Omega\subset\mathbb{R}^{d} is an open and bounded set, whose boundary will be supposed to be of class C2C^{2} in this paper, ν>0\nu>0 is a fixed constant, κ:(0,+)\kappa:\mathbb{R}\to(0,+\infty), and ρ00\rho_{0}\geq 0 is the initial density. The Dirichlet condition on φ\varphi comes as usual from the fact that, for agents who are already on the boundary, the remaining time to reach it is zero, and the Dirichlet condition on ρ\rho comes from the fact that we stop the evolution of a particle as soon as it touches the boundary (absorbing boundary conditions).

A crucial difference with the previous paper [27] concerns the time horizon. If we suppose that κ\kappa is bounded from below in the model without diffusion, it is not difficult to see that all agents will have left the domain after a common finite time, so that the final value of φ\varphi is not really relevant, and the problem can be studied on a finite interval [0,T][0,T]. This is not the case when there is diffusion, as a density following a Fokker–Planck equation with a bounded drift cannot fully vanish in finite time. As a consequence, the model should be studied on the unbounded interval [0,)[0,\infty). For every time t<t<\infty there is still mass everywhere, but this mass decreases to 0 as t+t\to+\infty, which suggests that the value function φ\varphi should converge to a function, that we call Ψ\Psi, which is the value function for the corresponding control problem with no mass, i.e. when κ=κ(0)\kappa=\kappa(0). Since in this control problem κ\kappa is independent of time, Ψ\Psi is a function of xx only and solves a stationary Hamilton–Jacobi–Bellman equation which takes the form of an elliptic PDE

νΔΨ+κ(0)|Ψ|1=0-\nu\Delta\Psi+\kappa(0)\lvert\nabla\Psi\rvert-1=0

with Dirichlet boundary conditions on Ω\partial\Omega. It is then reasonable to investigate whether solutions of the above system satisfy further ρt0\rho_{t}\to 0 and φtΨ\varphi_{t}\to\Psi as t+t\to+\infty.

In order to study the above system, we will first study an artificial finite-horizon setting, where we stop the game at time TT, choose a penalization ψ:Ω+\psi:\Omega\to\mathbb{R}_{+} with ψ=0\psi=0 on Ω\partial\Omega, and look at the stochastic optimal control problem

inf{𝔼[min{τ,T}+ψ(Xmin{τ,T})]:X(τ)Ω,X(0)=x0,dXt=αtdt+2νdBt,|αt|κ(ρ(t,Xt))}.\inf\Big{\{}\mathbb{E}[\min\{\tau,T\}+\psi(X_{\min\{\tau,T\}})]\,:\\ \,X(\tau)\in\partial\Omega,X(0)=x_{0},\,\operatorname{d\!}X_{t}=\alpha_{t}\operatorname{d\!}t+\sqrt{2\nu}\operatorname{d\!}B_{t},\,\lvert\alpha_{t}\rvert\leq\kappa(\rho(t,X_{t}))\Big{\}}.

This gives rise to the MFG system

(1.3) {tρνΔρ(ρκ(ρ)φ|φ|)=0, in (0,T)×Ω,tφνΔφ+κ(ρ)|φ|1=0, in (0,T)×Ω,ρ(0,x)=ρ0(x),φ(T,x)=ψ(x),ρ(t,x)=0,φ(t,x)=0, in Ω, on (0,T)×Ω,\left\{\begin{aligned} &\partial_{t}\rho-\nu\Delta\rho-\nabla\cdot\left(\rho\kappa(\rho)\frac{\nabla\varphi}{\lvert\nabla\varphi\rvert}\right)=0,&\quad&\text{ in }(0,T)\times\Omega,\\ &-\partial_{t}\varphi-\nu\Delta\varphi+\kappa(\rho)\lvert\nabla\varphi\rvert-1=0,&&\text{ in }(0,T)\times\Omega,\\ &\begin{aligned} \rho(0,x)&=\rho_{0}(x),\quad&\varphi(T,x)&=\psi(x),\\ \rho(t,x)&=0,&\varphi(t,x)&=0,\end{aligned}&&\begin{aligned} &\text{ in }\Omega,\\ &\text{ on }(0,T)\times\partial\Omega,\end{aligned}\end{aligned}\right.

which corresponds to (1.2) with the unbounded time interval +\mathbb{R}_{+} replaced by (0,T)(0,T) and the additional final condition φ(T,x)=ψ(x)\varphi(T,x)=\psi(x). We will prove the existence of a solution of the system for finite TT (note that for this system, as well as for its infinite-horizon counterpart, we are not able to prove uniqueness), and then consider the limit as TT\to\infty. In order to guarantee suitable bounds, we just need to choose a sequence of final data ψT\psi_{T}, possibly depending on TT, which is uniformly bounded. We will then get at the limit a solution of the limit system which automatically satisfies ρt0\rho_{t}\to 0 (in the sense of uniform convergence) and φtΨ\varphi_{t}\to\Psi (this convergence being both uniform and strong in H01H^{1}_{0}).

The paper is organized as follows. After this introduction, Section 2 presents the tools that we need to study the two separate equations appearing in System (1.3) on a finite horizon, which come from the classical theory of parabolic equations. Section 3 is devoted to the existence of solutions of (1.3). After providing a precise definition of solution of (1.3) taking care of the case φ=0\nabla\varphi=0, we use the estimates of Section 2 to prove existence via a fixed-point argument based on Kakutani’s theorem. Section 4 concerns the limit TT\to\infty. In this section, some estimates of Section 2 need to be made more precise, in order to see how constants depend on the time horizon TT. In this way we are able to prove existence of a limit of the solutions of (1.3) as the time horizon TT tends to ++\infty and that this limit solves the limit system (1.2). Then we consider the asymptotic behavior of a solution (ρ,φ)(\rho,\varphi) of (1.2) as t+t\to+\infty, proving first ρt0\rho_{t}\to 0 in L1L^{1} and, thanks to a parabolic regularization argument, also in LL^{\infty}. To prove convergence in L1L^{1}, which is true for general Fokker–Planck systems under very mild assumptions, we exploit the MFG nature of the system, i.e. the coupling between the two equations, which also provides exponential decrease. We then consider the limit in time of φ\varphi, and prove that any bounded solution of this equation, once we know κ(t,x)κ(0)\kappa(t,x)\to\kappa(0), can only converge as t+t\to+\infty to the stationary function Ψ\Psi. This convergence is a priori very weak, but we are able to improve it into LH01L^{\infty}\cap H^{1}_{0}, and to prove that the uniform convergences of both ρ\rho and φ\varphi occur exponentially fast. The paper is then completed by an appendix, which details some global LL^{\infty} estimates for a large class of parabolic equations, including the estimates that we use to prove uniform convergence in time of ρt\rho_{t} and φt\varphi_{t} to 0 and Ψ,\Psi, respectively. These estimates are not surprising and not difficult to prove, using standard Moser iterations, but are not easy to find in the literature under the sole assumption of boundedness of the drift term in the divergence. The computations and the results are essentially the same as in the appendix of [7], but the boundary conditions are different.

2. Preliminary results

This section presents some preliminary results on Fokker–Planck and Hamilton–Jacobi–Bellman equations which are useful for the analysis of the Mean Field Game systems (1.2) and (1.3). We recall that, in the whole paper, Ω\Omega denotes an open and bounded set whose boundary Ω\partial\Omega is assumed to be C2C^{2}. Even though some of the results presented in this preliminary section also hold without the smoothness assumption on Ω\partial\Omega (such as existence and uniqueness results for both Fokker–Planck and Hamilton–Jacobi–Bellman equations in Propositions 2.2 and 2.5), this assumption is first used to obtain higher regularity of solutions of Hamilton–Jacobi–Bellman equations in Proposition 2.5 and is required for almost all of the subsequent results, including in particular our main results in Sections 3 and 4, as a consequence of the need of higher regularity of φ\varphi.

2.1. Fokker–Planck equation

We recall some results on the Fokker–Planck equation on a bounded domain Ωd\Omega\subset\mathbb{R}^{d} in finite time horizon T(0,+)T\in(0,+\infty),

(2.1) {tρνΔρ+(ρV)=0 in (0,T)×Ω,ρ(0,x)=ρ0(x) in Ω,ρ(t,x)=0 in [0,T]×Ω,\left\{\begin{aligned} &\partial_{t}\rho-\nu\Delta\rho+\nabla\cdot(\rho V)=0&\quad&\text{ in }(0,T)\times\Omega,\\ &\rho(0,x)=\rho_{0}(x)&&\text{ in }\Omega,\\ &\rho(t,x)=0&&\text{ in }[0,T]\times\partial\Omega,\end{aligned}\right.

where V:(0,T)×ΩdV:(0,T)\times\Omega\to\mathbb{R}^{d} is a given velocity field. We will only focus on the case where VV is bounded, an assumption which is satisfied in the cases of interest for this paper and which strongly simplifies the analysis. The results presented in this short section are a mixture of classical results (for which we mainly refer to [8, Section 7.1] or [19, Chapter III]), recent results obtained by Porretta in [28], and extra computations which are not original but are difficult to find in the literature, which we present in the Appendix.

Definition 2.1.

Let ν>0\nu>0, VL((0,T)×Ω;d)V\in L^{\infty}((0,T)\times\Omega;\mathbb{R}^{d}), and ρ0L1(Ω)\rho_{0}\in L^{1}(\Omega). We say that ρL1((0,T)×Ω)\rho\in L^{1}((0,T)\times\Omega) is a weak solution of (2.1) if, for every ηC2([0,T]×Ω)\eta\in C^{2}([0,T]\times\Omega) such that η|[0,T]×Ω=0\eta\bigr{\rvert}_{[0,T]\times\partial\Omega}=0 and η|{T}×Ω=0\eta\bigr{\rvert}_{\{T\}\times\Omega}=0, one has

(2.2) 0TΩρtηdxdt0TΩ(νρΔη+ρVη)dxdt=Ωρ0(x)η(0,x)dx.-\int_{0}^{T}\int_{\Omega}\rho\partial_{t}\eta\operatorname{d\!}x\operatorname{d\!}t-\int_{0}^{T}\int_{\Omega}\left(\nu\rho\Delta\eta+\rho V\cdot\nabla\eta\right)\operatorname{d\!}x\operatorname{d\!}t=\int_{\Omega}\rho_{0}(x)\eta(0,x)\operatorname{d\!}x.

We observe that, whenever equality (2.2) holds for C2C^{2} functions, and if we have further that ρL2((t1,t2);H01(Ω))C0([t1,t2];L2(Ω))\rho\in L^{2}((t_{1},t_{2});H^{1}_{0}(\Omega))\cap C^{0}([t_{1},t_{2}];L^{2}(\Omega)) for some t1,t2(0,T)t_{1},t_{2}\in(0,T) with t1<t2t_{1}<t_{2}, then we also have

(2.3) t1t2(Ωρtη+νρηρVη)dxdt=Ωρ(t1,x)η(t1,x)dxΩρ(t2,x)η(t2,x)dx.\int_{t_{1}}^{t_{2}}\left(\int_{\Omega}-\rho\partial_{t}\eta+\nu\nabla\rho\cdot\nabla\eta-\rho V\cdot\nabla\eta\right)\operatorname{d\!}x\operatorname{d\!}t\\ =\int_{\Omega}\rho(t_{1},x)\eta(t_{1},x)\operatorname{d\!}x-\int_{\Omega}\rho(t_{2},x)\eta(t_{2},x)\operatorname{d\!}x.

for every ηCc1([0,T)×Ω))\eta\in C^{1}_{c}([0,T)\times\Omega)) and, by density, for every ηL2((t1,t2);H01(Ω))C0([t1,t2];L2(Ω))\eta\in L^{2}((t_{1},t_{2});H^{1}_{0}(\Omega))\cap C^{0}([t_{1},t_{2}];\allowbreak L^{2}(\Omega)) such that tηL2((t1,t2);H1(Ω))\partial_{t}\eta\in L^{2}((t_{1},t_{2});H^{-1}(\Omega)). Of course it is well-known that, in case ρ\rho is more regular, other test functions can also be accepted, and that if ρC2\rho\in C^{2} then the equation is satisfied in a classical sense.

We now state a proposition summarizing all the main results that we will use.

Proposition 2.2.

Let ν>0\nu>0, VL((0,T)×Ω;d)V\in L^{\infty}((0,T)\times\Omega;\mathbb{R}^{d}), and ρ0L1(Ω)\rho_{0}\in L^{1}(\Omega) be a given non-negative initial datum. Then (2.1) admits a unique weak solution ρ\rho. In addition, we have ρ0\rho\geq 0 and ρC0([0,T];L1(Ω))\rho\in C^{0}([0,T];L^{1}(\Omega)) with ρtL1ρ0L1\lVert\rho_{t}\rVert_{L^{1}}\leq\lVert\rho_{0}\rVert_{L^{1}}, as well as ρLq((0,T)×Ω)\nabla\rho\in L^{q}((0,T)\times\Omega) and tρLq((0,T);W1,q(Ω))\partial_{t}\rho\in L^{q}((0,T);W^{-1,q}(\Omega)) for all q<d+2d+1q<\frac{d+2}{d+1} and ρLr((0,T)×Ω)\rho\in L^{r}((0,T)\times\Omega) for all r<d+2dr<\frac{d+2}{d}, and the norms of ρ,ρ\rho,\nabla\rho and tρ\partial_{t}\rho in the above spaces are bounded by quantities only depending on ρ0L1\lVert\rho_{0}\rVert_{L^{1}}. Moreover, for every t0>0t_{0}>0, we also have ρL((t0,T)×Ω)L2((t0,T);H01(Ω))C0([t0,T];L2(Ω))\rho\in L^{\infty}((t_{0},T)\times\Omega)\cap L^{2}((t_{0},T);H^{1}_{0}(\Omega))\cap C^{0}([t_{0},T];L^{2}(\Omega)) and tρL2((t0,T);H1(Ω))\partial_{t}\rho\in L^{2}((t_{0},T);H^{-1}(\Omega)).

Of course we do not provide a full proof of the above results, but we explain below how to deduce the different parts of the statement from the most well-known literature and the relevant references.

Proof.

The definition of the solution is exactly the one used in [28], where the key assumption is ρ|V|2L1((0,T)×Ω)\rho|V|^{2}\in L^{1}((0,T)\times\Omega). In our case, where VV is bounded, this assumption is satisfied as soon as ρL1((0,T)×Ω)\rho\in L^{1}((0,T)\times\Omega). One of the main results of [28] is exactly the uniqueness of the solution in this class, and this can be applied to the present setting. The same paper also guarantees the estimates ρLq((0,T)×Ω)\nabla\rho\in L^{q}((0,T)\times\Omega), tρLq((0,T);W1,q(Ω))\partial_{t}\rho\in L^{q}((0,T);W^{-1,q}(\Omega)), ρLr((0,T)×Ω)\rho\in L^{r}((0,T)\times\Omega), and the L1L^{1} bound.

Existence is not included in [28] but in the particular case VLV\in L^{\infty} it is easy to obtain by regularization and compactness. Indeed, one can apply the classical L2L^{2} theory of [19, Chapter III] to an approximated initial datum, and obtain a sequence of solutions: the LrL^{r} bounds of [28], which only depend on the initial L1L^{1} norm in this setting, allow to obtain the compactness we need to pass the PDE to the limit. Note that this argument is specific to the case VLV\in L^{\infty} since, otherwise, we would need to control the L1L^{1} norm of ρ|V|2\rho|V|^{2}, which is non-trivial.

By approximating ρ0\rho_{0} and VV by smooth functions ρ0,ε\rho_{0,\varepsilon} and VεV_{\varepsilon} with ρ0,ε0\rho_{0,\varepsilon}\geq 0, the corresponding solution ρε\rho_{\varepsilon} of (2.1) satisfies ρε0\rho_{\varepsilon}\geq 0 thanks to the classical maximum principle, and then this property passes to the limit and also applies to the unique weak solution ρ\rho corresponding to the original ρ0\rho_{0} and VV.

The local LL^{\infty} bound can be obtained thanks to the Appendix of the present paper (even if we stress that similar computations are nowadays standard). For simplicity, the bound is presented under the assumption ρ0Lr\rho_{0}\in L^{r}, r>1r>1, and not ρ0L1\rho_{0}\in L^{1}. Yet, the time-space LrL^{r} summability already stated in the claim allows to deduce ρtLr\rho_{t}\in L^{r} for a.e. t>0t>0, and if we choose t<t0t<t_{0} we obtain the desired LL^{\infty} bound. Once we know that ρ\rho is locally (in time) LL^{\infty} (in space), it is also locally (in time) L2L^{2} (in space), and hence the classical L2L^{2} theory of [19, Chapter III] provides the last estimates of the statement. ∎

2.2. Hamilton–Jacobi–Bellman equation

We consider the non-linear Hamilton–Jacobi–Bellman equation in a finite time horizon

(2.4) {tφνΔφ+K|φ|1=0 in (0,T)×Ω,φ(T,x)=ψ(x) in Ω,φ(t,x)=0 in [0,T]×Ω,\left\{\begin{aligned} &-\partial_{t}\varphi-\nu\Delta\varphi+K\lvert\nabla\varphi\rvert-1=0&\quad&\text{ in }(0,T)\times\Omega,\\ &\varphi(T,x)=\psi(x)&&\text{ in }\Omega,\\ &\varphi(t,x)=0&&\text{ in }[0,T]\times\partial\Omega,\end{aligned}\right.

where K:(0,T)×ΩK:(0,T)\times\Omega\to\mathbb{R} is a given function.

Definition 2.3.

Let ν>0\nu>0, KL((0,T)×Ω;)K\in L^{\infty}((0,T)\times\Omega;\mathbb{R}), and ψL2(Ω)\psi\in L^{2}(\Omega). We say that φL((0,T);L2(Ω))L2((0,T);H01(Ω))\varphi\in L^{\infty}((0,T);L^{2}(\Omega))\cap L^{2}((0,T);H^{1}_{0}(\Omega)) is a weak solution of (2.4) if, for every ηC1([0,T]×Ω)\eta\in C^{1}([0,T]\times\Omega) such that η|[0,T]×Ω=0\eta\bigr{\rvert}_{[0,T]\times\partial\Omega}=0 and η|{0}×Ω=0\eta\bigr{\rvert}_{\{0\}\times\Omega}=0, one has

(2.5) 0TΩφtη+ν0TΩφη+0TΩ(K|φ|1)η=Ωψ(x)η(T,x)dx.\int_{0}^{T}\int_{\Omega}\varphi\partial_{t}\eta+\nu\int_{0}^{T}\int_{\Omega}\nabla\varphi\cdot\nabla\eta+\int_{0}^{T}\int_{\Omega}(K\left\lvert\nabla\varphi\right\rvert-1)\eta=\int_{\Omega}\psi(x)\eta(T,x)\operatorname{d\!}x.

As we did after Definition 2.1, we observe that, if (2.5) holds for every η\eta as before, and if we assume further that φC0([t1,t2];L2(Ω))\varphi\in C^{0}([t_{1},t_{2}];L^{2}(\Omega)) for some t1,t2(0,T)t_{1},t_{2}\in(0,T) with t1<t2t_{1}<t_{2}, then we also have

(2.6) t0t1Ω(φtη+νφη+(K|φ|1)η)=Ωφ(t1,x)η(t1,x)dxΩφ(t0,x)η(t0,x)dx,\int_{t_{0}}^{t_{1}}\int_{\Omega}\left(\varphi\partial_{t}\eta+\nu\nabla\varphi\cdot\nabla\eta+(K\left\lvert\nabla\varphi\right\rvert-1)\eta\right)\\ =\int_{\Omega}\varphi(t_{1},x)\eta(t_{1},x)\operatorname{d\!}x-\int_{\Omega}\varphi(t_{0},x)\eta(t_{0},x)\operatorname{d\!}x,

for every ηCc1((0,T]×Ω)\eta\in C^{1}_{c}((0,T]\times\Omega) and, by density, for every ηL2((t1,t2);H01(Ω))C0([t1,t2];L2(Ω))\eta\in L^{2}((t_{1},t_{2});H_{0}^{1}(\Omega))\cap C^{0}([t_{1},t_{2}];\allowbreak L^{2}(\Omega)) such that tηL2((t1,t2);H1(Ω))\partial_{t}\eta\in L^{2}((t_{1},t_{2});H^{-1}(\Omega)).

Remark 2.4.

Note that (2.4), as a Hamilton–Jacobi–Bellman equation of an optimal control problem, is backward in time: the final condition φ(T,x)=ψ(x)\varphi(T,x)=\psi(x) is given and one solves the equation in the time interval [0,T][0,T]. One can apply classical results on forward PDEs to (2.4) by using the standard time reversal tTtt\mapsto T-t.

The next proposition gathers the main results on solutions of (2.4) that will be needed in the paper.

Proposition 2.5.

Let ν>0\nu>0, KL((0,T)×Ω)K\in L^{\infty}((0,T)\times\Omega), and ψL2(Ω)\psi\in L^{2}(\Omega). Then (2.4) admits a unique weak solution φ\varphi. In addition, we have φC0([0,T];L2(Ω))\varphi\in C^{0}([0,T];L^{2}(\Omega)), and the norms of φ\varphi in L((0,T);L2(Ω))L^{\infty}((0,T);L^{2}(\Omega)) and L2((0,T);H01(Ω))L^{2}((0,T);H_{0}^{1}(\Omega)) are bounded by quantities depending only on dd, ν\nu, TT, Ω\Omega, an upper bound on KL((0,T)×Ω)\left\lVert K\right\rVert_{L^{\infty}((0,T)\times\Omega)}, and ψL2(Ω)\left\lVert\psi\right\rVert_{L^{2}(\Omega)}.

Moreover, if ψ0\psi\geq 0 a.e. in Ω\Omega, then the unique solution also satisfies φ0\varphi\geq 0 a.e. in (0,T)×Ω(0,T)\times\Omega. If K0K\geq 0, ψH01(Ω)L(Ω)\psi\in H_{0}^{1}(\Omega)\cap L^{\infty}(\Omega), and ψ0\psi\geq 0 a.e. in Ω\Omega, then there exists a constant C>0C>0 depending on ν\nu, Ω\Omega, and ψL\lVert\psi\rVert_{L^{\infty}} such that φC\varphi\leq C a.e. on (0,T)×Ω(0,T)\times\Omega.

Finally, if ψH01(Ω)\psi\in H_{0}^{1}(\Omega), then φC0([0,T];H01(Ω))L2((0,T);H2(Ω))\varphi\in C^{0}([0,T];H_{0}^{1}(\Omega))\cap L^{2}((0,T);H^{2}(\Omega)), tφL2((0,T)×Ω)\partial_{t}\varphi\in L^{2}((0,T)\times\Omega), and the norms of φ\varphi in these spaces are bounded by quantities depending only on dd, ν\nu, TT, Ω\Omega, an upper bound on KL((0,T)×Ω)\left\lVert K\right\rVert_{L^{\infty}((0,T)\times\Omega)}, and ψH01(Ω)\left\lVert\psi\right\rVert_{H_{0}^{1}(\Omega)}.

The results stated in Proposition 2.5 are classical and follow from more general results for nonlinear pseudo-monotone operators. Similarly to Proposition 2.2, we explain below how they can be retrieved from the relevant literature.

Proof.

Existence of a weak solution φ\varphi for ψL2(Ω)\psi\in L^{2}(\Omega) follows from [29, Theorem 2.1] and the corresponding bounds on the norms of φ\varphi are a consequence of [29, Lemma 4.1], whereas uniqueness follows from [9, Theorem 2.4].

The positivity of φ\varphi when ψ0\psi\geq 0 is classical for smooth solutions and can be obtained by an easy application of the maximum principle for parabolic equations. For solutions of HJB obtained as value functions of a stochastic control problem, the result is also straightforward, as the quantity which is minimized is positive. In our context of weak solutions, it can be deduced by applying, for instance, [2, Theorem 1] to φ-\varphi, after changing time orientation and paying attention to the observation at the end of the proof (page 98) that the inequality is enough (indeed, the source term 11 in the HJB equation has the good sign to preserve positivity).

The upper bound on φ\varphi under the positivity assumption on ψ\psi and KK and the fact that ψH01(Ω)L(Ω)\psi\in H_{0}^{1}(\Omega)\cap L^{\infty}(\Omega) can be obtained by applying a parabolic comparison principle (see [24, Theorem 9.1] for the smooth case) to φ\varphi and Φ+ψL\Phi+\lVert\psi\rVert_{L^{\infty}}, where Φ\Phi is the solution of the torsion equation νΔΦ=1-\nu\Delta\Phi=1 in Ω\Omega with Dirichlet boundary conditions.

Finally, higher regularity of φ\varphi when ψH01(Ω)\psi\in H_{0}^{1}(\Omega) can be obtained in a straightforward manner by noticing that tφνΔφ=1K|φ|-\partial_{t}\varphi-\nu\Delta\varphi=1-K\lvert\nabla\varphi\rvert, i.e., φ\varphi satisfies a linear backwards heat equation in Ω\Omega with source term 1K|φ|L2((0,T)×Ω)1-K\lvert\nabla\varphi\rvert\in L^{2}((0,T)\times\Omega). The conclusions then follow from classical improved regularity results for heat equations (such as [8, Section 7.1, Theorem 5] and [19, Chapter III, § 6, Equation (6.10) and Theorem 6.1]). ∎

We next state, for future reference, a standard parabolic comparison principle for (2.4) (see, e.g., [9, Corollary 2.2]).

Proposition 2.6.

Let φ1,φ2\varphi_{1},\varphi_{2} be two solutions of (2.4) with T<+T<+\infty, with final data such that φ1(T,)φ2(T,)\varphi_{1}(T,\cdot)\geq\varphi_{2}(T,\cdot). Then

φ1φ2 on (0,T)×Ω.\varphi_{1}\geq\varphi_{2}\quad\text{ on }\ (0,T)\times\Omega.

3. The MFG system with a finite time horizon

We now consider the MFG system with a finite time horizon (1.3). One of the difficulties in the analysis of (1.3) is that the velocity field in the continuity equation depends on φ|φ|\frac{\nabla\varphi}{\lvert\nabla\varphi\rvert}, which is defined only when φ0\nabla\varphi\neq 0. In order to handle this difficulty, we make use of the following definition of weak solution.

Definition 3.1.

Let ν>0\nu>0, T(0,+)T\in(0,+\infty), κ:(0,+)\kappa:\mathbb{R}\to(0,+\infty) be continuous and bounded, ρ0L1(Ω)\rho_{0}\in L^{1}(\Omega), and ψL2(Ω)\psi\in L^{2}(\Omega). We say that (ρ,φ)L1((0,T)×Ω)×L2((0,T);H01(Ω))(\rho,\varphi)\in L^{1}((0,T)\times\Omega)\times L^{2}((0,T);H_{0}^{1}(\Omega)) is a weak solution of (1.3) with initial condition ρ0\rho_{0} and final condition ψ\psi if there exists VL((0,T)×Ω;d)V\in L^{\infty}((0,T)\times\Omega;\mathbb{R}^{d}) such that |V(t,x)|κ(ρ(t,x))\lvert V(t,x)\rvert\leq\kappa(\rho(t,x)) and V(t,x)φ(t,x)=κ(ρ(t,x))|φ(t,x)|V(t,x)\cdot\nabla\varphi(t,x)=-\kappa(\rho(t,x))\left\lvert\nabla\varphi(t,x)\right\rvert a.e. on (0,T)×Ω(0,T)\times\Omega and such that ρ\rho is a solution of the Fokker–Planck equation (2.1) with initial datum ρ0\rho_{0} and vector field VV on [0,T]×Ω[0,T]\times\Omega in the sense of Definition 2.1, and φ\varphi is a solution of the Hamilton–Jacobi–Bellman equation (2.4) with final datum ψ\psi and K=κ(ρ)K=\kappa(\rho) in the sense of Definition 2.3 on the same domain.

Remark 3.2.

If (ρ,φ)(\rho,\varphi) is a weak solution of (1.3) and VV is any function satisfying the properties stated in Definition 3.1, then we have V(t,x)=κ(ρ(t,x))φ(t,x)|φ(t,x)|V(t,x)=-\kappa(\rho(t,x))\frac{\nabla\varphi(t,x)}{\lvert\nabla\varphi(t,x)\rvert} wherever φ(t,x)0\nabla\varphi(t,x)\neq 0. The introduction of the function VV in Definition 3.1 has the advantages of providing a meaning to the first equation of (1.3) and handling its velocity field even when φ(t,x)=0\nabla\varphi(t,x)=0, which might a priori happen in a set of positive measure.

The main result of this section is the following.

Theorem 3.3.

Let ν>0\nu>0, T(0,+)T\in(0,+\infty), κ:(0,+)\kappa:\mathbb{R}\to(0,+\infty) be continuous and bounded, ρ0L1(Ω)\rho_{0}\in L^{1}(\Omega), and ψH01(Ω)\psi\in H_{0}^{1}(\Omega). Then there exists a weak solution (ρ,φ)(\rho,\varphi) of (1.3) with initial condition ρ0\rho_{0} and final condition ψ\psi.

The proof of Theorem 3.3 relies on a fixed-point argument on the velocity field VV of the Fokker–Planck equation in (1.3). Before turning to the proof, we need some continuity results on solutions of (2.1) with respect to the velocity field VV and on solutions of (2.4) with respect to the function KK, which we state and prove now.

Proposition 3.4.

Let ν>0\nu>0 and ρ0L1(Ω)\rho_{0}\in L^{1}(\Omega). Given VL((0,T)×Ω;d)V\in L^{\infty}((0,T)\times\Omega;\mathbb{R}^{d}), let (Vn)n(V_{n})_{n\in\mathbb{N}} be a sequence in L((0,T)×Ω;d)L^{\infty}((0,T)\times\Omega;\mathbb{R}^{d}) such that VnVV_{n}\xrightharpoonup{\ast}V as nn\to\infty. For nn\in\mathbb{N}, let ρn\rho_{n} (resp. ρ\rho) be the unique weak solution of (2.1) in L1((0,T)×Ω)L^{1}((0,T)\times\Omega) with velocity field VnV_{n} (resp. VV). Then ρnρ\rho_{n}\to\rho in L1((0,T)×Ω)L^{1}((0,T)\times\Omega) as nn\to\infty.

Proof.

Since (Vn)n(V_{n})_{n\in\mathbb{N}} converges weakly-\ast to VV in LL^{\infty}, there exists a constant M>0M>0 such that VnL((0,T)×Ω)M\left\lVert V_{n}\right\rVert_{L^{\infty}((0,T)\times\Omega)}\leq M for every nn\in\mathbb{N} and thus, by Proposition 2.2, there exists C>0C>0 depending only on dd, ν\nu, MM, and ρ0L1(Ω)\left\lVert\rho_{0}\right\rVert_{L^{1}(\Omega)} such that, for every nn\in\mathbb{N},

(3.1) ρnL((0,T);L1(Ω))+ρnLq((0,T);W1,q(Ω))+tρnLq((0,T);W1,q(Ω))C.\left\lVert\rho_{n}\right\rVert_{L^{\infty}((0,T);L^{1}(\Omega))}+\left\lVert\rho_{n}\right\rVert_{L^{q}((0,T);W^{1,q}(\Omega))}+\left\lVert\partial_{t}\rho_{n}\right\rVert_{L^{q}((0,T);W^{-1,q}(\Omega))}\leq C.

It follows from (3.1) and Aubin–Lions Lemma (see, e.g., [32, Corollary 4]) that (ρn)n(\rho_{n})_{n\in\mathbb{N}} is relatively compact in L1((0,T)×Ω)L^{1}((0,T)\times\Omega). Let ρL1((0,T)×Ω)\rho^{\ast}\in L^{1}((0,T)\times\Omega) be a limit point of (ρn)n(\rho_{n})_{n\in\mathbb{N}} and (ρnk)k(\rho_{n_{k}})_{k\in\mathbb{N}} a subsequence of (ρn)n(\rho_{n})_{n\in\mathbb{N}} converging to ρ\rho^{\ast} in L1((0,T)×Ω)L^{1}((0,T)\times\Omega).

The weak convergence of VnV_{n} in LL^{\infty} together with the strong convergence of ρn\rho_{n} in L1L^{1} allow to pass to the limit the drift term (ρnVn)\nabla\cdot(\rho_{n}V_{n}) in the equation and we then easily obtain that ρ\rho^{\ast} is a weak solution of (2.1). By the uniqueness of such solution from Proposition 2.2, one concludes ρ=ρ\rho^{\ast}=\rho. In particular, ρ\rho is the unique limit point of the relatively compact sequence (ρn)n(\rho_{n})_{n\in\mathbb{N}} in L1((0,T)×Ω)L^{1}((0,T)\times\Omega), which yields the result. ∎

Proposition 3.5.

Let ν>0\nu>0 and ψH01(Ω)\psi\in H_{0}^{1}(\Omega). Given KL((0,T)×Ω)K\in L^{\infty}((0,T)\times\Omega), let (Kn)n(K_{n})_{n\in\mathbb{N}} be a sequence in L((0,T)×Ω)L^{\infty}((0,T)\times\Omega) such that KnKK_{n}\xrightharpoonup{\ast}K as nn\to\infty. For nn\in\mathbb{N}, let φn\varphi_{n} (resp. φ\varphi) be the unique weak solution of (2.4) in L((0,T);L2(Ω))L2((0,T);H01(Ω))L^{\infty}((0,T);L^{2}(\Omega))\cap L^{2}((0,T);H_{0}^{1}(\Omega)) with KnK_{n} (resp. KK). Then φnφ\varphi_{n}\to\varphi in L2((0,T);H01(Ω))L^{2}((0,T);H_{0}^{1}(\Omega)) as nn\to\infty.

Proof.

Again, there exists a constant M>0M>0 such that KnL((0,T)×Ω)M\left\lVert K_{n}\right\rVert_{L^{\infty}((0,T)\times\Omega)}\leq M for every nn\in\mathbb{N} and thus, by Proposition 2.5, there exists C>0C>0 depending only on dd, ν\nu, TT, Ω\Omega, MM, and ψH01(Ω)\left\lVert\psi\right\rVert_{H_{0}^{1}(\Omega)} such that

(3.2) φnL((0,T);H01(Ω))+φnL2((0,T);H2(Ω))+tφnL2((0,T)×Ω)C.\left\lVert\varphi_{n}\right\rVert_{L^{\infty}((0,T);H_{0}^{1}(\Omega))}+\left\lVert\varphi_{n}\right\rVert_{L^{2}((0,T);H^{2}(\Omega))}+\left\lVert\partial_{t}\varphi_{n}\right\rVert_{L^{2}((0,T)\times\Omega)}\leq C.

Hence, by Aubin–Lions Lemma (see, e.g., [32, Corollary 4]), (φn)n(\varphi_{n})_{n\in\mathbb{N}} is relatively compact in L2((0,T);H01(Ω))L^{2}((0,T);H_{0}^{1}(\Omega)). Let φ\varphi^{\ast} be a limit point of (φn)n(\varphi_{n})_{n\in\mathbb{N}} and (φnk)k(\varphi_{n_{k}})_{k\in\mathbb{N}} be a subsequence of (φn)n(\varphi_{n})_{n\in\mathbb{N}} converging to φ\varphi^{\ast} in L2((0,T);H01(Ω))L^{2}((0,T);H_{0}^{1}(\Omega)). By (3.2), we also have φL((0,T);H01(Ω)))\varphi^{\ast}\in L^{\infty}((0,T);\allowbreak H_{0}^{1}(\Omega))).

Now, because of the non-linearity in the equation, we prefer to provide details on how to pass it to the limit. For every kk and every ηH1((0,T)×Ω)\eta\in H^{1}((0,T)\times\Omega) such that η|[0,T]×Ω=0\eta\bigr{\rvert}_{[0,T]\times\partial\Omega}=0 and η|{0}×Ω=0\eta\bigr{\rvert}_{\{0\}\times\Omega}=0, one has

0TΩφnktη+ν0TΩφnkη+0TΩ(Knk|φnk|1)η=Ωψ(x)η(T,x)dx.\int_{0}^{T}\int_{\Omega}\varphi_{n_{k}}\partial_{t}\eta+\nu\int_{0}^{T}\int_{\Omega}\nabla\varphi_{n_{k}}\cdot\nabla\eta+\int_{0}^{T}\int_{\Omega}(K_{n_{k}}\left\lvert\nabla\varphi_{n_{k}}\right\rvert-1)\eta=\int_{\Omega}\psi(x)\eta(T,x)\operatorname{d\!}x.

Since KnkKK_{n_{k}}\xrightharpoonup{\ast}K in L((0,T)×Ω)L^{\infty}((0,T)\times\Omega) and φnkφ\varphi_{n_{k}}\to\varphi^{\ast} in L2((0,T);H01(Ω))L^{2}((0,T);H_{0}^{1}(\Omega)), one obtains, letting kk\to\infty, that

0TΩφtη+ν0TΩφη+0TΩ(K|φ|1)η=Ωψ(x)η(T,x)dx.\int_{0}^{T}\int_{\Omega}\varphi^{\ast}\partial_{t}\eta+\nu\int_{0}^{T}\int_{\Omega}\nabla\varphi^{\ast}\cdot\nabla\eta+\int_{0}^{T}\int_{\Omega}(K\left\lvert\nabla\varphi^{\ast}\right\rvert-1)\eta=\int_{\Omega}\psi(x)\eta(T,x)\operatorname{d\!}x.

Hence φL((0,T);L2(Ω))L2((0,T);H01(Ω))\varphi^{\ast}\in L^{\infty}((0,T);L^{2}(\Omega))\cap L^{2}((0,T);H_{0}^{1}(\Omega)) is a weak solution of (2.4) and, by the uniqueness of solutions of (2.4) from Proposition 2.5, one deduces that φ=φ\varphi^{\ast}=\varphi. Thus φ\varphi is the unique limit point in L2((0,T);H01(Ω))L^{2}((0,T);H_{0}^{1}(\Omega)) of the relatively compact sequence (φn)n(\varphi_{n})_{n\in\mathbb{N}}, yielding the conclusion. ∎

We now recall the statement of Kakutani’s fixed point theorem (see, e.g., [13, § 7, Theorem 8.6]), which is used in the proof of Theorem 3.3.

Theorem 3.6 (Kakutani’s fixed point theorem).

Let \mathcal{B} be a compact convex subset of a locally convex topological vector space EE, and let 𝒱\mathcal{V} be a set-valued map in \mathcal{B}, i.e., 𝒱\mathcal{V} associates, with each bb\in\mathcal{B}, a set 𝒱(b)\mathcal{V}(b)\subset\mathcal{B}. Assume that 𝒱\mathcal{V} is upper semi-continuous and that, for every bb\in\mathcal{B}, 𝒱(b)\mathcal{V}(b) is non-empty, compact, and convex. Then 𝒱\mathcal{V} admits a fixed point in \mathcal{B}, i.e., there exists bb\in\mathcal{B} such that bS(b)b\in S(b).

We are finally in position to provide the proof of Theorem 3.3.

Proof of Theorem 3.3.

Let κ0\kappa_{0} be an upper bound on κ\kappa. We endow the space L((0,T)×Ω;d)L^{\infty}((0,T)\times\Omega;\mathbb{R}^{d}) with its weak-\ast topology and consider the ball of radius κ0\kappa_{0} given by

={VL((0,T)×Ω;d)VL((0,T)×Ω;d)κ0}.\mathcal{B}=\left\{V\in L^{\infty}((0,T)\times\Omega;\mathbb{R}^{d})\mid\lVert V\rVert_{L^{\infty}((0,T)\times\Omega;\mathbb{R}^{d})}\leq\kappa_{0}\right\}.

Note that \mathcal{B} is clearly convex and, by the Banach–Alaoglu theorem, \mathcal{B} is a compact subset of L((0,T)×Ω;d)L^{\infty}((0,T)\times\Omega;\mathbb{R}^{d}).

Let 𝒮FP:L((0,T)×Ω;d)L1((0,T)×Ω)\mathcal{S}_{\mathrm{FP}}:L^{\infty}((0,T)\times\Omega;\mathbb{R}^{d})\to L^{1}((0,T)\times\Omega) be the function that associates, with each VL((0,T)×Ω;d)V\in L^{\infty}((0,T)\times\Omega;\mathbb{R}^{d}), the unique weak solution ρ=𝒮FP(V)L1((0,T)×Ω)\rho=\mathcal{S}_{\mathrm{FP}}(V)\in L^{1}((0,T)\times\Omega) of (2.1) with initial condition ρ0\rho_{0}. Note that, by Proposition 3.4, 𝒮FP\mathcal{S}_{\mathrm{FP}} is continuous with respect to the weak-\ast topology of L((0,T)×Ω;d)L^{\infty}((0,T)\times\Omega;\mathbb{R}^{d}) and the strong topology of L1((0,T)×Ω)L^{1}((0,T)\times\Omega). Similarly, we define 𝒮HJB:L((0,T)×Ω)L2((0,T);H01(Ω))\mathcal{S}_{\mathrm{HJB}}:L^{\infty}((0,T)\times\Omega)\to L^{2}((0,T);H_{0}^{1}(\Omega)) as the function that associates, with each KL((0,T)×Ω)K\in L^{\infty}((0,T)\times\Omega), the unique weak solution φ=𝒮HJB(K)L2((0,T);H01(Ω))\varphi=\mathcal{S}_{\mathrm{HJB}}(K)\in L^{2}((0,T);H_{0}^{1}(\Omega)) of (2.4) with terminal condition ψ\psi. Proposition 3.5 ensures that 𝒮HJB\mathcal{S}_{\mathrm{HJB}} is continuous with respect to the weak-\ast topology of L((0,T)×Ω)L^{\infty}((0,T)\times\Omega) and the strong topology of L2((0,T);H01(Ω))L^{2}((0,T);H_{0}^{1}(\Omega)).

We define the set-valued map 𝒱\mathcal{V} that, with each VV\in\mathcal{B}, associates the set 𝒱(V)\mathcal{V}(V)\subset\mathcal{B} given by

𝒱(V)={V~|\displaystyle\mathcal{V}(V)=\Big{\{}\widetilde{V}\in\mathcal{B}\mathrel{}\Big{|}\mathrel{} |V~(t,x)|κ(ρ(t,x)) for a.e. (t,x)(0,T)×Ω,\displaystyle\big{\lvert}\widetilde{V}(t,x)\big{\rvert}\leq\kappa(\rho(t,x))\text{ for a.e.\ }(t,x)\in(0,T)\times\Omega,
V~(t,x)φ(t,x)=κ(ρ(t,x))|φ(t,x)| for a.e. (t,x)(0,T)×Ω,\displaystyle\widetilde{V}(t,x)\cdot\nabla\varphi(t,x)=-\kappa(\rho(t,x))\left\lvert\nabla\varphi(t,x)\right\rvert\text{ for a.e.\ }(t,x)\in(0,T)\times\Omega,
where ρ=𝒮FP(V) and φ=𝒮HJB(κρ)}.\displaystyle\text{where }\rho=\mathcal{S}_{\mathrm{FP}}(V)\text{ and }\varphi=\mathcal{S}_{\mathrm{HJB}}(\kappa\circ\rho)\Big{\}}.

In order to prove the existence of a weak solution (ρ,φ)(\rho,\varphi) of (1.3), we first prove the existence of a fixed point of the set-valued map 𝒱\mathcal{V}, i.e., of a VL((0,T)×Ω;d)V\in L^{\infty}((0,T)\times\Omega;\mathbb{R}^{d}) such that V𝒱(V)V\in\mathcal{V}(V). This is done by applying Kakutani’s fixed point theorem to the set-valued map 𝒱\mathcal{V}. To do so, we first need to verify some properties of 𝒱\mathcal{V} and its graph 𝒢\mathcal{G} defined by

𝒢={(V,V~)×V~𝒱(V)}.\mathcal{G}=\left\{(V,\widetilde{V})\in\mathcal{B}\times\mathcal{B}\mid\widetilde{V}\in\mathcal{V}(V)\right\}.
Claim 1.

For every VL((0,T)×Ω;d)V\in L^{\infty}((0,T)\times\Omega;\mathbb{R}^{d}), the set 𝒱(V)\mathcal{V}(V) is non-empty and convex.

Proof.

It is immediate to verify that 𝒱(V)\mathcal{V}(V) is convex. To prove that it is non-empty, let VL((0,T)×Ω;d)V\in L^{\infty}((0,T)\times\Omega;\mathbb{R}^{d}), ρ=𝒮FP(V)\rho=\mathcal{S}_{\mathrm{FP}}(V), and φ=𝒮HJB(κρ)\varphi=\mathcal{S}_{\mathrm{HJB}}(\kappa\circ\rho). Then, the function V~L((0,T)×Ω;d)\widetilde{V}\in L^{\infty}((0,T)\times\Omega;\mathbb{R}^{d}) defined for a.e. (t,x)(0,T)×Ω(t,x)\in(0,T)\times\Omega by

V~(t,x)={κ(ρ(t,x))φ(t,x)|φ(t,x)|if φ(t,x)0,0otherwise,\widetilde{V}(t,x)=\begin{dcases*}-\kappa(\rho(t,x))\frac{\nabla\varphi(t,x)}{\left\lvert\nabla\varphi(t,x)\right\rvert}&if $\nabla\varphi(t,x)\neq 0$,\\ 0&otherwise,\end{dcases*}

clearly satisfies V~𝒱(V)\widetilde{V}\in\mathcal{V}(V). ∎

Claim 2.

The graph 𝒢\mathcal{G} is a closed subset of ×\mathcal{B}\times\mathcal{B}.

Proof.

Let (Vn,V~n)n(V_{n},\widetilde{V}_{n})_{n\in\mathbb{N}} be a sequence in 𝒢\mathcal{G} converging weakly-\ast in ×\mathcal{B}\times\mathcal{B} to a point (V,V~)(V,\widetilde{V}). We want to prove (V,V~)𝒢(V,\widetilde{V})\in\mathcal{G}, i.e., V~𝒱(V)\widetilde{V}\in\mathcal{V}(V).

Define, for nn\in\mathbb{N}, the functions ρnL1((0,T)×Ω)\rho_{n}\in L^{1}((0,T)\times\Omega) and φnL2((0,T);H01(Ω))\varphi_{n}\in L^{2}((0,T);H_{0}^{1}(\Omega)) by ρn=𝒮FP(Vn)\rho_{n}=\mathcal{S}_{\mathrm{FP}}(V_{n}) and φn=𝒮HJB(κρn)\varphi_{n}=\mathcal{S}_{\mathrm{HJB}}(\kappa\circ\rho_{n}) and, similarly, let ρ=𝒮FP(V)\rho=\mathcal{S}_{\mathrm{FP}}(V) and φ=𝒮HJB(κρ)\varphi=\mathcal{S}_{\mathrm{HJB}}(\kappa\circ\rho). Since 𝒮FP:L((0,T)×Ω;d)L1((0,T)×Ω)\mathcal{S}_{\mathrm{FP}}:L^{\infty}((0,T)\times\Omega;\mathbb{R}^{d})\to L^{1}((0,T)\times\Omega) is continuous with respect to the weak-\ast topology of L((0,T)×Ω;d)L^{\infty}((0,T)\times\Omega;\mathbb{R}^{d}) and the strong topology of L1((0,T)×Ω)L^{1}((0,T)\times\Omega), one deduces ρnρ\rho_{n}\to\rho in L1((0,T)×Ω)L^{1}((0,T)\times\Omega) as nn\to\infty. Hence, up to extracting subsequences (which we still denote using the same notation for simplicity), one has ρnρ\rho_{n}\to\rho a.e. in (0,T)×Ω(0,T)\times\Omega. Since κ\kappa is continuous, we deduce κρnκρ\kappa\circ\rho_{n}\to\kappa\circ\rho a.e. in (0,T)×Ω(0,T)\times\Omega, and it follows κρnκρ\kappa\circ\rho_{n}\xrightharpoonup{\ast}\kappa\circ\rho in L((0,T)×Ω)L^{\infty}((0,T)\times\Omega). The continuity of 𝒮HJB:L((0,T)×Ω;d)L2((0,T);H01(Ω))\mathcal{S}_{\mathrm{HJB}}:L^{\infty}((0,T)\times\Omega;\mathbb{R}^{d})\to L^{2}((0,T);H_{0}^{1}(\Omega)) with respect to the weak-\ast topology of L((0,T)×Ω;d)L^{\infty}((0,T)\times\Omega;\mathbb{R}^{d}) and the strong topology of L2((0,T);H01(Ω))L^{2}((0,T);H_{0}^{1}(\Omega)) implies φnφ\varphi_{n}\to\varphi in L2((0,T);H01(Ω))L^{2}((0,T);H_{0}^{1}(\Omega)) as nn\to\infty.

From the weak convergence of V~n\widetilde{V}_{n} to V~\widetilde{V}, the convexity of the function ||\lvert\cdot\rvert, and the (strong) convergence of κ(ρn)\kappa(\rho_{n}) to κ(ρ)\kappa(\rho), the inequality |V~n|κ(ρn)\big{\lvert}\widetilde{V}_{n}\big{\rvert}\leq\kappa(\rho_{n}) gives at the limit

(3.3) |V~(t,x)|κ(ρ(t,x)) for a.e. (t,x)(0,T)×Ω.\big{\lvert}\widetilde{V}(t,x)\big{\rvert}\leq\kappa(\rho(t,x))\qquad\text{ for a.e.\ }(t,x)\in(0,T)\times\Omega.

Since V~n𝒱(Vn)\widetilde{V}_{n}\in\mathcal{V}(V_{n}) for every nn\in\mathbb{N}, we have V~n(t,x)φn(t,x)=κ(ρn(t,x))|φn(t,x)|\widetilde{V}_{n}(t,x)\cdot\nabla\varphi_{n}(t,x)=-\kappa(\rho_{n}(t,x))\left\lvert\nabla\varphi_{n}(t,x)\right\rvert for a.e. (t,x)(0,T)×Ω(t,x)\in(0,T)\times\Omega. Then, for every vL2((0,T)×Ω)v\in L^{2}((0,T)\times\Omega), one has

0TΩV~n(t,x)φn(t,x)v(t,x)dxdt=0TΩκ(ρn(t,x))|φn(t,x)|v(t,x)dxdt.\int_{0}^{T}\int_{\Omega}\widetilde{V}_{n}(t,x)\cdot\nabla\varphi_{n}(t,x)v(t,x)\operatorname{d\!}x\operatorname{d\!}t=-\int_{0}^{T}\int_{\Omega}\kappa(\rho_{n}(t,x))\left\lvert\nabla\varphi_{n}(t,x)\right\rvert v(t,x)\operatorname{d\!}x\operatorname{d\!}t.

Recalling that, as nn\to\infty, one has V~nV~\widetilde{V}_{n}\xrightharpoonup{\ast}\widetilde{V} in L((0,T)×Ω)L^{\infty}((0,T)\times\Omega), φnφ\nabla\varphi_{n}\to\nabla\varphi in L2((0,T)×Ω)L^{2}((0,T)\times\Omega), and κρnκρ\kappa\circ\rho_{n}\xrightharpoonup{\ast}\kappa\circ\rho in L((0,T)×Ω)L^{\infty}((0,T)\times\Omega), we obtain, letting nn\to\infty, that

0TΩV~(t,x)φ(t,x)v(t,x)dxdt=0TΩκ(ρ(t,x))|φ(t,x)|v(t,x)dxdt\int_{0}^{T}\int_{\Omega}\widetilde{V}(t,x)\cdot\nabla\varphi(t,x)v(t,x)\operatorname{d\!}x\operatorname{d\!}t=-\int_{0}^{T}\int_{\Omega}\kappa(\rho(t,x))\left\lvert\nabla\varphi(t,x)\right\rvert v(t,x)\operatorname{d\!}x\operatorname{d\!}t

for every vL2((0,T)×Ω)v\in L^{2}((0,T)\times\Omega), which implies that

(3.4) V~(t,x)φ(t,x)=κ(ρ(t,x))|φ(t,x)| for a.e. (t,x)(0,T)×Ω.\widetilde{V}(t,x)\cdot\nabla\varphi(t,x)=-\kappa(\rho(t,x))\left\lvert\nabla\varphi(t,x)\right\rvert\qquad\text{ for a.e.\ }(t,x)\in(0,T)\times\Omega.

Combining (3.3) and (3.4), we conclude that V~𝒱(V)\widetilde{V}\in\mathcal{V}(V), as required. ∎

Claim 3.

For every VL((0,T)×Ω;d)V\in L^{\infty}((0,T)\times\Omega;\mathbb{R}^{d}), the set 𝒱(V)\mathcal{V}(V) is compact.

Proof.

This is a consequence of the fact that 𝒢\mathcal{G} is a closed subset of the compact set ×\mathcal{B}\times\mathcal{B}. ∎

Thanks to Claims 2 and 3, it follows from [3, Proposition 1.4.8] that the set-valued map 𝒱\mathcal{V} is upper semi-continuous. Using this fact and Claims 1 and 3, it follows from Kakutani’s fixed point theorem that 𝒱\mathcal{V} admits a fixed point VV\in\mathcal{B}. Let ρ=𝒮FP(V)\rho=\mathcal{S}_{\mathrm{FP}}(V) and φ=𝒮HJB(κρ)\varphi=\mathcal{S}_{\mathrm{HJB}}(\kappa\circ\rho). Using the facts that ρ\rho and φ\varphi are solutions of (2.1) and (2.4), respectively, and that V𝒱(V)V\in\mathcal{V}(V), it is immediate to verify, using Definitions 2.1, 2.3, and 3.1, that (ρ,φ)(\rho,\varphi) is a weak solution of (1.3) with initial condition ρ0\rho_{0} and final condition ψ\psi, as required. ∎

4. The MFG system with an infinite time horizon

Now that we have established in Section 3 the existence of solutions to the Mean Field Game system (1.3) in a finite time horizon TT, we consider in this section the Mean Field Game system (1.2) with an infinite time horizon. Let us first provide the definition of a weak solution in this setting.

Definition 4.1.

Let ν>0\nu>0, κ:(0,+)\kappa:\mathbb{R}\to(0,+\infty) be continuous and bounded, and ρ0L1(Ω)\rho_{0}\in L^{1}(\Omega). We say that (ρ,φ)Lloc(+;L1(Ω))×Lloc2(+;H01(Ω))(\rho,\varphi)\in L^{\infty}_{\mathrm{loc}}(\mathbb{R}_{+};L^{1}(\Omega))\times L^{2}_{\mathrm{loc}}(\mathbb{R}_{+};H_{0}^{1}(\Omega)) is a weak solution of (1.2) with initial condition ρ0\rho_{0} if φL(+×Ω)\varphi\in L^{\infty}(\mathbb{R}_{+}\times\Omega) and if there exists VL(+×Ω;d)V\in L^{\infty}(\mathbb{R}_{+}\times\Omega;\mathbb{R}^{d}) such that |V(t,x)|κ(ρ(t,x))\lvert V(t,x)\rvert\leq\kappa(\rho(t,x)) and V(t,x)φ(t,x)=κ(ρ(t,x))|φ(t,x)|V(t,x)\cdot\nabla\varphi(t,x)=-\kappa(\rho(t,x))\left\lvert\nabla\varphi(t,x)\right\rvert a.e. on +×Ω\mathbb{R}_{+}\times\Omega and such that, for every T>0T>0, ρ\rho is a solution of the Fokker–Planck equation (2.1) with initial datum ρ0\rho_{0} and vector field VV on [0,T]×Ω[0,T]\times\Omega in the sense of Definition 2.1 and φ\varphi is a solution of the Hamilton–Jacobi–Bellman equation (2.4) with K=κ(ρ)K=\kappa(\rho) in the sense of Definition 2.3 on the same domain111Note that Definition 2.3 requires to fix a final value, and we did not define the notion of solution independently of the final value ψ\psi. This could be formalized as “there exists ψL2(Ω)\psi\in L^{2}(\Omega) such that φ\varphi is a solution of (2.4)”. Yet, since the function φ\varphi will be finally continuous as a function valued into L2(Ω)L^{2}(\Omega), the final datum on [0,T][0,T] will be necessarily given by its own value φ(T,)\varphi(T,\cdot)..

Notice that, with respect to Definition 3.1, we make the additional requirement that φL(+×Ω)\varphi\in L^{\infty}(\mathbb{R}_{+}\times\Omega). This is done mainly for three reasons. Firstly, boundedness of the solution of a Hamilton–Jacobi–Bellman equation is a condition usually required in order to ensure that this solution is the value function of an optimal control problem (see, e.g., [5, Theorem 8.1.10] and [10, Chapter II, Corollary 9.1]). Secondly, the strategy we use in this section to prove existence of a solution of (1.2), based on a limit argument from solutions of (1.3) in finite time horizon TT as T+T\to+\infty, allows us to ensure that the function φ:+×Ω\varphi:\mathbb{R}_{+}\times\Omega\to\mathbb{R} we construct is indeed bounded. Finally, boundedness of φ\varphi is an important property in order to establish the results on the the asymptotic behavior of solutions to (1.2) provided in Theorem 4.2 and Propositions 4.5 and 4.6.

4.1. Existence of solutions and their asymptotic behavior

From now on, we let Ψ\Psi denote the solution of the (stationary) Hamilton–Jacobi–Bellman equation

(4.1) νΔΨ+κ(0)|Ψ|=1,xΩ,-\nu\Delta\Psi+\kappa(0)|\nabla\Psi|=1,\quad x\in\Omega,

with Dirichlet boundary conditions Ψ=0\Psi=0 on Ω\partial\Omega. Existence of such a solution Ψ\Psi follows from standard results on elliptic equations, and Ψ\Psi is continuous in the closure of Ω\Omega and C2C^{2} and positive in Ω\Omega (see, e.g., [11, Theorem 15.12], [20, Chapter 4, Section 8], [25]; these results require additional regularity properties on Ω\partial\Omega but they can be easily adapted to a C2C^{2} boundary using the techniques from [11, Section 15.6] or [20, Chapter 4, pp. 309–310]). Uniqueness of Ψ\Psi follows also from classical arguments for elliptic equations based on the maximum principle: the difference Ψ~\widetilde{\Psi} of two solutions of (4.1) is zero on Ω\partial\Omega and satisfies νΔΨ~κ(0)|Ψ~|0-\nu\Delta\widetilde{\Psi}-\kappa(0)\lvert\nabla\widetilde{\Psi}\rvert\leq 0 in Ω\Omega, and hence the maximum principle from [11, Theorem 10.9] allows one to conclude that Ψ~=0\widetilde{\Psi}=0 in Ω\Omega.

The main result of this section is the following.

Theorem 4.2.

Let ρ0L1(Ω)\rho_{0}\in L^{1}(\Omega). Then, there exists at least one solution (ρ,φ)(\rho,\varphi) to the Mean Field Game system with infinite time horizon (1.2).

In addition, any such solution satisfies

ρtt+0,φtt+Ψ,\rho_{t}\underset{t\to+\infty}{\longrightarrow}0,\quad\varphi_{t}\underset{t\to+\infty}{\longrightarrow}\Psi,

and the above convergences hold uniformly.

The sequel of this section is devoted to the proof of Theorem 4.2. Let us start by giving an idea of the proof. First, we will construct solutions to the problem with infinite time horizon as limits of solutions of the problem with finite time horizon TT by letting TT go to ++\infty. Then, to prove the long-time uniform convergence of the solutions, we shall make a crucial use of some regularity results for parabolic equations. More precisely, we will use local maximum principles for Fokker–Planck and for (forward) Hamilton–Jacobi–Bellman equations; roughly speaking, these results state that the L(Ω)L^{\infty}(\Omega) norm of solutions of such equations at some time t2t_{2} is controlled by some LpL^{p} norms of the same solution at some previous time t1<t2t_{1}<t_{2}. The results we use are proved in Appendix A, see Proposition A.1 and Corollaries A.2 and A.3.

We start with a lemma that gathers some useful estimates. These estimates have already been discussed in Section 2, but we need now to track possible dependencies of the constant on the time horizon TT.

Lemma 4.3.

Let (ρ,φ)(\rho,\varphi) be solution of the finite horizon MFG system (1.3) on [0,T]×Ω[0,T]\times\Omega in the sense of Definition 3.1, with final datum ψH01(Ω)L(Ω)\psi\in H_{0}^{1}(\Omega)\cap L^{\infty}(\Omega) with ψ0\psi\geq 0. Then, there are C1,C2>0C_{1},C_{2}>0, depending on ψL+ψH01\lVert\psi\rVert_{L^{\infty}}+\lVert\psi\rVert_{H^{1}_{0}}, supκ\sup\kappa, ν\nu, Ψ\Psi and Ω\Omega such that

(4.2) φ(t,)L2C1,for all t[0,T],\lVert\nabla\varphi(t,\cdot)\rVert_{L^{2}}\leq C_{1},\quad\text{for all }\ t\in[0,T],

and

(4.3) φL2((T1,T2);H2)C2(1+|T2T1|).\lVert\varphi\rVert_{L^{2}((T_{1},T_{2});H^{2})}\leq C_{2}(1+|T_{2}-T_{1}|).
Proof.

Step 1. A preliminary estimate.

Let us start with giving an estimate on the gradient of φ\varphi. First, multiplying by φ\varphi the equation satisfied by φ\varphi and integrating on Ω\Omega for a fixed t(0,T)t\in(0,T), we find

ddt(12Ωφ2)=νΩ|φ|2Ωκ(ρ)|φ|φ+Ωφ.-\frac{\operatorname{d\!}}{\operatorname{d\!}t}\left(\frac{1}{2}\int_{\Omega}\varphi^{2}\right)=-\nu\int_{\Omega}|\nabla\varphi|^{2}-\int_{\Omega}\kappa(\rho)|\nabla\varphi|\varphi+\int_{\Omega}\varphi.

Therefore, since φ\varphi is bounded, we have

(4.4) T1T2Ω|φ|2C(1+|T2T1|),\int_{T_{1}}^{T_{2}}\int_{\Omega}|\nabla\varphi|^{2}\leq C(1+|T_{2}-T_{1}|),

for some C>0C>0 depending on supφ\sup\varphi, |Ω|\lvert\Omega\rvert, supκ\sup\kappa, ν\nu and for every T1,T2[0,T]T_{1},T_{2}\in[0,T] with 0T1T2T0\leq T_{1}\leq T_{2}\leq T. Note that, from Proposition 2.5, supφ\sup\varphi is bounded in terms of ν\nu, Ω\Omega, and ψL\lVert\psi\rVert_{L^{\infty}}.

Step 2. Bound on φ(t,)L2\lVert\nabla\varphi(t,\cdot)\rVert_{L^{2}}.

We define, for t[0,T]t\in[0,T],

u(t):=12Ω|φ(t,x)|2dx.u(t):=\frac{1}{2}\int_{\Omega}|\nabla\varphi(t,x)|^{2}\operatorname{d\!}x.

We differentiate uu to obtain

(4.5) u(t)=νΩ(Δφ)2Ωκ(ρ)|φ|Δφ+ΩΔφ.u^{\prime}(t)=\nu\int_{\Omega}(\Delta\varphi)^{2}-\int_{\Omega}\kappa(\rho)\lvert\nabla\varphi\rvert\Delta\varphi+\int_{\Omega}\Delta\varphi.

Using Young’s inequality, we find that there are K1,K2>0K_{1},K_{2}>0 depending only on |Ω|\lvert\Omega\rvert, supκ\sup\kappa, ν\nu such that

u(t)+K1u(t)+K20.u^{\prime}(t)+K_{1}u(t)+K_{2}\geq 0.

This implies, for any 0t<sT0\leq t<s\leq T,

(4.6) u(s)+K2K1(1eK1(st))u(t)eK1(st).u(s)+\frac{K_{2}}{K_{1}}\left(1-e^{-K_{1}(s-t)}\right)\geq u(t)e^{-K_{1}(s-t)}.

We integrate (4.6) for s(t,t+1)s\in(t,t+1) to get

12tt+1Ω|φ|2(s,x)dxds+K2K101(1eK1r)dr12(01eK1rdr)Ω|φ|2(t,x)dx.\frac{1}{2}\int_{t}^{t+1}\int_{\Omega}|\nabla\varphi|^{2}(s,x)\operatorname{d\!}x\operatorname{d\!}s+\frac{K_{2}}{K_{1}}\int_{0}^{1}(1-e^{-K_{1}r})\operatorname{d\!}r\geq\frac{1}{2}\left(\int_{0}^{1}e^{-K_{1}r}\operatorname{d\!}r\right)\int_{\Omega}|\nabla\varphi|^{2}(t,x)\operatorname{d\!}x.

Using (4.4) yields the L(H1)L^{\infty}(H^{1}) bound (4.2) for t[0,T1)t\in[0,T-1). To get the L(H1)L^{\infty}(H^{1}) bound (4.2) for t[T1,T]t\in[T-1,T], we use (4.6) with s=Ts=T. The result follows, with a constant also depending on u(T)=12Ω|ψ|2<+u(T)=\frac{1}{2}\int_{\Omega}|\nabla\psi|^{2}<+\infty.

Step 3. Bound in L2((T1,T2);H2)L^{2}((T_{1},T_{2});H^{2}).

Let us integrate (4.5) on (T1,T2)(T_{1},T_{2}). We find

νT1T2Ω(Δφ)2=u(T2)u(T1)+T1T2Ωκ(ρ)|φ|ΔφT1T2ΩΔφ.\nu\int_{T_{1}}^{T_{2}}\int_{\Omega}(\Delta\varphi)^{2}=u(T_{2})-u(T_{1})+\int_{T_{1}}^{T_{2}}\int_{\Omega}\kappa(\rho)|\nabla\varphi|\Delta\varphi-\int_{T_{1}}^{T_{2}}\int_{\Omega}\Delta\varphi.

Using Young’s inequality on T1T2Ωκ(ρ)|φ|Δφ\int_{T_{1}}^{T_{2}}\int_{\Omega}\kappa(\rho)|\nabla\varphi|\Delta\varphi and T1T2ΩΔφ\int_{T_{1}}^{T_{2}}\int_{\Omega}\Delta\varphi and the estimate (4.4), we get the desired bound (4.3) on L2((T1,T2);H2)L^{2}((T_{1},T_{2});H^{2}). ∎

The next lemma shows that the time derivative of Ωρφ\int_{\Omega}\rho\varphi is equal to Ωρ-\int_{\Omega}\rho. Differentiating the average value of the value function is a classical computation in Mean Field Game theory. Since here the value function is an exit time, it is expected that it should decrease with rate 11, and one can guess the result from the fact that the total mass of the agents in this model is not fixed but decreases in time and is equal to Ωρ\int_{\Omega}\rho.

Lemma 4.4.

Let (ρ,φ)(\rho,\varphi) be a solution of the finite-horizon MFG (1.3) on [0,T]×Ω[0,T]\times\Omega in the sense of Definition 3.1. Then, for a.e. tt, we have

ddt(Ωρ(t,x)φ(t,x)dx)=Ωρ(t,x)dx.\frac{\operatorname{d\!}}{\operatorname{d\!}t}\left(\int_{\Omega}\rho(t,x)\varphi(t,x)\operatorname{d\!}x\right)=-\int_{\Omega}\rho(t,x)\operatorname{d\!}x.
Proof.

Let us fix two instants of times t1<t2Tt_{1}<t_{2}\leq T, with t1>0t_{1}>0. On the interval (t1,t2)(t_{1},t_{2}) we can use φ\varphi as a test function in (2.3) and ρ\rho in (2.6) since both φ\varphi and ρ\rho are continuous as curves valued in L2L^{2}, belong to L2((t1,t2);H01(Ω))L^{2}((t_{1},t_{2});H^{1}_{0}(\Omega)), and their time-derivatives belong to L2((t1,t2);H1(Ω))L^{2}((t_{1},t_{2});H^{-1}(\Omega)). We subtract the two equalities that we obtain, which provides

t1t2Ωρtφdxdtt1t2Ω(νρρV)φdxdt+t1t2Ω(φtρ+νφρ+(κ(ρ)|φ|1)ρ)=2Ωφ(t2,x)ρ(t2,x)dx2Ωφ(t1,x)ρ(t1,x)dx.\int_{t_{1}}^{t_{2}}\int_{\Omega}\rho\partial_{t}\varphi\operatorname{d\!}x\operatorname{d\!}t-\int_{t_{1}}^{t_{2}}\int_{\Omega}\left(\nu\nabla\rho-\rho V\right)\cdot\nabla\varphi\operatorname{d\!}x\operatorname{d\!}t\\ +\int_{t_{1}}^{t_{2}}\int_{\Omega}\left(\varphi\partial_{t}\rho+\nu\nabla\varphi\cdot\nabla\rho+(\kappa(\rho)\left\lvert\nabla\varphi\right\rvert-1)\rho\right)\\ =2\int_{\Omega}\varphi(t_{2},x)\rho(t_{2},x)\operatorname{d\!}x-2\int_{\Omega}\varphi(t_{1},x)\rho(t_{1},x)\operatorname{d\!}x.

After canceling the terms with ρφ\nabla\rho\cdot\nabla\varphi and using Vφ+κ(ρ)|φ|=0V\cdot\nabla\varphi+\kappa(\rho)\lvert\nabla\varphi\rvert=0 we are left with

t1t2Ω(ρtφ+φtρ)dxdtt1t2Ωρdxdt=2Ωφ(t2,x)ρ(t2,x)dx2Ωφ(t1,x)ρ(t1,x)dx.\int_{t_{1}}^{t_{2}}\int_{\Omega}(\rho\partial_{t}\varphi+\varphi\partial_{t}\rho)\operatorname{d\!}x\operatorname{d\!}t-\int_{t_{1}}^{t_{2}}\int_{\Omega}\rho\operatorname{d\!}x\operatorname{d\!}t\\ =2\int_{\Omega}\varphi(t_{2},x)\rho(t_{2},x)\operatorname{d\!}x-2\int_{\Omega}\varphi(t_{1},x)\rho(t_{1},x)\operatorname{d\!}x.

It is then easy to see, by approximation via smooth functions, that for every pair (ρ,φ)(\rho,\varphi) such that ρ,φL2((t1,t2);H1(Ω))\rho,\varphi\in L^{2}((t_{1},t_{2});H^{1}(\Omega)) and tρ,tφL2((t1,t2);H1(Ω))\partial_{t}\rho,\partial_{t}\varphi\in L^{2}((t_{1},t_{2});H^{-1}(\Omega)), we have

t1t2Ω(ρtφ+φtρ)dxdt=Ωφ(t2,x)ρ(t2,x)dxΩφ(t1,x)ρ(t1,x)dx.\int_{t_{1}}^{t_{2}}\int_{\Omega}(\rho\partial_{t}\varphi+\varphi\partial_{t}\rho)\operatorname{d\!}x\operatorname{d\!}t=\int_{\Omega}\varphi(t_{2},x)\rho(t_{2},x)\operatorname{d\!}x-\int_{\Omega}\varphi(t_{1},x)\rho(t_{1},x)\operatorname{d\!}x.

We are then left with

Ωφ(t2,x)ρ(t2,x)dxΩφ(t1,x)ρ(t1,x)dx=t1t2Ωρdxdt,\int_{\Omega}\varphi(t_{2},x)\rho(t_{2},x)\operatorname{d\!}x-\int_{\Omega}\varphi(t_{1},x)\rho(t_{1},x)\operatorname{d\!}x=-\int_{t_{1}}^{t_{2}}\int_{\Omega}\rho\operatorname{d\!}x\operatorname{d\!}t,

which is equivalent to the claim. ∎

We are now in position to prove Theorem 4.2.

Proof of Theorem 4.2.

Let ρ0L1(Ω)\rho_{0}\in L^{1}(\Omega) be fixed.

Step 1. Existence.

For T>0T>0, we let (ρT,φT)(\rho^{T},\varphi^{T}) denote a solution of (1.3) with T>0T>0, with initial datum ρ0\rho_{0} for ρ\rho and with final datum ψT\psi^{T} for φ\varphi, where (ψT)T>0(\psi^{T})_{T>0} is any family of non-negative functions, bounded in L(Ω)H01(Ω)L^{\infty}(\Omega)\cap H^{1}_{0}(\Omega).

Recall that, by Proposition 2.5, φTL((0,T)×Ω)\lVert\varphi^{T}\rVert_{L^{\infty}((0,T)\times\Omega)} is bounded independently of TT. Let 0<T1<T20<T_{1}<T_{2} be fixed. Lemma 4.3 implies that, as soon as T>T2T>T_{2}, φT\varphi^{T} is bounded in L2((T1,T2);H2(Ω))L^{2}((T_{1},T_{2});H^{2}(\Omega)) independently of T>0T>0. Moreover, because tφTL2((T1,T2)×Ω)\partial_{t}\varphi^{T}\in L^{2}((T_{1},T_{2})\times\Omega) owing to Proposition 2.5, we can apply Aubin–Lions Lemma to the sequence (φT)T>0(\varphi^{T})_{T>0} to get that, up to extraction, it converges strongly in Lloc2((0,+);H1(Ω))L^{2}_{loc}((0,+\infty);H^{1}(\Omega)) to some limit φ\varphi_{\infty}. Up to another extraction, we ensure that the convergence of φT,φT\varphi^{T},\nabla\varphi^{T} also holds pointwise.

Using Aubin–Lions Lemma for the sequence (ρT)T>0(\rho^{T})_{T>0} as in the proof of Proposition 3.4, we find that, up to another extraction, it converges strongly to a limit ρ\rho_{\infty} in L2((T1,T2)×Ω)L^{2}((T_{1},T_{2})\times\Omega) and weakly in L2((T1,T2);H01(Ω))L^{2}((T_{1},T_{2});H^{1}_{0}(\Omega)). The solutions (ρT,φT)(\rho^{T},\varphi^{T}) are associated with a bounded vector field VTV_{T}, which will converge weakly-\ast in LL^{\infty} to a vector field VV_{\infty}. Using the same arguments as in the proof of Theorem 3.3, Claim 2, we can pass to the limit T+T\to+\infty in the equation to find that the pair (ρ,φ)(\rho_{\infty},\varphi_{\infty}) solves (1.2).

Step 2. Long-time behavior of ρ\rho.

Let (ρ,φ)(\rho,\varphi) be a solution of (1.2), as built in the previous step. The integral version of Lemma 4.4, which is valid for (ρT,φT)(\rho^{T},\varphi^{T}), also applies to (ρ,φ)(\rho,\varphi) at the limit, and we have

ddtΩρ(t,x)φ(t,x)dx1supφ(Ωρ(t,x)φ(t,x)dx),\frac{\operatorname{d\!}}{\operatorname{d\!}t}\int_{\Omega}\rho(t,x)\varphi(t,x)\operatorname{d\!}x\leq\frac{-1}{\sup\varphi}\left(\int_{\Omega}\rho(t,x)\varphi(t,x)\operatorname{d\!}x\right),

hence, for all t0t\geq 0, we have

Ωρ(t,x)φ(t,x)dx(Ωρ0(x)φ(0,x)dx)e1supφt.\int_{\Omega}\rho(t,x)\varphi(t,x)\operatorname{d\!}x\leq\left(\int_{\Omega}\rho_{0}(x)\varphi(0,x)\operatorname{d\!}x\right)e^{-\frac{1}{\sup\varphi}t}.

Moreover, using the fact that tΩρ(t,x)dxt\mapsto\int_{\Omega}\rho(t,x)\operatorname{d\!}x is non-increasing, we get, integrating the relation from Lemma 4.4,

Ωρ(t,x)dxt1tΩρ(τ,x)dxdτΩρ(t1,x)φ(t1,x)dx,\int_{\Omega}\rho(t,x)\operatorname{d\!}x\leq\int_{t-1}^{t}\int_{\Omega}\rho(\tau,x)\operatorname{d\!}xd\tau\leq\int_{\Omega}\rho(t-1,x)\varphi(t-1,x)\operatorname{d\!}x,

from which we get that there are α,β>0\alpha,\beta>0 such that

Ωρ(t,x)dxβeαt.\int_{\Omega}\rho(t,x)\operatorname{d\!}x\leq\beta e^{-\alpha t}.

Now, let us denote u(t):=Ωρ2(t,x)dxu(t):=\int_{\Omega}\rho^{2}(t,x)\operatorname{d\!}x. This is well defined for all t>0t>0. We have

u(t)=2νΩ|ρ|22ΩρVρ,u^{\prime}(t)=-2\nu\int_{\Omega}|\nabla\rho|^{2}-2\int_{\Omega}\rho V\cdot\nabla\rho,

and, using Young’s inequality, we get that there is δ>0\delta>0 (depending on supκ\sup\kappa and ν\nu) such that

u2δu0.u^{\prime}-2\delta u\leq 0.

Hence

Ωρ2(t,x)dx(Ωρ2(1,x)dx)e2δ(t1).\int_{\Omega}\rho^{2}(t,x)\operatorname{d\!}x\leq\left(\int_{\Omega}\rho^{2}(1,x)\operatorname{d\!}x\right)e^{2\delta(t-1)}.

Now, let θ(0,1)\theta\in(0,1) be close enough to 11 so that αθ>δ(1θ)\alpha\theta>\delta(1-\theta). Let pθ:=θ+2(1θ)>1p_{\theta}:=\theta+2(1-\theta)>1. By classical interpolation arguments on LpL^{p} spaces, one has

ρ(t,)Lpθρ(t,)L1θρ(t,)L21θAe(αθδ(1θ))t,\lVert\rho(t,\cdot)\rVert_{L^{p_{\theta}}}\leq\lVert\rho(t,\cdot)\rVert_{L^{1}}^{\theta}\lVert\rho(t,\cdot)\rVert_{L^{2}}^{1-\theta}\leq Ae^{-(\alpha\theta-\delta(1-\theta))t},

where A=βθe(1θ)δρ(1,)L21θA=\beta^{\theta}e^{-(1-\theta)\delta}\lVert\rho(1,\cdot)\rVert_{L^{2}}^{1-\theta}. Now that we have that the LpθL^{p_{\theta}} norm of ρ(t,)\rho(t,\cdot) goes to zero as tt goes to ++\infty, Corollary A.2 gives us that the LL^{\infty} norm of ρ(t,)\rho(t,\cdot) also goes to zero when tt goes to ++\infty.

Step 3. Long-time behavior of φ\varphi.

We now turn to the convergence of φ\varphi as t+t\to+\infty. Let (tn)n(t_{n})_{n\in\mathbb{N}} be a sequence of positive real numbers diverging to ++\infty. Define

φn(t,x):=φ(t+tn,x).\varphi_{n}(t,x):=\varphi(t+t_{n},x).

Then, φn\varphi_{n} solves

tφnνΔφn+κ(ρ(t+tn,x))|φn|1=0,t>tn,xΩ.-\partial_{t}\varphi_{n}-\nu\Delta\varphi_{n}+\kappa(\rho(t+t_{n},x))|\nabla\varphi_{n}|-1=0,\quad t>-t_{n},\ x\in\Omega.

Using the same estimates as in the first step, we find that, up to a subsequence, φn\varphi_{n} converges to some φ¯(t,x)\overline{\varphi}(t,x) in the Lloc2(H1)L_{loc}^{2}(H^{1}) sense, that satisfies

tφ¯νΔφ¯+κ(0)|φ¯|1=0,t,xΩ,-\partial_{t}\overline{\varphi}-\nu\Delta\overline{\varphi}+\kappa(0)|\nabla\overline{\varphi}|-1=0,\quad t\in\mathbb{R},\ x\in\Omega,

where we have used the uniform convergence ρ(t,)0\rho(t,\cdot)\to 0 as t+t\to+\infty from the previous step in order to get the convergence of κ(ρ(t+tn,x))\kappa(\rho(t+t_{n},x)) to κ(0)\kappa(0) as n+n\to+\infty. We now want to prove φ¯=Ψ\overline{\varphi}=\Psi. From the boundedness of φ\varphi, the function φ¯\overline{\varphi} is also bounded.

Let T>0T>0 be fixed. Let uT,vTu_{T},v_{T} be the solutions of

(4.7) tuνΔu+κ(0)|u|1=0,t(0,T),xΩ,-\partial_{t}u-\nu\Delta u+\kappa(0)|\nabla u|-1=0,\quad t\in(0,T),\ x\in\Omega,

with homogeneous Dirichlet boundary conditions and final data uT(T,)=0u_{T}(T,\cdot)=0 and vT(T,)=Φ|Ω+Mv_{T}(T,\cdot)=\Phi|_{\Omega}+M, where Mφ¯M\geq\overline{\varphi} and Φ|Ω0\Phi|_{\Omega}\geq 0 is the restriction to Ω\Omega of the solution of the torsion equation νΔΦ=1-\nu\Delta\Phi=1 in Ω+\Omega^{+} (with Ω+\Omega^{+} a domain that contains Ω\Omega, say Ω+:=Ω+B1\Omega^{+}:=\Omega+B_{1}) with Dirichlet boundary conditions. We recall that the existence of uT,vTu_{T},v_{T} is guaranteed by Proposition 2.5.

The parabolic comparison principle, Proposition 2.6, implies that, for every T>0T>0,

uT(t,)φ¯(t,)vT(t,), for t(0,T).u_{T}(t,\cdot)\leq\overline{\varphi}(t,\cdot)\leq v_{T}(t,\cdot),\quad\text{ for }\ t\in(0,T).

Let us prove that uT,vTu_{T},v_{T} converge to Ψ\Psi, the stationary solution of (4.7), as TT goes to ++\infty. To get this, let us show that the sequences of functions (uT)T>0(u_{T})_{T>0} and (vT)T>0(v_{T})_{T>0} are non-decreasing and non-increasing respectively, in the sense that uTuT+hu_{T}\leq u_{T+h} and vTvT+hv_{T}\geq v_{T+h} on (0,T)×Ω(0,T)\times\Omega for every h(0,T)h\in(0,T).

Let T>0T>0 be fixed and let h(0,T)h\in(0,T). Because (4.7) is autonomous, uT+hu_{T+h} and uTu_{T} are both solutions of (4.7) on (0,T)×Ω(0,T)\times\Omega, with final data uT+h(T,)u_{T+h}(T,\cdot) and uT(T,)=0u_{T}(T,\cdot)=0 respectively. However, because uT+h(t,)0u_{T+h}(t,\cdot)\geq 0 for t(0,T+h)t\in(0,T+h) (as recalled in Proposition 2.5), we have uT+h(T,)uT(T,)u_{T+h}(T,\cdot)\geq u_{T}(T,\cdot). To phrase it differently, uT+hu_{T+h} and uTu_{T} are solutions of the same equation with ordered final data, hence, we can apply the comparison principle Proposition 2.6 to find that uT+huTu_{T+h}\geq u_{T} on (0,T)×Ω(0,T)\times\Omega.

Similarly , we have that vT+hv_{T+h} and vTv_{T} solve (4.7) on (0,T)×Ω(0,T)\times\Omega, with final data vT+h(T,)v_{T+h}(T,\cdot) and vT(T,)=Φ|Ω+Mv_{T}(T,\cdot)=\Phi|_{\Omega}+M. By a standard comparison principle, we have that vT+hΦ|Ω+Mv_{T+h}\leq\Phi|_{\Omega}+M. Therefore, we can apply the parabolic comparison principle Proposition 2.6 to get that vT+hvTv_{T+h}\leq v_{T} on (0,T)×Ω(0,T)\times\Omega.

Therefore, owing to theses monotonicities, the sequences (uT)T>0(u_{T})_{T>0} and (vT)T>0(v_{T})_{T>0} converge a.e. as TT goes to ++\infty to functions that do not depend on the tt variable (this last fact comes from the equality uT(,)=uT+h(+h,)u_{T}(\cdot,\cdot)=u_{T+h}(\cdot+h,\cdot), which is true because (4.7) is autonomous and because the solutions are unique). Moreover, arguing as in the first step, we have that these limiting functions are solutions of (4.7). The only stationary solution of (4.7) being Ψ\Psi, we get that φ¯(t,)=Ψ\overline{\varphi}(t,\cdot)=\Psi for every tt. We have thus proven that

wn(t,x):=φ(t+tn,x)Ψ(x)n+0,w_{n}(t,x):=\varphi(t+t_{n},x)-\Psi(x)\underset{n\to+\infty}{\longrightarrow}0,

in the Lloc2(H1)L^{2}_{loc}(H^{1}) sense.

Let us prove that this convergence is actually uniform. To this aim, observe that wnw_{n} is a weak solution of

twn=νΔwnκ(ρ(+tn,))znwn+(κ(0)κ(ρ(+tn,)))|Ψ|,-\partial_{t}w_{n}=\nu\Delta w_{n}-\kappa(\rho(\cdot+t_{n},\cdot))z_{n}\cdot\nabla w_{n}+(\kappa(0)-\kappa(\rho(\cdot+t_{n},\cdot)))|\nabla\Psi|,

where zn:=φn+Ψ|φn|+|Ψ|z_{n}:=\frac{\nabla\varphi_{n}+\nabla\Psi}{|\nabla\varphi_{n}|+|\nabla\Psi|} is bounded. Then, for every t1,t2t_{1},t_{2} such that t2+1<t1<t2+2t_{2}+1<t_{1}<t_{2}+2, using Corollary A.3, we find that

wn(t2,)LC(wn(t1,)L2+(κ(0)κ(ρ(+tn,)))|Ψ|L((t1,t2)×Ω))).\lVert w_{n}(t_{2},\cdot)\rVert_{L^{\infty}}\leq C\left(\lVert w_{n}(t_{1},\cdot)\rVert_{L^{2}}+\lVert(\kappa(0)-\kappa(\rho(\cdot+t_{n},\cdot)))\lvert\nabla\Psi\rvert\rVert_{L^{\infty}((t_{1},t_{2})\times\Omega))}\right).

Integrating this for t1(t2+1,t2+2)t_{1}\in(t_{2}+1,t_{2}+2), we find

wn(t2,)LC(wnL2((t2+1,t2+2)×Ω)+(κ(0)κ(ρ(+tn,)))|Ψ|L((t1,t2)×Ω))).\lVert w_{n}(t_{2},\cdot)\rVert_{L^{\infty}}\leq C\left(\lVert w_{n}\rVert_{L^{2}((t_{2}+1,t_{2}+2)\times\Omega)}+\lVert(\kappa(0)-\kappa(\rho(\cdot+t_{n},\cdot)))\lvert\nabla\Psi\rvert\rVert_{L^{\infty}((t_{1},t_{2})\times\Omega))}\right).

Because wnw_{n} goes to zero in the Lloc2(H1)L_{loc}^{2}(H^{1}) sense and |Ψ||\nabla\Psi| is bounded, observing that |κ(ρ(+tn,))κ(0)||\kappa(\rho(\cdot+t_{n},\cdot))-\kappa(0)| converges uniformly to zero (this comes from the uniform convergence to zero of ρ\rho from Step 2), we obtain that wnw_{n} goes to zero uniformly, whence

φ(t,x)t+Ψ(x)\varphi(t,x)\underset{t\to+\infty}{\longrightarrow}\Psi(x)

in the LL^{\infty} sense. ∎

4.2. Improved convergence results

In the previous section, Theorem 4.2 proved the existence of solutions (ρ,φ)(\rho,\varphi) to the MFG system with infinite time horizon (1.2) and characterized the asymptotic behavior of any such solution by providing uniform convergence ρt0\rho_{t}\to 0 and φtΨ\varphi_{t}\to\Psi. We want here to improve this result in two ways: first, we will prove that this convergence is actually exponential (in what concerns φ\varphi this requires a very small extra assumption on the function κ\kappa); second, we will prove that the convergence of φ(t,)\varphi(t,\cdot) to Ψ\Psi as t+t\to+\infty, in addition to being uniform, is also a strong convergence in H01(Ω)H_{0}^{1}(\Omega). This last result is natural to evoke, because of the role played by φ\nabla\varphi in the dynamics.

Proposition 4.5.

Suppose that the function κ:++\kappa:\mathbb{R}_{+}\to\mathbb{R}_{+} is Hölder continuous. Then, there exist constants C,α>0C,\alpha>0 (depending on κ\kappa, ν\nu, and Ω\Omega), such that we have, for any (t,x)[0,+)×Ω(t,x)\in[0,+\infty)\times\Omega,

|ρ(t,x)|+|φ(t,x)Ψ(x)|Ceαt.\lvert\rho(t,x)\rvert+\lvert\varphi(t,x)-\Psi(x)\rvert\leq Ce^{-\alpha t}.
Proof.

The exponential convergence of ρ\rho to 0 is indeed part of the proof of Theorem 4.2, since we proved that, for pp close to 11, the LpL^{p} norm of ρt\rho_{t} tends exponentially to 0, and we then used the parabolic regularization estimate ρtLCρt1Lp\left\lVert\rho_{t}\right\rVert_{L^{\infty}}\leq C\left\lVert\rho_{t-1}\right\rVert_{L^{p}}.

Thanks to the assumption that κ\kappa is Hölder continuous, up to modifying the coefficient in the exponent, we obtain |K(t,x)κ(0)|Ceαt\lvert K(t,x)-\kappa(0)\rvert\leq Ce^{-\alpha t}, where K(t,x)=κ(ρ(t,x))K(t,x)=\kappa(\rho(t,x)).

We need now to discuss the exponential convergence of φ\varphi. Let us fix a time t1t_{1} and define

a±:=1±3CψLeαt1,Ψ±:=a±Ψ±eαt1.a_{\pm}:=1\pm 3C\left\lVert\nabla\psi\right\rVert_{L^{\infty}}e^{-\alpha t_{1}},\quad\Psi_{\pm}:=a_{\pm}\Psi\pm e^{-\alpha t_{1}}.

We will use a comparison principle between φ\varphi and Ψ±\Psi_{\pm}. The functions Ψ±\Psi_{\pm} solve

tΨ±νΔΨ±+κ(0)|Ψ±|a±=0,-\partial_{t}\Psi_{\pm}-\nu\Delta\Psi_{\pm}+\kappa(0)|\nabla\Psi_{\pm}|-a_{\pm}=0,

where the time-derivative term is actually 0 since they are functions of the xx variable only. If we set v±:=φΨ±v_{\pm}:=\varphi-\Psi_{\pm}, the functions v±v_{\pm} solve a linear PDE of the form

tv±νΔv±+w±v±±3CψLeαt1+(K(t,x)κ(0))a±|Ψ|=0,-\partial_{t}v_{\pm}-\nu\Delta v_{\pm}+w_{\pm}\cdot\nabla v_{\pm}\pm 3C\left\lVert\nabla\psi\right\rVert_{L^{\infty}}e^{-\alpha t_{1}}+(K(t,x)-\kappa(0))a_{\pm}|\nabla\Psi|=0,

where the vector fields w±w_{\pm} are such that |w±(t,x)|K(t,x)|w_{\pm}(t,x)|\leq K(t,x). In particular, if we note that, for t1t_{1} large enough, we have 0a±20\leq a_{\pm}\leq 2, we have (K(t,x)κ(0))a±|Ψ|2CψLeαt1(K(t,x)-\kappa(0))a_{\pm}|\nabla\Psi|\leq 2C\left\lVert\nabla\psi\right\rVert_{L^{\infty}}e^{-\alpha t_{1}}. Hence, for v+v_{+} we have

tv+νΔv++w+v+<0-\partial_{t}v_{+}-\nu\Delta v_{+}+w_{+}\cdot\nabla v_{+}<0

and for vv_{-}

tvνΔv+wv>0.-\partial_{t}v_{-}-\nu\Delta v_{-}+w_{-}\cdot\nabla v_{-}>0.

Let us look now at the boundary conditions of v±v_{\pm} relative to the parabolic domain [t1,t2]×Ω[t_{1},t_{2}]\times\Omega. If t2t_{2} is large enough, using the uniform convergence φtΨ\varphi_{t}\to\Psi, we can infer v+(t2,x)<0v_{+}(t_{2},x)<0 for every xΩx\in\Omega. Moreover, we also have v+(t,x)<0v_{+}(t,x)<0 for every tt and every xΩx\in\partial\Omega. The inequalities are opposite for vv_{-}, i.e. we have v(t,x)>0v_{-}(t,x)>0 for t=t2t=t_{2} or xΩx\in\partial\Omega. This implies, by the maximum principle in [2] (see [2, Theorem 1], adapted to this backward equation, and using again the version with the inequality presented at the end of the proof, page 98), the inequalities v+(t1,x)0v(t1,x)v_{+}(t_{1},x)\leq 0\leq v_{-}(t_{1},x), i.e.

(13CψLeαt1)Ψeαt1φ(1+3CψLeαt1)Ψ+eαt1.(1-3C\left\lVert\nabla\psi\right\rVert_{L^{\infty}}e^{-\alpha t_{1}})\Psi-e^{-\alpha t_{1}}\leq\varphi\leq(1+3C\left\lVert\nabla\psi\right\rVert_{L^{\infty}}e^{-\alpha t_{1}})\Psi+e^{-\alpha t_{1}}.

This shows φtΨLCeαt1\left\lVert\varphi_{t}-\Psi\right\rVert_{L^{\infty}}\leq Ce^{-\alpha t_{1}}, for a new constant CC. ∎

We can now pass to the following statement, which proves the convergence of the gradient of φ\varphi.

Proposition 4.6.

Let (ρ,φ)(\rho,\varphi) be a solution to the Mean Field Game system with infinite time horizon (1.2). Then

φ(t,)t+Ψ\varphi(t,\cdot)\underset{t\to+\infty}{\longrightarrow}\Psi

in H01(Ω)H_{0}^{1}(\Omega).

Proof.

We first observe that, by Lemma 4.3, the family of functions (φ(t,))t0(\varphi(t,\cdot))_{t\geq 0} is bounded in H01(Ω)H_{0}^{1}(\Omega). This, together with the uniform convergence to Ψ\Psi, implies that one has the weak convergence φ(t,)Ψ\varphi(t,\cdot)\xrightharpoonup{}\Psi in H01(Ω)H_{0}^{1}(\Omega) as t+t\to+\infty.

In order to conclude the proof, it suffices to show φ(t,)H01(Ω)ΨH01(Ω)\lVert\varphi(t,\cdot)\rVert_{H_{0}^{1}(\Omega)}\to\lVert\Psi\rVert_{H_{0}^{1}(\Omega)} as t+t\to+\infty. Since φ(t,)Ψ\varphi(t,\cdot)\xrightharpoonup{}\Psi in H01(Ω)H_{0}^{1}(\Omega) as tt\to\infty, one has

ΨH01(Ω)2lim inft+φ(t,)H01(Ω)2.\lVert\Psi\rVert_{H_{0}^{1}(\Omega)}^{2}\leq\liminf_{t\to+\infty}\lVert\varphi(t,\cdot)\rVert_{H_{0}^{1}(\Omega)}^{2}.

Let us argue by contradiction and assume that there exists ε>0\varepsilon>0 and an increasing sequence (sn)n(s_{n})_{n\in\mathbb{N}} with sn+s_{n}\to+\infty as n+n\to+\infty such that

(4.8) φ(sn)H01(Ω)2ΨH01(Ω)2+2ε\left\lVert\varphi(s_{n})\right\rVert_{H_{0}^{1}(\Omega)}^{2}\geq\lVert\Psi\rVert_{H_{0}^{1}(\Omega)}^{2}+2\varepsilon

for every nn\in\mathbb{N}. Recall that, from (4.6) in the proof of Lemma 4.3, there exists C>0C>0 depending only on supφ\sup\varphi, supκ\sup\kappa, ν\nu, and |Ω|\lvert\Omega\rvert such that

φ(t,)H01(Ω)2φ(sn)H01(Ω)2eC(tsn)C(1eC(tsn))\lVert\varphi(t,\cdot)\rVert_{H_{0}^{1}(\Omega)}^{2}\geq\lVert\varphi(s_{n})\rVert_{H_{0}^{1}(\Omega)}^{2}e^{-C(t-s_{n})}-C(1-e^{-C(t-s_{n})})

for every nn\in\mathbb{N} and t(sn,sn+1)t\in(s_{n},s_{n}+1). Combining this with (4.8), one obtains that there exists δ(0,1)\delta\in(0,1) depending only on ΨH01(Ω)\lVert\Psi\rVert_{H_{0}^{1}(\Omega)}, ε\varepsilon, and CC such that

(4.9) φ(t,)H01(Ω)2ΨH01(Ω)2+ε\left\lVert\varphi(t,\cdot)\right\rVert_{H_{0}^{1}(\Omega)}^{2}\geq\lVert\Psi\rVert_{H_{0}^{1}(\Omega)}^{2}+\varepsilon

for every nn\in\mathbb{N} and t(sn,sn+δ)t\in(s_{n},s_{n}+\delta). By Lemma 4.3, there exists a constant C>0C^{\prime}>0 depending only on supφ\sup\varphi, supκ\sup\kappa, ν\nu, |Ω|\lvert\Omega\rvert, and Ψ\Psi such that

φL2((sn,sn+δ);H2(Ω))2C for every n.\lVert\varphi\rVert_{L^{2}((s_{n},s_{n}+\delta);H^{2}(\Omega))}^{2}\leq C^{\prime}\qquad\text{ for every }n\in\mathbb{N}.

In particular, for every nn\in\mathbb{N}, there exists tn(sn,sn+δ)t_{n}\subset(s_{n},s_{n}+\delta) such that φ(tn)H2(Ω)2C/δ\lVert\varphi(t_{n})\rVert_{H^{2}(\Omega)}^{2}\leq C^{\prime}/\delta. Hence (φ(tn))n(\varphi(t_{n}))_{n\in\mathbb{N}} is bounded in H2(Ω)H^{2}(\Omega) and thus, up to extracting subsequences (which we still denote by (sn)n(s_{n})_{n\in\mathbb{N}} and (tn)n(t_{n})_{n\in\mathbb{N}} for simplicity), (φ(tn))n(\varphi(t_{n}))_{n\in\mathbb{N}} converges strongly in H01(Ω)H_{0}^{1}(\Omega) as nn\to\infty. Since φ(t,)Ψ\varphi(t,\cdot)\xrightharpoonup{}\Psi as t+t\to+\infty, the strong limit of (φ(tn))n(\varphi(t_{n}))_{n\in\mathbb{N}} in H01(Ω)H_{0}^{1}(\Omega) is necessarily Ψ\Psi, and thus, in particular, φ(tn)H01(Ω)2ΨH01(Ω)2\lVert\varphi(t_{n})\rVert_{H_{0}^{1}(\Omega)}^{2}\to\lVert\Psi\rVert_{H_{0}^{1}(\Omega)}^{2} as n+n\to+\infty. This, however, contradicts (4.9), and establishes the desired result. ∎

Appendix A Regularizing effects of parabolic equations

This appendix is concerned with the regularizing properties of a class of parabolic equations including both the Fokker–Planck and the Hamilton–Jacobi–Bellman equations we consider in this paper. More precisely, we consider the increase of the exponent pp of the LpL^{p} integrability in space of the solution of the system. As recalled in the introduction, the computations and results presented here are very similar to those from the appendix of [7], the main difference lying in the boundary condition. The main result of this appendix is the following.

Proposition A.1.

Let T(0,+]T\in(0,+\infty]. Let V,F,f,g,uC((0,T)×Ω)V,F,f,g,u\in C^{\infty}((0,T)\times\Omega) with V,gL((0,T)×Ω)V,g\in L^{\infty}((0,T)\times\Omega), u0u\geq 0, u=0u=0 on Ω\partial\Omega, such that

(A.1) tuνΔu+(uV)+F+f+gu0,on(0,T)×Ω.\partial_{t}u-\nu\Delta u+\nabla\cdot(uV)+\nabla\cdot F+f+g\cdot\nabla u\leq 0,\quad\text{on}\quad(0,T)\times\Omega.

Then, for every p>1p>1, every number a(0,1)a\in(0,1) and t1,t2t_{1},t_{2} such that 0<t1<t2<T0<t_{1}<t_{2}<T and a<|t1t2|<a1a<\lvert t_{1}-t_{2}\rvert<a^{-1}, there is C>0C>0, depending only on p,a,VL,gLp,a,\lVert V\rVert_{L^{\infty}},\lVert g\rVert_{L^{\infty}} such that

u(t2,)LC(u(t1,)Lp+FL((t1,t2)×Ω)+fL((t1,t2)×Ω)).\lVert u(t_{2},\cdot)\rVert_{L^{\infty}}\leq C\left(\lVert u(t_{1},\cdot)\rVert_{L^{p}}+\lVert F\rVert_{L^{\infty}((t_{1},t_{2})\times\Omega)}+\lVert f\rVert_{L^{\infty}((t_{1},t_{2})\times\Omega)}\right).

The same result is true omitting the assumption u0u\geq 0 if the PDE (A.1) is satisfied as an equality instead of an inequality.

The proof follows a standard method based on Moser’s iterations that will be detailed here. This appendix is included for completeness: the experienced reader will recognize well-known computations, which are simplified in this setting thanks in particular to the Dirichlet boundary conditions we use.

Proof.

Let uu be as in the proposition. For k>1k>1, we define

mk(t):=Ωuk(t,x)dx.m_{k}(t):=\int_{\Omega}u^{k}(t,x)\operatorname{d\!}x.

We also define α:=22=nn2\alpha:=\frac{2^{\star}}{2}=\frac{n}{n-2} if n>2n>2 (here 22^{\star} is the Sobolev exponent in dimension nn). When n=1,2n=1,2 we set α:=2\alpha:=2 (but any number larger than 11 and smaller than ++\infty could be used here). Moreover, we set

M:=FL((t1,t2)×Ω)+fL((t1,t2)×Ω)M:=\lVert F\rVert_{L^{\infty}((t_{1},t_{2})\times\Omega)}+\lVert f\rVert_{L^{\infty}((t_{1},t_{2})\times\Omega)}

Step 1. LpL^{p} estimates.

Let us start with proving that, for k0>1k_{0}>1, there is C>0C>0 depending on k0k_{0} and on the LL^{\infty} norms of V,gV,g, such that, for every k>k0>1k>k_{0}>1,

(A.2) ddt(mkeCk2t)+1Cmαk1αeCk2tCeCk2tk2Mk.\frac{\operatorname{d\!}}{\operatorname{d\!}t}(m_{k}e^{-Ck^{2}t})+\frac{1}{C}m_{\alpha k}^{\frac{1}{\alpha}}e^{-Ck^{2}t}\leq Ce^{-Ck^{2}t}k^{2}M^{k}.

In order to do so, we differentiate mkm_{k} with respect to tt, to get

mk(t)kΩ(νΔu(uV)guFf)uk1k(k1)νΩ|u|2uk2+k(k1)Ω(Vu)uk1kΩ(gu)uk1+k(k1)Ω(Fu)uk2kΩfuk1.m_{k}^{\prime}(t)\leq k\int_{\Omega}(\nu\Delta u-\nabla\cdot(uV)-g\cdot\nabla u-\nabla\cdot F-f)u^{k-1}\\ \leq-k(k-1)\nu\int_{\Omega}|\nabla u|^{2}u^{k-2}+k(k-1)\int_{\Omega}(V\cdot\nabla u)u^{k-1}-k\int_{\Omega}(g\cdot\nabla u)u^{k-1}\\ +k(k-1)\int_{\Omega}(F\cdot\nabla u)u^{k-2}-k\int_{\Omega}fu^{k-1}.

Now, owing to Young’s inequality, we can find C1,C2,C3>0C_{1},C_{2},C_{3}>0 depending only on VL\lVert V\rVert_{L^{\infty}}, gL\lVert g\rVert_{L^{\infty}}, k0k_{0}, ν\nu such that

mk(t)C1k2Ω|u|2uk2+C2k2Ωuk+C3k2Ω|F|2uk2+kΩ|f|uk1m_{k}^{\prime}(t)\leq-C_{1}k^{2}\int_{\Omega}|\nabla u|^{2}u^{k-2}+C_{2}k^{2}\int_{\Omega}u^{k}+C_{3}k^{2}\int_{\Omega}|F|^{2}u^{k-2}+k\int_{\Omega}|f|u^{k-1}

(note that we replaced the coefficient k(k1)k(k-1) with k2k^{2}, as these two numbers are equivalent up to multiplicative constants as far as k>k0>1k>k_{0}>1). Moreover, thanks again to a Young inequality, we have

|F|2uk22k|F|k+k2kuk and |f|uk11k|f|k+k1kuk.|F|^{2}u^{k-2}\leq\frac{2}{k}|F|^{k}+\frac{k-2}{k}u^{k}\quad\text{ and }\quad|f|u^{k-1}\leq\frac{1}{k}|f|^{k}+\frac{k-1}{k}u^{k}.

Therefore, up to increasing C2,C3C_{2},C_{3}, we get

mk(t)C1k2Ω|u|2uk2+C2k2Ωuk+C3k2Ω|F|k+kΩ|f|kC1k2Ω|u|2uk2+C2k2Ωuk+C3k2Mk.m_{k}^{\prime}(t)\leq-C_{1}k^{2}\int_{\Omega}|\nabla u|^{2}u^{k-2}+C_{2}k^{2}\int_{\Omega}u^{k}+C_{3}k^{2}\int_{\Omega}|F|^{k}+k\int_{\Omega}|f|^{k}\\ \leq-C_{1}k^{2}\int_{\Omega}|\nabla u|^{2}u^{k-2}+C_{2}k^{2}\int_{\Omega}u^{k}+C_{3}k^{2}M^{k}.

Now, owing to the Gagliardo–Nirenberg–Sobolev inequality, we have, for some C4>0C_{4}>0,

k2Ω|u|2uk2=4Ω|(uk2)|2C4(Ωukα)1α=C4mkα1α.k^{2}\int_{\Omega}|\nabla u|^{2}u^{k-2}=4\int_{\Omega}|\nabla(u^{\frac{k}{2}})|^{2}\geq C_{4}\left(\int_{\Omega}u^{k\alpha}\right)^{\frac{1}{\alpha}}=C_{4}m_{k\alpha}^{\frac{1}{\alpha}}.

Hence

mk(t)+C1C4mαk1αC2k2mk+C3k2Mk.m_{k}^{\prime}(t)+C_{1}C_{4}m_{\alpha k}^{\frac{1}{\alpha}}\leq C_{2}k^{2}m_{k}+C_{3}k^{2}M^{k}.

Let us denote C:=max{1C1C4,C2,C3}C:=\max\{\frac{1}{C_{1}C_{4}},C_{2},C_{3}\}. Then, the above equation gives us

mk(t)Ck2mk+1Cmαk1αCk2Mk,m_{k}^{\prime}(t)-Ck^{2}m_{k}+\frac{1}{C}m_{\alpha k}^{\frac{1}{\alpha}}\leq Ck^{2}M^{k},

which we can rewrite in order to get (A.2).

Step 2. Estimates on mαkm_{\alpha k}.

We show in this step that, for k>k0>1k>k_{0}>1 and for 0<t1<t2<T0<t_{1}<t_{2}<T, we have

(A.3) mαk1α(t2)eCαk2(t2t1)1t2t1eCk2t2t1t2mαk1α(s)eCk2sds+eCαk2(t2t1)Mk,m_{\alpha k}^{\frac{1}{\alpha}}(t_{2})\leq e^{C\alpha k^{2}(t_{2}-t_{1})}\frac{1}{t_{2}-t_{1}}e^{Ck^{2}t_{2}}\int_{t_{1}}^{t_{2}}m_{\alpha k}^{\frac{1}{\alpha}}(s)e^{-Ck^{2}s}\operatorname{d\!}s+e^{C\alpha k^{2}(t_{2}-t_{1})}M^{k},

for some CC depending on k0k_{0} and on the LL^{\infty} norms of V,gV,g.

The relation (A.2) provides

ddt(mkeCk2t)CeCk2tk2Mk.\frac{\operatorname{d\!}}{\operatorname{d\!}t}(m_{k}e^{-Ck^{2}t})\leq Ce^{-Ck^{2}t}k^{2}M^{k}.

Let us take s(t1,t2)s\in(t_{1},t_{2}). We integrate the above inequality for t(s,t2)t\in(s,t_{2}) to get:

mk(t2)eCk2t2mk(s)eCk2s+CMkst2eCk2tdtmk(s)eCk2s+MkeCk2s.m_{k}(t_{2})e^{-Ck^{2}t_{2}}\leq m_{k}(s)e^{-Ck^{2}s}+CM^{k}\int_{s}^{t_{2}}e^{-Ck^{2}t}\operatorname{d\!}t\leq m_{k}(s)e^{-Ck^{2}s}+M^{k}e^{-Ck^{2}s}.

Taking the power 1α<1\frac{1}{\alpha}<1 and using its subadditivity yields

mk1α(t2)eCαk2t2mk1α(s)eCαk2s+MkαeCαk2s.m_{k}^{\frac{1}{\alpha}}(t_{2})e^{-\frac{C}{\alpha}k^{2}t_{2}}\leq m_{k}^{\frac{1}{\alpha}}(s)e^{-\frac{C}{\alpha}k^{2}s}+M^{\frac{k}{\alpha}}e^{-\frac{C}{\alpha}k^{2}s}.

We replace kk by αk\alpha k so as to re-write the above inequality as

mαk1α(t2)eCαk2t2mαk1α(s)eCαk2s+MkeCαk2s.m_{\alpha k}^{\frac{1}{\alpha}}(t_{2})e^{-C\alpha k^{2}t_{2}}\leq m_{\alpha k}^{\frac{1}{\alpha}}(s)e^{-C\alpha k^{2}s}+M^{k}e^{-C\alpha k^{2}s}.

We multiply by eCk2s(α1)e^{Ck^{2}s(\alpha-1)} and integrate this for s(t1,t2)s\in(t_{1},t_{2}) in order to obtain

mαk1α(t2)t1t2eCαk2(st2)eCk2sdst1t2mαk1α(s)eCk2sds+Mkt1t2eCk2sds.m_{\alpha k}^{\frac{1}{\alpha}}(t_{2})\int_{t_{1}}^{t_{2}}e^{C\alpha k^{2}(s-t_{2})}e^{-Ck^{2}s}\operatorname{d\!}s\leq\int_{t_{1}}^{t_{2}}m_{\alpha k}^{\frac{1}{\alpha}}(s)e^{-Ck^{2}s}\operatorname{d\!}s+M^{k}\int_{t_{1}}^{t_{2}}e^{-Ck^{2}s}\operatorname{d\!}s.

We then use eCαk2(st2)eCαk2(t1t2)e^{C\alpha k^{2}(s-t_{2})}\geq e^{C\alpha k^{2}(t_{1}-t_{2})} in order to obtain

mαk1α(t2)eCαk2(t2t1)(t1t2eCk2sds)1t1t2mαk1α(s)eCk2sds+eCαk2(t2t1)Mkm_{\alpha k}^{\frac{1}{\alpha}}(t_{2})\leq e^{C\alpha k^{2}(t_{2}-t_{1})}\left(\int_{t_{1}}^{t_{2}}e^{-Ck^{2}s}\operatorname{d\!}s\right)^{-1}\int_{t_{1}}^{t_{2}}m_{\alpha k}^{\frac{1}{\alpha}}(s)e^{-Ck^{2}s}\operatorname{d\!}s+e^{C\alpha k^{2}(t_{2}-t_{1})}M^{k}

and finally we use t1t2eCk2sds(t2t1)eCk2t2\int_{t_{1}}^{t_{2}}e^{-Ck^{2}s}\operatorname{d\!}s\geq(t_{2}-t_{1})e^{-Ck^{2}t_{2}}, which provides the desired inequality.

Step 3. Higher integrability estimates.

Let us now show that, for k>k0k>k_{0}, there is C>0C>0 (depending on the same quantities as in the previous steps), such that

(A.4) mαk1αk(t2)eCk(t2t1)(C1|t2t1|)1/k(mk(t1)+Mk)1k.m_{\alpha k}^{\frac{1}{\alpha k}}(t_{2})\leq\frac{e^{Ck(t_{2}-t_{1})}}{(C^{-1}|t_{2}-t_{1}|)^{1/k}}\left(m_{k}(t_{1})+M^{k}\right)^{\frac{1}{k}}.

First of all, integrating (A.2) for t(t1,t2)t\in(t_{1},t_{2}), and discharging the final value mk(t2)eCt2m_{k}(t_{2})e^{-Ct_{2}}, we obtain

1Ct1t2mαk1α(t)eCk2tdtmk(t1)eCk2t1+Mkt1t2Ck2eCk2tdteCk2t1(Mk+mk(t1)).\frac{1}{C}\int_{t_{1}}^{t_{2}}m_{\alpha k}^{\frac{1}{\alpha}}(t)e^{-Ck^{2}t}\operatorname{d\!}t\leq m_{k}(t_{1})e^{-Ck^{2}t_{1}}+M^{k}\int_{t_{1}}^{t_{2}}Ck^{2}e^{-Ck^{2}t}\operatorname{d\!}t\leq e^{-Ck^{2}t_{1}}\left(M^{k}+m_{k}(t_{1})\right).

Combining this with (A.3), we get

mαk1α(t2)eCαk2(t2t1)(CeCk2(t2t1)t2t1(mk(t1)+Mk)+Mk).m_{\alpha k}^{\frac{1}{\alpha}}(t_{2})\leq e^{C\alpha k^{2}(t_{2}-t_{1})}\left(C\frac{e^{Ck^{2}(t_{2}-t_{1})}}{t_{2}-t_{1}}(m_{k}(t_{1})+M^{k})+M^{k}\right).

Up to enlarging the constant CC and using 0<t2t1<a10<t_{2}-t_{1}<a^{-1}, we can write the above inequality in a simpler form, i.e.

mαk1α(t2)eC(α+1)k2(t2t1)C1(t2t1)(mk(t1)+Mk),m_{\alpha k}^{\frac{1}{\alpha}}(t_{2})\leq\frac{e^{C(\alpha+1)k^{2}(t_{2}-t_{1})}}{C^{-1}(t_{2}-t_{1})}\left(m_{k}(t_{1})+M^{k}\right),

hence, (A.4) holds true, after raising to the power 1/k1/k and including α+1\alpha+1 in the constant CC.

Step 4. Iterations.

We conclude the proof in this step by proving that, for p,t1,t2p,t_{1},t_{2} as in the statement of the proposition, there is C>0C>0 such that

(A.5) u(t2,)LC(u(t1,)Lp+FL+fL).\lVert u(t_{2},\cdot)\rVert_{L^{\infty}}\leq C(\lVert u(t_{1},\cdot)\rVert_{L^{p}}+\lVert F\rVert_{L^{\infty}}+\lVert f\rVert_{L^{\infty}}).

We denote

sn:=t2t2t1(2α)n,kn:=αnp,βn:=eCkn(sn+1sn)(C1(sn+1sn))1kn,s_{n}:=t_{2}-\frac{t_{2}-t_{1}}{(2\alpha)^{n}},\ k_{n}:=\alpha^{n}p,\ \beta_{n}:=\frac{e^{Ck_{n}(s_{n+1}-s_{n})}}{(C^{-1}(s_{n+1}-s_{n}))^{\frac{1}{k_{n}}}},

and

an:=mkn1kn(sn),a~n:=max{an,M}.a_{n}:=m_{k_{n}}^{\frac{1}{k_{n}}}(s_{n}),\quad\tilde{a}_{n}:=\max\{a_{n},M\}.

Then, (A.4) gives us that

an+1βn(ankn+Mkn)1knβn21kna~n.a_{n+1}\leq\beta_{n}(a_{n}^{k_{n}}+M^{k_{n}})^{\frac{1}{k_{n}}}\leq\beta_{n}2^{\frac{1}{k_{n}}}\tilde{a}_{n}.

Hence, up to replacing the constant CC with a larger one so as to suppose βn21kn1\beta_{n}2^{\frac{1}{k_{n}}}\geq 1, we find

a~n+1βn21kna~n.\tilde{a}_{n+1}\leq\beta_{n}2^{\frac{1}{k_{n}}}\tilde{a}_{n}.

We observe that we have n=0+βn21kn<+\prod_{n=0}^{+\infty}\beta_{n}2^{\frac{1}{k_{n}}}<+\infty as a consequence of the logarithmic estimate

n=0+log(βn21kn)n=0+Cknt2t1(2α)n+1kn(log2+nlog(2α)log(t2t1)+logC)<+.\sum_{n=0}^{+\infty}\log(\beta_{n}2^{\frac{1}{k_{n}}})\leq\sum_{n=0}^{+\infty}Ck_{n}\frac{t_{2}-t_{1}}{(2\alpha)^{n}}+\frac{1}{k_{n}}(\log 2+n\log(2\alpha)-\log(t_{2}-t_{1})+\log C)<+\infty.

Therefore, we obtain

max{limn+an,M}(n=0+βn21kn)max{a0,M}C(a0+M).\max\{\lim_{n\to+\infty}a_{n},M\}\leq\left(\prod_{n=0}^{+\infty}\beta_{n}2^{\frac{1}{k_{n}}}\right)\max\{a_{0},M\}\leq C(a_{0}+M).

Hence, thanks to limn+an=u(t2,)L\lim_{n\to+\infty}a_{n}=\lVert u(t_{2},\cdot)\rVert_{L^{\infty}}, we obtain (A.5). This concludes the proof. ∎

Corollary A.2.

Let T(0,+]T\in(0,+\infty]. Let VL((0,T)×Ω)V\in L^{\infty}((0,T)\times\Omega). Let uL1((0,T)×Ω)u\in L^{1}((0,T)\times\Omega) be a positive distributional solution of

tuνΔu(uV)0, on (0,T)×Ω,\partial_{t}u-\nu\Delta u-\nabla\cdot(uV)\leq 0,\quad\text{ on }\ (0,T)\times\Omega,

satisfying the following mild regularity assumption: uu is obtained as a measurable curve (ut)t(u_{t})_{t} of functions of the xx variable, which is such that tΩη(x)ut(x)dxt\mapsto\int_{\Omega}\eta(x)u_{t}(x)\operatorname{d\!}x is continuous in time for every ηC(Ω)\eta\in C^{\infty}(\Omega) (note that we do not restrict to ηCc(Ω)\eta\in C^{\infty}_{c}(\Omega)). Then, for every p>1p>1 and a(0,1)a\in(0,1), there is C>0C>0, depending only on pp, aa, VL\lVert V\rVert_{L^{\infty}}, such that we have

u(t2,)LCu(t1,)Lp\lVert u(t_{2},\cdot)\rVert_{L^{\infty}}\leq C\lVert u(t_{1},\cdot)\rVert_{L^{p}}

for every 0<t1<t2<T0<t_{1}<t_{2}<T with a<|t2t1|<a1a<\lvert t_{2}-t_{1}\rvert<a^{-1}.

Proof.

To prove this estimate the only important point is to regularize the equation so as to apply Proposition A.1. In order for the proof to be self-contained, we detail a two-step approximation argument.

We convolve the equation by an approximation of the identity and to apply Proposition A.1. However, convolving will not preserve the Dirichlet boundary conditions, so we first have to extend uu by zero on a bigger set.

We define Ω+\Omega^{+} to be a open bounded regular set such that Ω+B1Ω+\Omega+B_{1}\subset\Omega^{+}, where B1B_{1} is unit ball in N\mathbb{R}^{N}.

We define u+(t,x):=u(t,x)u^{+}(t,x):=u(t,x) if xΩx\in\Omega, and u+(t,x)=0u^{+}(t,x)=0 elsewhere. Let ηε(x)\eta_{\varepsilon}(x) be an approximation of the identity whose support is included in B1B_{1}. We define uε:=u+ηεu_{\varepsilon}:=u^{+}\star\eta_{\varepsilon} (here, \star is the convolution in space only). It is a function which is smooth in xx and continuous in tt. We then convolve in time as well, taking χδ(t)\chi_{\delta}(t) an approximation of the identity whose support is included in +\mathbb{R}_{+}. Defining uε,δ:=χδuεu_{\varepsilon,\delta}:=\chi_{\delta}\star u_{\varepsilon} we have now a function which is smooth in time and space. It satisfies, in the classical sense,

tuε,δνΔuε,δ(uε,δVε,δ)0, for t(0,T),xΩ+,\partial_{t}u_{\varepsilon,\delta}-\nu\Delta u_{\varepsilon,\delta}-\nabla\cdot(u_{\varepsilon,\delta}V_{\varepsilon,\delta})\leq 0,\quad\text{ for }\ t\in(0,T),\ x\in\Omega^{+},

with Vε,δ:=χδηε(uV)uε,δCV_{\varepsilon,\delta}:=\frac{\chi_{\delta}\star\eta_{\varepsilon}\star(uV)}{u_{\varepsilon,\delta}}\in C^{\infty}. Moreover, the LL^{\infty} norm of Vε,δV_{\varepsilon,\delta} is bounded independently of ε\varepsilon and δ\delta. Then, uε,δu_{\varepsilon,\delta} is positive, regular and is a (classical) subsolution of a Fokker–Planck equation with regular coefficients, hence we can apply Proposition A.1. We then take the limit δ0\delta\to 0, and we observe that we have

uε(t,)Lp=limδ0uε,δ(t,)Lp\lVert u_{\varepsilon}(t,\cdot)\rVert_{L^{p}}=\lim_{\delta\to 0}\lVert u_{\varepsilon,\delta}(t,\cdot)\rVert_{L^{p}}

for every tt, since uεu_{\varepsilon} is continuous. Then, we have

u(t,)Lp=limε0uε(t,)Lp\lVert u(t,\cdot)\rVert_{L^{p}}=\lim_{\varepsilon\to 0}\lVert u_{\varepsilon}(t,\cdot)\rVert_{L^{p}}

from standard properties of convolutions (with the possibility, of course, that this limit and this norm take the value ++\infty). ∎

Corollary A.3.

Let T(0,+]T\in(0,+\infty]. Let f,gLf,g\in L^{\infty} and let uL((0,T);L2(Ω))L2((0,T);H01(Ω))u\in L^{\infty}((0,T);L^{2}(\Omega))\cap L^{2}((0,T);H^{1}_{0}(\Omega)) be solution (in the weak sense) of

tuνΔu+f+gu=0,on(0,T)×Ω,\partial_{t}u-\nu\Delta u+f+g\cdot\nabla u=0,\quad\text{on}\quad(0,T)\times\Omega,

with Dirichlet boundary conditions and initial datum u(0,)=u0L2u(0,\cdot)=u_{0}\in L^{2}.

Then, for every p>1p>1, and a(0,1)a\in(0,1) there is C>0C>0, depending only on p,a,gLp,a,\lVert g\rVert_{L^{\infty}} such that

u(t2,)LC(u(t1,)Lp+fL)\lVert u(t_{2},\cdot)\rVert_{L^{\infty}}\leq C\left(\lVert u(t_{1},\cdot)\rVert_{L^{p}}+\lVert f\rVert_{L^{\infty}}\right)

for every t1<t2t_{1}<t_{2} with a<|t2t1|<a1a<\lvert t_{2}-t_{1}\rvert<a^{-1}.

Proof.

Let fn,gnf_{n},g_{n} be CC^{\infty} and such that fnff_{n}\to f and gngg_{n}\to g in the L2L^{2} norm. Assume moreover that we have fnLfL\lVert f_{n}\rVert_{L^{\infty}}\to\lVert f\rVert_{L^{\infty}} and gnLgL\lVert g_{n}\rVert_{L^{\infty}}\to\lVert g\rVert_{L^{\infty}}. Let unu_{n} be the solution of

tunνΔun+fn+gnun=0,on(0,+)×Ω,\partial_{t}u_{n}-\nu\Delta u_{n}+f_{n}+g_{n}\cdot\nabla u_{n}=0,\quad\text{on}\quad(0,+\infty)\times\Omega,

with Dirichlet boundary condition and with initial datum u0nu_{0}^{n}, where u0nu_{0}^{n} is a smooth L2L^{2} approximation of u0u_{0}.

Then, unu_{n} is smooth enough to apply Proposition A.1 to unu_{n}, to get, for p,t1,t2p,t_{1},t_{2} as in the statement of the corollary,

(A.6) un(t2,)LC(un(t1,)Lp+fnL).\lVert u_{n}(t_{2},\cdot)\rVert_{L^{\infty}}\leq C\left(\lVert u_{n}(t_{1},\cdot)\rVert_{L^{p}}+\lVert f_{n}\rVert_{L^{\infty}}\right).

Then, as nn goes to ++\infty, unu_{n} converges (the arguments to prove this are standard and based on the weak L2L^{2} convergence of un\nabla u_{n}) to a solution (in the weak sense) of

tuνΔu+f+gu=0,on(0,+)×Ω,\partial_{t}u-\nu\Delta u+f+g\cdot\nabla u=0,\quad\text{on}\quad(0,+\infty)\times\Omega,

with Dirichlet boundary conditions and with initial datum u0u_{0}. The convergence is also strong in the L2L^{2} sense. Because this solution is unique, it necessarily coincides with the original solution uu of the statement. In order to obtain the result, we need to pass to the limit the inequality (A.6). The left-hand side can easily be dealt with by semicontinuity, while for the right-hand side, we suppose p2p\leq 2 and we use strong L2L^{2} convergence. Since this convergence is L2L^{2} in space-time, we have convergence of the right-hand side only for a.e. t1t_{1}. Yet, using the fact that the solution uu is continuous in time as a function valued into L2(Ω)L^{2}(\Omega), the result extends to all t1t_{1}. The inequality for p=2p=2 implies that with p>2p>2, up to modifying the constant in a way depending on |Ω|\lvert\Omega\rvert. ∎

The reader may observe that we used different regularization strategies to prove the two above corollaries. Indeed, the linear behavior of the Fokker–Planck equation allowed to directly regularize the solution (up to modifying the drift vector field: we convolve the solution and define a new drift vector field which preserves the same LL^{\infty} bound, a trick which is completely standard for curves in the Wasserstein space, see for instance [31, Chapter 5]). This is not possible for the Hamilton–Jacobi–Bellman equation. However, when uniqueness of the solution is known, regularizing the coefficients and the data of the equation is another option, and it is what we did in our last corollary.


Acknowledgments. The authors wish to thank many colleagues for useful discussions and suggestions, and in particular Alessio Porretta. Without the comments he made after a talk the second author gave on the topic of the present paper, the strategy to achieve convergence to a solution in the limit T+T\to+\infty would have been completely different, the result less general, and the time needed to achieve it much longer.

The authors acknowledge the financial support of French ANR project “MFG”, reference ANR-16-CE40-0015-01, and of a public grant as part of the “Investissement d’avenir” project, reference ANR-11-LABX-0056-LMH, LabEx LMH, PGMO project VarPDEMFG. The first author was also partially supported by the by the French IDEXLYON project Impulsion “Optimal Transport and Congestion Games” PFI 19IA106udl and the second author was also partially supported by the Hadamard Mathematics LabEx (LMH) through the grant number ANR-11-LABX-0056-LMH in the “Investissement d’avenir” project.

References

  • [1] Y. Achdou and A. Porretta. Mean field games with congestion. Ann. Inst. H. Poincaré Anal. Non Linéaire, 35(2):443–480, 2018.
  • [2] D. G. Aronson and J. Serrin. Local behavior of solutions of quasilinear parabolic equations. Arch. Rational Mech. Anal., 25:81–122, 1967.
  • [3] J.-P. Aubin and H. Frankowska. Set-valued analysis. Modern Birkhäuser Classics. Birkhäuser Boston, Inc., Boston, MA, 2009. Reprint of the 1990 edition.
  • [4] J.-D. Benamou, G. Carlier, and F. Santambrogio. Variational mean field games. In Active particles. Vol. 1. Advances in theory, models, and applications, Model. Simul. Sci. Eng. Technol., pages 141–171. Birkhäuser/Springer, Cham, 2017.
  • [5] P. Cannarsa and C. Sinestrari. Semiconcave functions, Hamilton–Jacobi equations, and optimal control. Progress in Nonlinear Differential Equations and their Applications, 58. Birkhäuser Boston, Inc., Boston, MA, 2004.
  • [6] P. Cardaliaguet. Notes on mean field games (from P.-L. Lions’ lectures at Collège de France). Available at https://www.ceremade.dauphine.fr/~cardaliaguet/MFG20130420.pdf, 2013.
  • [7] P. Cardaliaguet, J.-M. Lasry, P.-L. Lions, and A. Porretta. Long time average of mean field games with a nonlocal coupling. SIAM J. Control Optim., 51(5):3558–3591, 2013.
  • [8] L. C. Evans. Partial differential equations, volume 19 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 1998.
  • [9] F. Feo. A remark on uniqueness of weak solutions for some classes of parabolic problems. Ric. Mat., 63(1, suppl.):S143–S155, 2014.
  • [10] W. H. Fleming and H. M. Soner. Controlled Markov processes and viscosity solutions, volume 25 of Stochastic Modelling and Applied Probability. Springer, New York, second edition, 2006.
  • [11] D. Gilbarg and N. S. Trudinger. Elliptic partial differential equations of second order. Classics in Mathematics. Springer-Verlag, Berlin, 2001. Reprint of the 1998 edition.
  • [12] D. A. Gomes and V. K. Voskanyan. Short-time existence of solutions for mean-field games with congestion. Journal of the London Mathematical Society, 92(3):778–799, 11 2015.
  • [13] A. Granas and J. Dugundji. Fixed point theory. Springer Monographs in Mathematics. Springer-Verlag, New York, 2003.
  • [14] M. Huang, P. E. Caines, and R. P. Malhamé. Individual and mass behaviour in large population stochastic wireless power control problems: centralized and Nash equilibrium solutions. In 42nd IEEE Conference on Decision and Control, 2003. Proceedings, volume 1, pages 98–103. IEEE, 2003.
  • [15] M. Huang, P. E. Caines, and R. P. Malhamé. Large-population cost-coupled LQG problems with nonuniform agents: individual-mass behavior and decentralized ϵ\epsilon-Nash equilibria. IEEE Trans. Automat. Control, 52(9):1560–1571, 2007.
  • [16] M. Huang, R. P. Malhamé, and P. E. Caines. Large population stochastic dynamic games: closed-loop McKean–Vlasov systems and the Nash certainty equivalence principle. Commun. Inf. Syst., 6(3):221–251, 2006.
  • [17] R. L. Hughes. A continuum theory for the flow of pedestrians. Transportation Research Part B: Methodological, 36(6):507–535, jul 2002.
  • [18] R. L. Hughes. The flow of human crowds. In Annual review of fluid mechanics, Vol. 35, volume 35 of Annu. Rev. Fluid Mech., pages 169–182. Annual Reviews, Palo Alto, CA, 2003.
  • [19] O. A. Ladyženskaja, V. A. Solonnikov, and N. N. Ural\cprimeceva. Linear and quasilinear equations of parabolic type. Translations of Mathematical Monographs, Vol. 23. American Mathematical Society, Providence, R.I., 1968. Translated from the Russian by S. Smith.
  • [20] O. A. Ladyzhenskaya and N. N. Ural’tseva. Linear and quasilinear elliptic equations. Academic Press, New York-London, 1968. Translated from the Russian by Scripta Technica, Inc, Translation editor: Leon Ehrenpreis.
  • [21] J.-M. Lasry and P.-L. Lions. Jeux à champ moyen. I. Le cas stationnaire. C. R. Math. Acad. Sci. Paris, 343(9):619–625, 2006.
  • [22] J.-M. Lasry and P.-L. Lions. Jeux à champ moyen. II. Horizon fini et contrôle optimal. C. R. Math. Acad. Sci. Paris, 343(10):679–684, 2006.
  • [23] J.-M. Lasry and P.-L. Lions. Mean field games. Jpn. J. Math., 2(1):229–260, 2007.
  • [24] G. M. Lieberman. Second order parabolic differential equations. World Scientific Publishing Co., Inc., River Edge, NJ, 1996.
  • [25] P.-L. Lions. Résolution de problèmes elliptiques quasilinéaires. Arch. Rational Mech. Anal., 74(4):335–353, 1980.
  • [26] P.-L. Lions. Courses at Collège de France, 2006–2012. http://www.college-de-france.fr/site/pierre-louis-lions/_course.htm.
  • [27] G. Mazanti and F. Santambrogio. Minimal-time mean field games. Math. Models Methods Appl. Sci., 29(8):1413–1464, 2019.
  • [28] A. Porretta. Weak solutions to Fokker-Planck equations and mean field games. Arch. Ration. Mech. Anal., 216(1):1–62, 2015.
  • [29] M. M. Porzio. Existence of solutions for some “noncoercive” parabolic equations. Discrete Contin. Dynam. Systems, 5(3):553–568, 1999.
  • [30] F. Santambrogio. Lecture notes on variational mean field games. Preprint cvgmt. http://cvgmt.sns.it/paper/4646/.
  • [31] F. Santambrogio. Optimal transport for applied mathematicians, volume 87 of Progress in Nonlinear Differential Equations and their Applications. Birkhäuser/Springer, Cham, 2015. Calculus of variations, PDEs, and modeling.
  • [32] J. Simon. Compact sets in the space Lp(0,T;B)L^{p}(0,T;B). Ann. Mat. Pura Appl. (4), 146:65–96, 1987.