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

The ergodic Mean Field Game system for a type of state constraint condition

Mariya Sardarli

1 Introduction

The object of this paper is to establish well-posedness (existence and uniqueness) results for a type of state constraint ergodic Mean Field Game (MFG) system

{Δu+H(Du)+ρ=F(x;m) in Ω,Δm+div(mDpH(Du))=0 in Ω,m0,Ωm=1,\displaystyle\begin{cases}-\Delta u+H(Du)+\rho=F(x;m)&\text{ in }\Omega,\\ \Delta m+div(mD_{p}H(Du))=0&\text{ in }\Omega,\\ m\geq 0,\quad\int_{\Omega}m=1,\end{cases} (1.1)

supplemented with the state constraint-type infinite Dirichlet boundary condition

limd(x)0u(x)=,\lim\limits_{d(x)\to 0}u(x)=\infty, (1.2)

where d(x):=d(x,Ω)d(x):=d(x,\partial\Omega) is the distance to the boundary. Here Ωn\Omega\subset\mathbb{R}^{n} is an open bounded subset of n\mathbb{R}^{n}, H:n×Ω¯H:\mathbb{R}^{n}\times\overline{\Omega}\to\mathbb{R} is the Hamiltonian associated with the cost function of an individual agent, ρ\rho is the ergodic constant, mm is the distribution of the agents, and F:Ω×L1(Ω)F:\Omega\times L^{1}(\Omega)\to\mathbb{R} is the interaction term. The Hamilton-Jacobi-Bellman (HJB) equation is understood in the viscosity sense and the Kolmogorov-Fokker-Planck (KFP) equation, in the distributional sense.

The MFG (1.1)(\ref{eq:proto}) with the boundary condition (1.2)(\ref{intro:bds}) can be interpreted as a state constraint-type problem. Indeed, the infinite boundary condition prevents the underlying stochastic trajectories from reaching the boundary, forcing them to stay in the domain.

The theory of MFGs was introduced by Lasry and Lions [8] and, in a particular setting, by Caines, Huang, and Malhamé [4], to describe the interactions of a large number of small and indistinguishable rational agents. In the absence of common noise, the behavior of the agents leads to a forward-backward coupled system of PDEs, consisting of a backward HJB equation describing the individual agent’s value function, and a forward KFP equation for the distribution of the law (density) of the population.

The forward-backward system with periodic boundary conditions and either local or non-local coupling, as well its ergodic stationary counterpart, was first studied in [8]. Summaries of results may also be found in Cardaliguet [2], Lions [9], and Gomes, Pimentel, and Voskanyan [3]. Fewer results exist, however, in the case of either Dirichlet, state constraint or Neumann boundary conditions.

We next state the main result of this paper for superlinear power-like Hamiltonians, that is,

H(p,x)=|p|q with q>1,H(p,x)=|p|^{q}\text{ with }\quad q>1, (1.3)

and suitable assumptions on the coupling FF (see (F1)(F1), (F2)(F2) and (F3)(F3) in Section 2.2). We emphasize that the particular form of the Hamiltonian is by no means essential to the analysis that follows. It simply provides a sufficiently general model problem that results in the necessary asymptotics of uu and mm near Ω\partial\Omega. In particular, for 1<q21<q\leq 2, Hamiltonians H(p)H(p) for which

δq/(q1)H(δ1/(q1)p)|p|q0,\delta^{q/(q-1)}H(\delta^{-1/(q-1)}p)-|p|^{q}\to 0, (1.4)

locally uniformly in pp as δ0\delta\to 0 are permissible. We state the result below, deferring the precise meaning of a solution to the system until Section 3.

Theorem.

Let 1<q21<q\leq 2 and assume either (F1)(F1) or (F2)(F2).
Then, for all r>1r>1, there exists a solution (u,ρ,m)Wloc2,r(Ω)××Wloc1,r(Ω)(u,\rho,m)\in W^{2,r}_{loc}(\Omega)\times\mathbb{R}\times W^{1,r}_{loc}(\Omega) to the system

{Δu+|Du|q+ρ=F(x;m) in Ω,limd(x)0u(x)=,Δm+div(mq|Du|q2Du)=0 in Ω,m0,Ωm=1.\displaystyle\begin{cases}-\Delta u+|Du|^{q}+\rho=F(x;m)&\text{ in }\Omega,\\ \lim\limits_{d(x)\to 0}u(x)=\infty,\\ \Delta m+div(mq|Du|^{q-2}Du)=0&\text{ in }\Omega,\\ m\geq 0,\quad\int_{\Omega}m=1.\end{cases} (1.5)

Moreover, if FF satisfies (F3)(F3), then the solution is unique.

As will be explained below, the conditions (F1)(F1) and (F2)(F2) describe non-local and local couplings respectively, while (F3)(F3) is the usual monotonicity condition.

1.1 Background

Heuristically, HJB equations with convex (in the momentum) Hamiltonian HH are associated with the stochastic control problem governed by the dynamics

dXt=αtdt+2dBt,dX_{t}=\alpha_{t}dt+\sqrt{2}dB_{t},

where (αt)t0(\alpha_{t})_{t\geq 0} is a non-anticipating process. In feedback form, the optimal policy is DpH(Du)-D_{p}H(Du).

In settings in which the trajectories of the agents reach the boundary but do not leave the domain or, as is the case we consider here, do not reach the boundary at all, the solutions of the HJB equation are referred to as state constraint solutions.

If the drift DpH(Du)-D_{p}H(Du) is bounded, then for all t>0t>0, P(XtΩc)>0P(X_{t}\in\Omega^{c})>0. It follows that for trajectories to remain inside Ω\Omega, the drift must become unbounded near the boundary, pushing the agents back into the domain. For Hamiltonians of the form (1.3), or more generally (1.4)(\ref{eq:gen}), state constraint solutions exist when the equation is paired with an infinite Neumann condition (reflection cost) or infinite Dirichlet condition (exit cost), the “correct” choice of boundary condition depending on whether the Hamiltonian is sub- or super-quadratic. Indeed, when q>2q>2, solutions are bounded but the normal component of the drift blows up at the boundary. In the sub-quadratic case, the solution itself also blows up.

In the sub-quadratic case, the trajectories never reach the boundary. Hence, the population density is “almost zero” near Ω\partial\Omega. As the behavior of mm at the boundary is coupled with the blowup of DuDu on Ω\partial\Omega, no boundary condition is required on mm in (1.5).

A type of MFG state-constraint problem has been studied by Porretta and Ricciardi in [10]. The authors impose structural assumptions on their Hamiltonian so that, for all choices of the control, an agent governed by these dynamics does not exit the domain Ω\Omega. In this instance, no explicit boundary conditions on the value function are necessary to achieve well posedness. However, the assumptions of [10] do not permit coercive Hamiltonians, and, in particular, do not apply to the ones considered in this paper.

1.2 Infinite Dirichlet boundary conditions

In the context of HJB equations, infinite boundary conditions with power-like Hamiltonians were studied by Lasry and Lions in [7], who considered the HJB equation

Δu+|Du|q+λu=g in Ω,-\Delta u+|Du|^{q}+\lambda u=g\text{ in }\Omega, (1.6)

with 1<q21<q\leq 2 and gL(Ω)g\in L^{\infty}(\Omega), as well as the ergodic problem

Δu+|Du|q+ρ=g in Ω,-\Delta u+|Du|^{q}+\rho=g\text{ in }\Omega,

both coupled with the infinite Dirichlet boundary condition (1.2)(\ref{intro:bds}).

It was shown in [7] that the boundary value problem

{Δu+|Du|q+ρ=g in Ω,limd(x)0u(x)=,\displaystyle\begin{cases}-\Delta u+|Du|^{q}+\rho=g\text{ in }\Omega,\\ \lim\limits_{d(x)\to 0}u(x)=\infty,\end{cases} (1.7)

has a unique viscosity solution (u,ρ)(u,\rho), where uniqueness for uu is understood to be modulo an additive constant. Such a solution has unbounded drift in the direction of the boundary and the agents remain in the domain.

Due to the coupled nature of the MFG system, the existence and well posedness of the corresponding KFP equation depends in an essential way on the asymptotics of uu and DuDu near the boundary.

In general, if bLloc1(Ω)b\in L^{1}_{loc}(\Omega), the distributional solution mm of

Δm+div(mb)=0 in Ω,Ωm=1,m0,\displaystyle\Delta m+div(mb)=0\text{ in }\Omega,\quad\int_{\Omega}m=1,\quad m\geq 0, (1.8)

need not be unique unless a boundary condition is specified.

On the other hand, if

limd(x)0b(x)d(x)=Cν,\displaystyle\lim\limits_{d(x)\to 0}b(x)d(x)=C\nu, (1.9)

where C>0C>0 and ν\nu is the outward unit normal to the boundary at the point closest to xx, Huang, Ji, Liu and Yi showed in [5] that solutions are unique, despite the absence of boundary conditions.

To show (1.9)(\ref{eq:13}) for b=DpH(Du)b=D_{p}H(Du) and apply this result, we need to make precise the asymptotics of uu and DuDu near the boundary.

In [7], the authors proved that if (u,ρ)(u,\rho) is a solution of (1.7)(\ref{eq:hjb}), the asymptotics are precisely of this type. Namely, there exists a constant C>0C>0 such that

limd(x)0Du(x)ν(x)d1/(q1)=C.\displaystyle\lim\limits_{d(x)\to 0}\dfrac{Du(x)\cdot\nu(x)}{d^{-1/(q-1)}}=C.

In this paper, we refine the result and show that the rate of convergence of the above limit is controlled by the LL^{\infty}-norm of gg. This becomes important when proving stability results related to (1.5)(\ref{eq:mainintro}).

1.3 The methodology

As usual, the existence of a solution to the MFG system (1.5)(\ref{eq:mainintro}) is proven via a fixed point argument. To apply such an argument, it is crucial to obtain bounds on mm up to Ω\partial\Omega. Toward that end, for δ>0\delta>0, we define the subdomain ΩδΩ\Omega_{\delta}\subset\Omega by

Ωδ:={xΩ:d(x,Ω)>δ}.\Omega_{\delta}:=\{x\in\Omega:d(x,\partial\Omega)>\delta\}.

As a first step, we show that if bb satisfies (1.9)(\ref{eq:13}) with constant CC, then there exists γ=γ(C)>0\gamma=\gamma(C)>0, δ=δ(C)>0\delta=\delta(C)>0, and L=L(C,bL(Ωδ))>0L=L(C,||b||_{L^{\infty}(\Omega_{\delta})})>0 such that any solution mm of (1.8)(\ref{eq:mbla}) satisfies

ΩdγmL.\int_{\Omega}d^{-\gamma}m\leq L. (1.10)

This estimate is needed in both the local and non-local cases.

The space in which the fixed point argument is carried out differs in the non-local and local coupling settings. In the former, the fixed point is obtained in the space

𝒲γ(Ω):={mL1(Ω):Ωdγm<},\mathcal{W}_{\gamma}(\Omega):=\{m\in L^{1}(\Omega):\int_{\Omega}d^{-\gamma}m<\infty\}, (1.11)

where γ\gamma is as in (1.10)(\ref{eq:gam|}) and, here, depends only on qq. The map T:𝒲γ(Ω)𝒲γ(Ω)T:\mathcal{W}_{\gamma}(\Omega)\to\mathcal{W}_{\gamma}(\Omega) formed by composing the solution maps of (1.7)(\ref{eq:hjb}) and (1.8(\ref{eq:mbla}) satisfies the conditions of Schaefer’s fixed point theorem. In particular, the continuity of TT is straightforward due to the regularizing effect of the coupling.

By contrast, the continuity of the solution map is far from obvious in the local case, where F(x;m)=f(m(x))F(x;m)=f(m(x)) for a continuous and bounded function ff. Convergence of a sequence mnm_{n} in 𝒲γ(Ω)\mathcal{W}_{\gamma}(\Omega) does not yield local uniform convergence of F(x;mn)F(x;m_{n}) as it does in the non-local case. Thus the fixed point argument must be performed in a different space.

To deal with this issue, we introduce, for δ>0\delta>0, the approximating coupled ergodic system

{Δuδ+|Duδ|q+ρδ=f(m~δ(x)), in Ω,limd(x)0uδ(x)=,Δmδ+div(mδq|Duδ|q2Duδ)=0 in Ωδ,(Dmδ+mδq|Duδ|q2Duδ)ν=0 on Ωδ,Ωmδ=1,\displaystyle\begin{cases}-\Delta u_{\delta}+|Du_{\delta}|^{q}+\rho_{\delta}=f(\tilde{m}_{\delta}(x)),&\text{ in }\Omega,\\ \lim\limits_{d(x)\to 0}u_{\delta}(x)=\infty,\\ \Delta m_{\delta}+div\left(m_{\delta}q|Du_{\delta}|^{q-2}Du_{\delta}\right)=0&\text{ in }\Omega_{\delta},\\ \left(Dm_{\delta}+m_{\delta}q|Du_{\delta}|^{q-2}Du_{\delta}\right)\cdot\nu=0&\text{ on }\partial\Omega_{\delta},\\ \int_{\Omega}m_{\delta}=1,\end{cases} (1.12)

where (uδ,ρδ,mδ)W1,r××C0,α(Ωδ)(u_{\delta},\rho_{\delta},m_{\delta})\in W^{1,r}\times\mathbb{R}\times C^{0,\alpha}(\Omega_{\delta}) and m~δ\tilde{m}_{\delta} is a suitable extension of mδ{m_{\delta}} to all of Ω\Omega satisfying

m~δC0,α(Ω)=mδC0,α(Ωδ).||\tilde{m}_{\delta}||_{C^{0,\alpha}(\Omega)}=||m_{\delta}||_{C^{0,\alpha}(\Omega_{\delta})}.

We note that (1.12) is not the standard MFG system with Neumann boundary conditions, in which the HJB and KFP equations are set in the same domain and both have Neumann boundary conditions. Instead, the HJB equation is set in the entire domain Ω\Omega, allowing us to to take advantage of the known asymptotics of uu and DuDu near the boundary. On the other hand, the Neumann conditions on the KFP equation allow us to exploit in regularity of mδm_{\delta} up to Ωδ\partial\Omega_{\delta}.

As we discuss below, for bL(Ω)b\in L^{\infty}(\Omega), the KFP equation

{Δmδ+div(mδb)=0 in Ωδ,(Dmδ+mδb)ν=0 on Ωδ,\displaystyle\begin{cases}\Delta m_{\delta}+div(m_{\delta}b)=0&\text{ in }\Omega_{\delta},\\ (Dm_{\delta}+m_{\delta}b)\cdot\nu=0&\text{ on }\partial\Omega_{\delta},\end{cases}

has a a unique (modulo a multiplicative constant) distributional solution, which is positive and Hölder continuous up to the boundary with Hölder constant depending on the LL^{\infty}-norm of the drift. If uu is the solution of a HJB equation like (1.7)(\ref{eq:hjb}), DuDu is bounded on Ωδ\Omega_{\delta} and this result applies to the KFP equation in (1.12)(\ref{eq:deltap}). Moreover, as shown in [7], the local bounds on the solutions uδu_{\delta} of the HJB equation are uniform in the LL^{\infty}-norm of ff, which is independent of mδm_{\delta}. Therefore, as we show below, it is possible to carry out a fixed point argument in C0,α(Ωδ)C^{0,\alpha}(\Omega_{\delta}).

We are then able to pass from the solution (uδ,ρδ,mδ)(u_{\delta},\rho_{\delta},m_{\delta}) of the approximating system to a solution (u,ρ,m)(u,\rho,m) of the original system on Ω\Omega by the stability of the HJB equation and the local uniform bounds on mδm_{\delta}, which in turn follow from local uniform bounds on DuδDu_{\delta}.

Lastly, under the usual monotonicity assumption we establish uniqueness in both the local and non-local case. The key observation in the argument is that, for solutions (ρ,u,m)(\rho,u,m) of (1.5), mm decays near the boundary as in (1.10).

1.4 Organization of the paper

The paper is organized as follows. The assumptions on the domain Ω\Omega and the coupling FF are stated in Section 2. Section 3 introduces the notions of weak solutions of the KFP and HJB equations and proves (local) regularity and stability results. Section 4 treats the non-local coupling case. In Section 5, we show the approximating problem (1.12) has a solution and pass to the limit to obtain a solution to (1.5)(\ref{eq:mainintro}) on the entire domain. Section 6 establishes the uniqueness of the system and quantifies the sense in which mm vanishes at the boundary. The Appendix contains two technical lemmata used in Sections 3 and 5.

2 Assumptions and definitions

2.1 The domain

Throughout this paper it is assumed that

Ω is a bounded open subset of n with C2-boundary.\Omega\text{ is a bounded open subset of }\mathbb{R}^{n}\text{ with }C^{2}\text{-boundary}.

In particular, the domain satisfies the uniform interior ball condition. Namely, there exists a δ0>0\delta_{0}>0 such that for every xΩx\in\partial\Omega, the ball Bδ0(xδ0ν(x))B_{\delta_{0}}(x-\delta_{0}\nu(x)) is contained in Ω\Omega.

Recall the definition of the subdomains

Ωδ:={xΩ:d(x,Ω)>δ}.\Omega_{\delta}:=\{x\in\Omega:d(x,\partial\Omega)>\delta\}.

Due to the C2C^{2}-regularity of Ω\Omega, there exists ϵ0(0,1)\epsilon_{0}\in(0,1) such that d(x,Ω)C2(Ω\Ωϵ0)d(x,\partial\Omega)\in C^{2}(\Omega\backslash\Omega_{\epsilon_{0}}) and for all xΩ\Ωϵ0x\in\Omega\backslash\Omega_{\epsilon_{0}} there exists a unique point x¯Ω\overline{x}\in\partial\Omega such that d(x,Ω)=|xx¯|d(x,\partial\Omega)=|x-\overline{x}|. Moreover, on Ω\Ωϵ0\Omega\backslash\Omega_{\epsilon_{0}},

Dd(x,Ω)=Dd(x¯,Ω)=ν(x¯).Dd(x,\partial\Omega)=Dd(\overline{x},\partial\Omega)=-\nu(\overline{x}).

Lastly, we define dd to be a C2(Ω¯)C^{2}(\overline{\Omega}) extension of d(x,Ω)d(x,\partial\Omega) which is equal to d(x,Ω)d(x,\partial\Omega) on Ω\Ωϵ0\Omega\backslash\Omega_{\epsilon_{0}}. We additionally require that d(x)1d(x)\leq 1 in Ω\Omega and d(x)M0d(x)\geq M_{0} in Ωϵ0\Omega_{\epsilon_{0}} for some positive constant M0M_{0}.

2.2 The coupling

Our analysis permits a coupling F:Ω×L1(Ω)F:\Omega\times L^{1}(\Omega)\to\mathbb{R} which is either non-local, on the one hand, or local, continuous, and bounded on the other. Namely, we consider couplings FF satisfying one of the following:

  1. (F1)

    The map mF(;m)m\mapsto F(\cdot\ ;m) sends bounded sets in L1(Ω)L^{1}(\Omega) to bounded sets in L(Ω)L^{\infty}(\Omega), and is continuous from L1(Ω)L^{1}(\Omega) into L(Ω)L^{\infty}(\Omega).

  2. (F2)

    F(x;m)=f(m(x))F(x;m)=f(m(x)) for ff a continuous, bounded function.

A coupling satisfying assumption (F1)(F1) has an L(Ω)L^{\infty}(\Omega)-bound depending only on Ωm\int_{\Omega}m. A coupling satisfying (F2)(F2) inherits regularity from the regularity of mm and is bounded by fL()||f||_{L^{\infty}(\mathbb{R})}.

When proving uniqueness we will require the additional monotonicity assumption

  1. (F3)

    Ω(F(x;m1)F(x;m2))(m1(x)m2(x))𝑑x0\int_{\Omega}(F(x;m_{1})-F(x;m_{2}))(m_{1}(x)-m_{2}(x))dx\geq 0, with equality only if m1=m2m_{1}=m_{2} a.e.

2.3 The space of measures for the non-local setting

In the introduction, we defined a space 𝒲γ(Ω)\mathcal{W}_{\gamma}(\Omega) which quantifies the boundary behavior of solutions to KFP equations with an unbounded drift term. We repeat the definition here for completeness.

Definition 2.1.

For γ>0\gamma>0, define a norm ||||γ,Ω||\cdot||_{\gamma,\Omega} by

mγ,Ω=Ωdγ|m|,\displaystyle||m||_{\gamma,\Omega}=\int_{\Omega}d^{-\gamma}|m|,

and recall that 𝒲γ(Ω)\mathcal{W}_{\gamma}(\Omega) is the Banach space induced by this norm, as in (1.11)(\ref{wgamma}).

2.4 The extension of measures defined on subdomains

In the local coupling setting discussed in Section 5 we will consider measures defined on the subdomain Ωδ\Omega_{\delta}. Since FF is defined on Ω×L1(Ω)\Omega\times L^{1}(\Omega), it will be necessary to define a suitable extension of measures on Ωδ\Omega_{\delta} to Ω\Omega that preserves their Hölder regularity.

For μC0,α(Ωδ)\mu\in C^{0,\alpha}(\Omega_{\delta}) and xΩx\in\Omega we define μ~\tilde{\mu} by

μ~(x):=infyΩδ{μ(y)+||μ||C0,α(Ωδ)|xy|α}.\displaystyle\tilde{\mu}(x):=\inf_{y\in\Omega_{\delta}}\{\mu(y)+||\mu||_{C^{0,\alpha}(\Omega_{\delta})}|x-y|^{\alpha}\}. (2.1)

It is straightforward to check that μ~(x)=μ(x)\tilde{\mu}(x)=\mu(x) on Ωδ\Omega_{\delta} and that μ~C0,α(Ω)\tilde{\mu}\in C^{0,\alpha}(\Omega) with the same Hölder modulus as μ\mu.

In the sections that follow, we abuse notation slightly and define the coupling F:Ω×C0,α(Ωδ)F:\Omega\times C^{0,\alpha}(\Omega_{\delta})\to\mathbb{R} by

F(x;μ):=F(x;μ~).\displaystyle F(x;\mu):=F(x;\tilde{\mu}).

3 Preliminaries

In this section, we recall some existence and regularity results for the HJB equation and the KFP equation. At the end we prove stability results that will be used in Section 5.

3.1 The Kolmogorov Fokker-Planck equation

Throughout this paper we will need to consider the KFP equation in subdomains of Ω\Omega. To that end, we study the following KFP boundary value problem for a general open bounded domain VV and BL(V)B\in L^{\infty}(V).

{Δμ+div(μB)=0 in V,(Dμ+μB)ν=0 on V,Vμ=1,μ>0 in V.\displaystyle\begin{cases}\Delta\mu+div(\mu B)=0\ &\text{ in }V,\\ (D\mu+\mu B)\cdot\nu=0&\text{ on }\partial V,\\ \int_{V}\mu=1,\,\mu>0\,\text{ in }V.\end{cases} (3.1)

Later we will apply these results when V=ΩδV=\Omega_{\delta} and B=Dp(H(Du))B=D_{p}(H(Du)).

Definition 3.1.

Given BL(V)B\in L^{\infty}(V), we say that μW1,2(V)\mu\in W^{1,2}(V) is a Neumann weak solution of (3.1)(\ref{mdelta}) if μ\mu is a positive probability measure such that, for all ϕW1,2(V)\phi\in W^{1,2}(V),

V[Dϕ(x)Dμ(x)+B(x)Dϕ(x)μ(x)]𝑑x=0.\displaystyle\int_{V}[D\phi(x)\cdot D\mu(x)+B(x)\cdot D\phi(x)\mu(x)]\ dx=0.

The next lemma gathers existence and uniqueness results for Neumann weak solutions of (3.1)(\ref{mdelta}) as well as Wloc1,r(V)W_{loc}^{1,r}(V) estimates.

Lemma 3.1.

For BL(V)B\in L^{\infty}(V), (3.1)(\ref{mdelta}) admits a unique Neumann weak solution μW1,2(V)\mu\in W^{1,2}(V). In addition, μW1,r(V)\mu\in W^{1,r}(V) for all r>1r>1 and, for compact sets KKVK\subset K^{\prime}\subset V, μW1,r(K)||\mu||_{W^{1,r}(K)} is bounded uniformly by a constant that depends only on K,KK,K^{\prime} and BL(K)||B||_{L^{\infty}(K^{\prime})}.

Proof.

The existence, positivity, and uniqueness are proven in [6]. To obtain the local uniform W1,rW^{1,r} estimate, we recall that Theorem 1.1 of [1] yields a constant C1=C1(K,BL(K))>0C_{1}=C_{1}(K,||B||_{L^{\infty}(K^{\prime})})>0 such that

μW1,r(K)C1μL(K).||\mu||_{W^{1,r}(K)}\leq C_{1}||\mu||_{L^{\infty}(K)}. (3.2)

Then Harnack inequality and the normalization condition Vμ=1\int_{V}\mu=1 give some C2=C2(K,K)>0C_{2}=C_{2}(K,K^{\prime})>0, such that

supKμC2infKμC21|K|.\sup_{K}\mu\leq C_{2}\inf_{K}\mu\leq C_{2}\frac{1}{|K|}. (3.3)

Combining (3.2)(\ref{eq:r}) and (3.3)(\ref{eq:r2}) completes the proof. ∎

If the drift in the KFP equation blows-up at the boundary, it is necessary to modify the notion of weak solution as follows.

Definition 3.2.

Given bLloc1(Ω)b\in L^{1}_{loc}(\Omega) and r>1r>1, we say μWloc1,r(Ω)\mu\in W^{1,r}_{loc}(\Omega) is a weak solution of

Δμ+div(μb)=0 in Ω,\displaystyle\Delta\mu+div(\mu b)=0\text{ in }\Omega, (3.4)

if μ\mu is a non-negative function such that, for all ϕCc2(Ω)\phi\in C^{2}_{c}(\Omega),

Ω[Δϕ(x)b(x)Dϕ(x)]μ(x)𝑑x=0.\displaystyle\int_{\Omega}[\Delta\phi(x)-b(x)\cdot D\phi(x)]\mu(x)\ dx=0.

Moreover, μ\mu is a proper weak solution if Ωμ=1\int_{\Omega}\mu=1.

Remark.

The class of test functions in Definition 3.2 can be extended to C2C^{2} functions that are constant in a neighborhood of Ω\partial\Omega.

The following lemma shows that weak solutions of (3.4)(\ref{eq:m}) may arise as the limit of Neumann weak solutions. The rest of the section provides conditions under which the limit is a proper weak solution.

Lemma 3.2.

Let mδW1,r(Ωδ)m_{\delta}\in W^{1,r}(\Omega_{\delta}) be a Neumann weak solution of (3.1)(\ref{mdelta}) for velocity fields bδL(Ωδ)b_{\delta}\in L^{\infty}(\Omega_{\delta}). Assume that, as δ0\delta\to 0, bδbb_{\delta}\to b locally uniformly. Then, along subsequences, mδmWloc1,r(Ω)m_{\delta}\to m\in W^{1,r}_{loc}(\Omega) and mm is a weak solution of

Δm+div(mb)=0 in Ω.\displaystyle\Delta m+div(mb)=0\text{ in }\Omega.
Proof.

Recalling that mδW1,r(K)||m_{\delta}||_{W^{1,r}(K)} is uniformly bounded for each compact KΩK\subset\Omega, it is possible to extract a subsequence such that mδm_{\delta} converges locally uniformly to a non-negative Wloc1,r(Ω)W^{1,r}_{loc}(\Omega) function mm. To see that mm satisfies the desired equation, consider a test function ϕCc2(Ω)\phi\in C^{2}_{c}(\Omega). Since bδb_{\delta} converges uniformly to bb in the support of ϕ\phi, ΔϕbδDϕ\Delta\phi-b_{\delta}\cdot D\phi converges uniformly to ΔϕbDϕ\Delta\phi-b\cdot D\phi. As mδmm_{\delta}\to m uniformly in the support of ϕ\phi, letting δ0\delta\to 0 in the weak formulation Ωδ(ΔϕbδDϕ)mδ\int_{\Omega_{\delta}}(\Delta\phi-b_{\delta}\cdot D\phi)m_{\delta} leads to the desired result. ∎

It follows from Fatou’s lemma that the limit mm in the above lemma is integrable on Ω\Omega and Ωm1\int_{\Omega}m\leq 1. To ensure that mm is a proper weak solution it is necessary to impose additional structural conditions on the vector fields bδb_{\delta}. In [1], Bogachev and Rökner give conditions on the drifts under which the sequence of weak solutions mδm_{\delta} is uniformly tight, which in turn yields L1(Ω)L^{1}(\Omega) convergence of mδm_{\delta} to mm. The version of the theorem relevant to our setting appears below.

Lemma 3.3 (Lemma 1.1 in [1]).

Let UkU_{k} by a sequence of nested domains and μk\mu_{k} the corresponding sequence of Neumann weak solutions to Δμk+div(μkbk)=0\Delta\mu_{k}+div(\mu_{k}b_{k})=0 on UkU_{k}. Suppose that bkLloc1(Uk)b_{k}\in L^{1}_{loc}(U_{k}) and that, for every R>0R>0, there exists αR>n\alpha_{R}>n such that

κR:=supkUR|bk(x)|αR𝑑x<.\displaystyle\kappa_{R}:=\sup_{k}\int_{U_{R}}|b_{k}(x)|^{\alpha_{R}}dx<\infty.

If, in addition, there exists VC2(U)V\in C^{2}(U) such that, for some ckc_{k}\to\infty, Uk={V<ck}U_{k}=\{V<c_{k}\}, {V=ck}\{V=c_{k}\} has Lebesgue measure zero,

limd(x,U)0V(x)=,\displaystyle\lim_{d(x,\partial U)\to 0}V(x)=\infty, (3.5)

and

limd(x,U)0supk[ΔVDVbk]=,\displaystyle\lim_{d(x,\partial U)\to 0}\sup_{k}[\Delta V-DV\cdot b_{k}]=-\infty, (3.6)

then the sequence μk\mu_{k} is uniformly tight.

A function VV satisfying the above conditions is referred to as a Lyapanov function. The connection between Lyapanov functions and existence and uniqueness of weak solutions of (3.4)(\ref{eq:m}) has been extensively studied in [1], as well as in [5].

It follows from the Lemma 3.3 that, if there exists a VV satisfying (3.5)(\ref{v1}), and

limd(x,U)0ΔVDVb=,\displaystyle\lim_{d(x,\partial U)\to 0}\Delta V-DV\cdot b=-\infty, (3.7)

then, setting bk=bχUkb_{k}=b\chi_{U_{k}}, we obtain a sequence of measures mkm_{k} that converge to a proper weak solution mm of

Δm+div(mb)=0 in Ω.\displaystyle\Delta m+div(mb)=0\text{ in }\Omega. (3.8)

It is proven in [5] that such a solution is unique.

In the present setting, these results will be applied by considering b(x)b(x) satisfying, for some C>1C>1 and ϵ(0,C1)\epsilon\in(0,C-1),

limd(x)0(b(x)Dd(x))d(x)=C,\displaystyle\lim_{d(x)\to 0}(b(x)\cdot Dd(x))d(x)=-C, (3.9)

and V(x)=d(x)C+1+ϵV(x)=d(x)^{-C+1+\epsilon}, which satisfies (3.5)(\ref{v1}) and (3.7)(\ref{vb2}). This is stated precisely in the next result, the proof of which appears in the Appendix.

Lemma 3.4.

Given b:Ωdb:\Omega\to\mathbb{R}^{d} satisfying (3.9)(\ref{eq:asym}) with C>1C>1, there exists a unique proper weak solution mWloc1,r(Ω)m\in W^{1,r}_{loc}(\Omega) of

Δm+div(mb)=0 in Ω,\displaystyle\Delta m+div(mb)=0\text{ in }\Omega,

and, for every ϵ(0,C1)\epsilon\in(0,C-1), there exists a δ=δ(C,ϵ)\delta=\delta(C,\epsilon), such that

ΩdC1+ϵmC^=C^(C,bL(Ωδ)).\displaystyle\int_{\Omega}d^{-C-1+\epsilon}m\leq\hat{C}=\hat{C}(C,||b||_{L^{\infty}(\Omega_{\delta})}).
Remark.

The bound on ΩdC1+ϵm\int_{\Omega}d^{-C-1+\epsilon}m makes precise the intuition that mm “approaches 0” near the boundary Ω\partial\Omega.

It is proven in [7] that, for solutions uu of the HJB equation in (1.5)(\ref{eq:mainintro}), b=q|Du|q2Dub=q|Du|^{q-2}Du satisfies precisely (3.9)(\ref{eq:asym}) for C=q/(q1)C=q/(q-1). This is discussed further in the next section.

3.2 The Hamilton-Jacobi-Bellman equation

In this section, we fix q(1,2]q\in(1,2] and gL(Ω)g\in L^{\infty}(\Omega) and consider the stationary HJB equation with infinite Dirichlet conditions

{Δu+|Du|q+ρ=g in Ω,limd(x)0u(x)=.\displaystyle\begin{cases}-\Delta u+|Du|^{q}+\rho=g\text{ in }\Omega,\\ \lim\limits_{d(x)\to 0}u(x)=\infty.\end{cases} (3.10)

The estimates detailed here will be used to prove the HJB stability results in Theorem 3.8.

We begin by stating the definition of solutions for concreteness.

Definition 3.3.

For 1<q21<q\leq 2 and gL(Ω)g\in L^{\infty}(\Omega), (u,ρ)Wloc2,r(Ω)×(u,\rho)\in W^{2,r}_{loc}(\Omega)\times\mathbb{R} is an explosive solution of (3.10)(\ref{udef}) if uu is a viscosity solution in Ω\Omega and

limd(x)0+u(x)=.\displaystyle\lim_{d(x)\to 0^{+}}u(x)=\infty.

The existence and uniqueness of solutions to such HJB equations was proven in [7], along with several local estimates. These results are collected in the next two theorems. The second theorem is a straightforward consequence of the bounds established in Theorem IV.I of [7], with a slight modification to fit our ergodic framework.

Theorem 3.5.

(Theorem VI.I in [7])) Let 1<q21<q\leq 2 and gL(Ω)g\in L^{\infty}(\Omega). The equation (3.10)(\ref{udef}) admits an explosive solution (u,ρ)Wloc2,r(Ω)×(u,\rho)\in W^{2,r}_{loc}(\Omega)\times\mathbb{R} for all 1<r<1<r<\infty.

Theorem 3.6.

Let 1<q<21<q<2, gL(Ω)g\in L^{\infty}(\Omega) and (u,ρ)Wloc2,r(Ω)×(u,\rho)\in W^{2,r}_{loc}(\Omega)\times\mathbb{R} be an explosive solution of (3.10)(\ref{eq:hj}) with u(x0)=0u(x_{0})=0, for some fixed x0Ωx_{0}\in\Omega . There exist positive constants C1,C2,C3C_{1},C_{2},C_{3}, depending only on Ω\Omega and gL(Ω)||g||_{L^{\infty}(\Omega)}, such that,

  1. (i)

    |ρ|C1|\rho|\leq C_{1},

  2. (ii)

    |Du(x)|C2d(x)1/(q1)|Du(x)|\leq C_{2}d(x)^{-1/(q-1)}, and

  3. (iii)

    |u(x)|C3d(x)(q2)/(q1)|u(x)|\leq C_{3}d(x)^{(q-2)/(q-1)}.

In addition, for any KKK\subset K^{\prime} compact subsets of Ω\Omega, there exists a constant C4=C4(d(K,K),n,r)>0C_{4}=C_{4}(d(K,K^{\prime}),n,r)>0 such that

  1. (iv)

    uW2,r(K)C4(uL(K)+gLq(K))||u||_{W^{2,r}(K)}\leq C_{4}(||u||_{L^{\infty}(K^{\prime})}+||g||_{{L^{q}}(K^{\prime})}).

If q=2q=2, the same estimates hold but (iii)(iii) must be replaced by

  1. (iii)’

    |u(x)|C3log(d(x))|u(x)|\leq C_{3}\log(d(x)).

Sketch of Proof of Theorem 3.6.

The first result follows from the construction of ρ\rho in Theorem VI.I of [7] which shows that ρ\rho arises as the local uniform limit of λuλ\lambda u_{\lambda} as λ0\lambda\to 0, where uλu_{\lambda} is the unique viscosity solution of

{Δuλ+|Duλ|q+λuλ=g in Ω,limd(x)0uλ(x)=.\displaystyle\begin{cases}-\Delta u_{\lambda}+|Du_{\lambda}|^{q}+\lambda u_{\lambda}&=g\text{ in }\Omega,\\ \lim\limits_{d(x)\to 0}u_{\lambda}(x)=\infty.\end{cases}

If 1<q<21<q<2, for arbitrary ϵ>0\epsilon>0, uλu_{\lambda} has as sub- and super-solutions (Cqϵ)d(q2)/(q1)Cϵ/λ(C_{q}\mp\epsilon)d^{(q-2)/(q-1)}\mp C_{\epsilon}/\lambda, where Cq=(q1)(q2)/(q1)(2q)1C_{q}=(q-1)^{(q-2)/(q-1)}(2-q)^{-1} and CϵC_{\epsilon} is a function of gL(Ω)||g||_{L^{\infty}(\Omega)}. Similarly, for q=2q=2, uλu_{\lambda} has as sub- and super-solutions (1ϵ)log(d)Cϵ/λ-(1\mp\epsilon)\log(d)\mp C_{\epsilon}/\lambda, where CϵC_{\epsilon} is a function of gL(Ω)||g||_{L^{\infty}(\Omega)}. Sending λ0\lambda\to 0 yields the desired bound on ρ\rho.

The second result (ii)(ii) is proven in Theorem IV.I of [7] and (iii)(iii) is an immediate consequence of (ii)(ii) as u(x0)u(x_{0}) is fixed. The last interior estimate follows from a bootstrap argument. ∎

In order to study the solution of the KFP equation near the boundary, it will be necessary to know the precise asymptotics of DuDu. In [7], Lasry and Lions, and later Porretta and Véron in [11], established the following results.

Theorem 3.7.

(Theorem I.1 and II.3 in [7]; Theorem 2.3 in [11]) Let 1<q<21<q<2, gL(Ω)g\in L^{\infty}(\Omega) and (u,ρ)Wloc2,r(Ω)×(u,\rho)\in W^{2,r}_{loc}(\Omega)\times\mathbb{R} be a solution of (3.10)(\ref{eq:hj}). Then,

  1. (i)

    limd(x)0u(x)d(q2)/(q1)=(q1)(q2)/(q1)(2q)1\lim\limits_{d(x)\to 0}\dfrac{u(x)}{d^{(q-2)/(q-1)}}=(q-1)^{(q-2)/(q-1)}(2-q)^{-1},

  2. and
  3. (ii)

    limd(x)0Du(x)ν(x)d1/(q1)=(q1)1/(q1).\lim\limits_{d(x)\to 0}\dfrac{Du(x)\cdot\nu(x)}{d^{-1/(q-1)}}=(q-1)^{-1/(q-1)}.

The second result also holds for q=2q=2, while the former must be replaced by

  1. (i)’

    limd(x)0u(x)log(d(x))=1\lim\limits_{d(x)\to 0}\dfrac{u(x)}{-\log(d(x))}=1.

Remark.

Lemma 7.1 in the Appendix is a slightly stronger version of this theorem, where the convergence is shown to be uniform in gL(Ω)||g||_{L^{\infty}(\Omega)}. This stronger version will be necessary in Section 5.

Lastly, we provide a stability result particular to the types of couplings we are studying.

Theorem 3.8.

Assume (F1)(F1) and fix x0Ωx_{0}\in\Omega. Let μn\mu_{n} be a sequence of probability measures on Ω\Omega which converge in L1(Ω)L^{1}(\Omega) to μ\mu, and (un,ρn)Wloc2,r(Ω)×(u_{n},\rho_{n})\in W^{2,r}_{loc}(\Omega)\times\mathbb{R} be the unique explosive solution of

{Δun+|Dun|q+ρn=F(x;μn) in Ω,limd(x)0un(x)=,\displaystyle\begin{cases}-\Delta u_{n}+|Du_{n}|^{q}+\rho_{n}=F(x;\mu_{n})\text{ in }\Omega,\\ \lim\limits_{d(x)\to 0}u_{n}(x)=\infty,\end{cases}

subject to the normalization condition un(x0)=0u_{n}(x_{0})=0. Then there exists (u,ρ)Wloc2,r(Ω)×(u,\rho)\in W^{2,r}_{loc}(\Omega)\times\mathbb{R}, such that ρnρ\rho_{n}\to\rho and unuu_{n}\to u locally uniformly as nn\to\infty. Moreover, (u,ρ)(u,\rho) is the unique explosive solution of

{Δu+|Du|q+ρ=F(x;μ) in Ω,limd(x)0u(x)=,u(x0)=0.\displaystyle\begin{cases}-\Delta u+|Du|^{q}+\rho=F(x;\mu)\text{ in }\Omega,\\ \lim\limits_{d(x)\to 0}u(x)=\infty,u(x_{0})=0.\end{cases} (3.11)
Proof.

It follows from (F1)(F1) that F(x;μn)F(x;\mu_{n}) and F(x;μ)F(x;\mu) are bounded in L(Ω)L^{\infty}(\Omega). Theorem 3.6 yields that ρn\rho_{n} is bounded and hence converges along a subsequence to a constant ρ\rho. The sequence unu_{n} is bounded in Wloc2,r(Ω)W_{loc}^{2,r}(\Omega) so, by Arzelà-Ascoli, we can extract a subsequence converging locally uniformly to uWloc2,r(Ω)u\in W^{2,r}_{loc}(\Omega) such that DunDuDu_{n}\to Du locally uniformly. By pointwise convergence, u(x0)=0u(x_{0})=0.

That u(x)u(x)\to\infty as d(x)0d(x)\to 0 follows from the fact that the sequence of equations have a common explosive subsolution. When 1<q<21<q<2, one such subsolution is of the form (Cqϵ)d(x)(q2)/(q1)Cϵ(C_{q}-\epsilon)d(x)^{(q-2)/(q-1)}-C_{\epsilon}, where Cq=(q1)(q2)/(q1)(2q)1C_{q}=(q-1)^{(q-2)/(q-1)}(2-q)^{-1}, ϵ>0\epsilon>0, and Cϵ=Cϵ(F(x;μ)L(Ω;L1(Ω)),ϵ,Cq)C_{\epsilon}=C_{\epsilon}(||F(x;\mu)||_{L^{\infty}(\Omega;L^{1}(\Omega))},\epsilon,C_{q}). When q=2q=2 the corresponding subsolution is (1ϵ)log(d(x))-(1-\epsilon)\log(d(x)).

Lastly, in view of (F1)(F1), F(x;μn)F(x;\mu_{n}) converges uniformly to F(x;μ)F(x;\mu). It follows from classic viscosity theory that (u,ρ)(u,\rho) is a viscosity solution of (3.11)(\ref{ustab}). As the explosive solution satisfying u(x0)=0u(x_{0})=0 is unique, the limit is indepedent of the subsequence. ∎

3.3 Definition and Main Result

We finish this section with the definition of a solution to the system (1.5)(\ref{eq:mainintro}) and the main existence and uniqueness result stated in the introduction.

Definition 3.4.

A triplet (u,ρ,m)Wloc2,r(Ω)××Wloc1,r(Ω)(u,\rho,m)\in W^{2,r}_{loc}(\Omega)\times\mathbb{R}\times W^{1,r}_{loc}(\Omega) is a solution of (1.5)(\ref{eq:mainintro}) if (u,ρ)(u,\rho) is an explosive solution of

{Δu+|Du|q+ρ=F(x;m) in Ω,limd(x)0u(x)=,\displaystyle\begin{cases}-\Delta u+|Du|^{q}+\rho=F(x;m)\text{ in }\Omega,\\ \lim\limits_{d(x)\to 0}u(x)=\infty,\end{cases}

and mm is a proper weak solution of

Δm+div(mq|Du|q2Du)=0 in Ω.\displaystyle\Delta m+div\left(mq|Du|^{q-2}Du\right)=0\text{ in }\Omega.

We are now able to state the main result of the paper.

Theorem 3.9.

Assume that 1<q21<q\leq 2 and either (F1)(F1) or (F2)(F2) holds. Then (1.5)(\ref{eq:mainintro}) has a solution for all r>1r>1. Moreover, if (F3)(F3) also holds, then the solution is unique.

4 Fixed point for the non-local case

In this section we will use Schaefer’s fixed point theorem in 𝒲γ(Ω)\mathcal{W}_{\gamma}(\Omega) to prove the existence of a solution to (1.5)(\ref{eq:mainintro}) when (F1)(F1) holds.

Let 2<γ<(2q1)/(q1)2<\gamma<(2q-1)/(q-1). Define T1:𝒲γ(Ω)Wloc1,r(Ω)T_{1}:\mathcal{W}_{\gamma}(\Omega)\to W^{1,r}_{loc}(\Omega) to be the map that sends measures μ\mu to q|Du|q2Duq|Du|^{q-2}Du, where (u,ρ)(u,\rho) is any explosive solution of

{Δu+|Du|q+ρ=F(x;μ) in Ω,limd(x)0u(x)=.\displaystyle\begin{cases}-\Delta u+|Du|^{q}+\rho=F(x;{\mu})\text{ in }\Omega,\\ \lim\limits_{d(x)\to 0}u(x)=\infty.\end{cases} (4.1)

It is immediate from Theorem 3.5 that T1T_{1} is well defined as explosive solutions are unique up to a constant.

Define T2:T1(𝒲γ(Ω))𝒲γ(Ω)T_{2}:T_{1}(\mathcal{W}_{\gamma}(\Omega))\to\mathcal{W}_{\gamma}(\Omega) to be the map that sends bb in the image of T1T_{1} to the proper weak solution mm of the associated continuity equation

Δm+div(mb)=0 in Ω.\displaystyle\Delta m+div(mb)=0\text{ in }\Omega. (4.2)

In view of Lemma 3.4 and Theorem 3.7, (4.2)(\ref{eq:fpb}) has a unique proper weak solution mm for bb in the image of T1T_{1}.

Finally define T=T2T1T=T_{2}\circ T_{1}. In the next result, we apply Schaefer’s fixed point theorem to TT, thereby obtaining a solution of the MFG system (1.5)(\ref{eq:mainintro}).

Theorem 4.1.

For FF satisfying (F1)(F1) and 2<γ<(2q1)/(q1)2<\gamma<(2q-1)/(q-1), T:𝒲γ(Ω)𝒲γ(Ω)T:\mathcal{W}_{\gamma}(\Omega)\to\mathcal{W}_{\gamma}(\Omega) has a fixed point.

Proof.

Recall that to apply Schaefer’s fixed point theorem, we need to show that TT is continuous, TT maps bounded sets to precompact sets, and if FWγF\subset W_{\gamma} is the set defined by

F={m𝒲γ(Ω)|m=λT(m) for some 0λ1},\displaystyle F=\{m\in\mathcal{W}_{\gamma}(\Omega)|m=\lambda T(m)\text{ for some }0\leq\lambda\leq 1\}, (4.3)

then FF is bounded.

We start by showing that TT is continuous. Consider a sequence μn\mu_{n} that converges to μ\mu in 𝒲γ(Ω)\mathcal{W}_{\gamma}(\Omega). To simplify the notation, define bn:=T1(μn)b_{n}:=T_{1}(\mu_{n}), b:=T1(μ)b:=T_{1}(\mu), and mn:=T2(bn)m_{n}:=T_{2}(b_{n}). We first show that bnbb_{n}\to b locally uniformly. We then show that the sequence mnm_{n} converges in 𝒲γ(Ω)\mathcal{W}_{\gamma}(\Omega) to a measure mm and prove that m=T2(b)=T(μ)m=T_{2}(b)=T(\mu).

Fix x0Ωx_{0}\in\Omega and let (un,ρn)(u_{n},\rho_{n}) be the unique explosive solution of (4.1)(\ref{eq:hjmu}) that corresponds to F(x;μn)F(x;{\mu}_{n}) and satisfies un(x0)=0u_{n}(x_{0})=0. Note that T1(μn)=q|Dun|q2DunT_{1}(\mu_{n})=q|Du_{n}|^{q-2}Du_{n} by the remark following (4.1)(\ref{eq:hjmu}). Since μn\mu_{n} converges to μ\mu in L1(Ω)L^{1}(\Omega), the local uniform convergence of T1(μn)T_{1}(\mu_{n}) to T1(μ)T_{1}(\mu) follows from Theorem 3.8.

In view of the local uniform bounds on bnb_{n}, Theorem 3.1 implies the sequence of measures mnm_{n} are uniformly bounded in Wloc1,r(Ω)W^{1,r}_{loc}(\Omega) for all r>1r>1 and, hence, locally uniformly bounded in C0,α(Ω)C^{0,\alpha}(\Omega) for α<1n/r\alpha<1-n/r. Along subsequences, the measures mnm_{n} converge locally uniformly to a measure mm. Consequently, we can pass to the limit in the weak formulation of mnm_{n} and obtain that mm is a non-negative measure that satisfies the KFP equation with drift b=T1(μ)b=T_{1}(\mu). It remains to show that mnm_{n} converges to mm in 𝒲γ(Ω)\mathcal{W}_{\gamma}(\Omega).

Lemma 3.4 yields that if (2q1)/(q1)>γ>γ(2q-1)/(q-1)>\gamma^{\prime}>\gamma, then the sequence mnm_{n} is bounded in 𝒲γ(Ω)\mathcal{W}_{\gamma^{\prime}}(\Omega). This implies that the sequence dγmnd^{-\gamma}m_{n} is uniformly tight. Indeed, for some constant CC, independent of the choice of γ\gamma and γ\gamma^{\prime},

C\displaystyle C Ωdγmnδγ+γ(Ω\Ωδdγmn)\displaystyle\geq\int_{\Omega}d^{-\gamma^{\prime}}m_{n}\geq\delta^{-\gamma^{\prime}+\gamma}\left(\int_{\Omega\backslash\Omega_{\delta}}d^{-\gamma}m_{n}\right)

It then follows from the bound

Ωdγ|mnm|Ωδdγ|mnm|+Ω\Ωδdγmn+Ω\Ωδdγm,\displaystyle\int_{\Omega}d^{-\gamma}|m_{n}-m|\leq\int_{\Omega_{\delta}}d^{-\gamma}|m_{n}-m|+\int_{\Omega\backslash\Omega_{\delta}}d^{-\gamma}m_{n}+\int_{\Omega\backslash\Omega_{\delta}}d^{-\gamma}m,

that mnmγ,Ω0||m_{n}-m||_{\gamma,\Omega}\to 0 as nn\to\infty.

Since convergence in 𝒲γ(Ω)\mathcal{W}_{\gamma}(\Omega) implies convergence in L1(Ω)L^{1}(\Omega), this also proves that Ωm=1\int_{\Omega}m=1. The uniqueness of mm then yields the convergence of the full sequence mn=T(μn)m_{n}=T(\mu_{n}) to T(μ)T(\mu).

Next we show that TT maps bounded sets to precompact ones. Indeed let μn\mu_{n} be a bounded sequence in 𝒲γ(Ω)\mathcal{W}_{\gamma}(\Omega), and consider unu_{n}, bnb_{n}, and mnm_{n} defined as before. It suffices to show that there is a convergent subsequence of mnm_{n}. The sequence μn\mu_{n} is bounded in L1(Ω)L^{1}(\Omega) and F(x;μn)F(x;\mu_{n}) is bounded in L(Ω;L1(Ω))L^{\infty}(\Omega;L^{1}(\Omega)). Moreover, Theorem 3.6 yields that bnb_{n} is locally uniformly bounded. Finally, Lemma 3.1 implies that mnm_{n} is uniformly bounded in Cloc0,α(Ω)C^{0,\alpha}_{loc}(\Omega). Passing to subsequences, the measures mnm_{n} converge locally uniformly to a measure mm. Proceeding as in the continuity argument, we find that the sequence dγmnd^{-\gamma}m_{n} is uniformly tight and mnmγ,Ω0||m_{n}-m||_{\gamma,\Omega}\to 0.

It remains to check that the set FF defined above is bounded. To see this, first observe that if m=λT(m)m=\lambda T(m), then Ωm=λ1\int_{\Omega}m=\lambda\leq 1. Hence F(x;m)L(Ω;L1(Ω))||F(x;m)||_{L^{\infty}(\Omega;L^{1}(\Omega))} is bounded by assumption (F1)(F1). Next we apply Theorem 3.6 to find that q|Du|q2Duq|Du|^{q-2}Du is bounded on Ωδ\Omega_{\delta} by a constant C(δ,F(x;m)L(Ω;L1(Ω)))C(\delta,||F(x;m)||_{L^{\infty}(\Omega;L^{1}(\Omega))}). Lemma 3.4 then implies that mγ,Ω||m||_{\gamma,\Omega} is bounded. ∎

5 Fixed point argument for the local setting

In this section we will show existence of solutions to (1.5)(\ref{eq:mainintro}) under assumption (F2)(F2) by considering the system (1.12)(\ref{eq:deltap}), introduced in Section 1, which approximates (1.5)(\ref{eq:mainintro}). First, we will use Schaefer’s fixed point theorem to show the system (1.12)(\ref{eq:deltap}) has a solution. Then, using the uniform tightness results of Lemma 3.3 and the local uniform bounds of Theorem 3.6, we will show that solutions of (1.12)(\ref{eq:deltap}) converge locally uniformly to solutions of (1.5)(\ref{eq:mainintro}).

5.1 The approximate problem

We begin by making precise the notion of solution to the approximating system (1.12)(\ref{eq:deltap}). We then show the existence of a solution using a fixed point theorem argument similar to that used in the non-local setting.

To utilize the asymptotic behavior of the explosive solutions to HJB equations at the boundary of Ω\Omega as well as the Wloc1,rW_{loc}^{1,r} bounds of Neumann weak solutions to KFP equations on Ωδ\Omega_{\delta}, the KFP equation of (1.12)(\ref{eq:deltap}) is solved in the subdomain Ωδ\Omega_{\delta} and the HJB equation is solved in Ω\Omega.

Definition 5.1.

A triplet (uδ,ρδ,mδ)Wloc2,r(Ω)××W1,r(Ωδ)(u_{\delta},\rho_{\delta},m_{\delta})\in W^{2,r}_{loc}(\Omega)\times\mathbb{R}\times W^{1,r}(\Omega_{\delta}) is a solution of (1.12)(\ref{eq:deltap}) if (uδ,ρδ)(u_{\delta},\rho_{\delta}) is an explosive solution of

{Δuδ+|Duδ|q+ρδ=F(x;mδ) in Ω,limd(x)0uδ(x)=,\displaystyle\begin{cases}-\Delta u_{\delta}+|Du_{\delta}|^{q}+\rho_{\delta}=F(x;{m}_{\delta})\text{ in }\Omega,\\ \lim\limits_{d(x)\to 0}u_{\delta}(x)=\infty,\end{cases}

and mδm_{\delta} is a Neumann weak solution of

Δmδ+div(mδq|Duδ|q2Duδ)=0 in Ωδ.\displaystyle\Delta m_{\delta}+div\left(m_{\delta}q|Du_{\delta}|^{q-2}Du_{\delta}\right)=0\text{ in }\Omega_{\delta}.

When FF satisfies (F2)(F2), existence of a solution to this system follows from Schaefer’s fixed point theorem.

Indeed, let T1:C0,α(Ωδ)Wloc1,r(Ω)T_{1}:C^{0,\alpha}(\Omega_{\delta})\to W^{1,r}_{loc}(\Omega) be the map that sends measures μ\mu to q|Du|q2Duq|Du|^{q-2}Du, where (u,ρ)(u,\rho) is the explosive solution of

{Δu+|Du|q+ρ=F(x;μ) in Ωlimd(x)0u(x)=,\displaystyle\begin{cases}-\Delta u+|Du|^{q}+\rho=F(x;{\mu})\text{ in }\Omega\\ \lim\limits_{d(x)\to 0}u(x)=\infty,\end{cases}

which is well defined for all α>0\alpha>0 and 1<r<1<r<\infty. It follows from Theorem 3.5 that the system has a solution uWloc2,r(Ω)u\in W^{2,r}_{loc}(\Omega) for every r>1r>1, and DuDu is unique.

Let T2:Wloc1,r(Ω)C0,α(Ωδ)T_{2}:W^{1,r}_{loc}(\Omega)\to C^{0,\alpha}(\Omega_{\delta}) be the map that sends bb to the Neumann weak solution mδm_{\delta} of

{Δmδ+div(mδbδ)=0 in Ωδ(Dmδ+mδbδ)ν=0 on Ωδ,\displaystyle\begin{cases}\Delta m_{\delta}+div(m_{\delta}b_{\delta})=0&\text{ in }\Omega_{\delta}\\ (Dm_{\delta}+m_{\delta}b_{\delta})\cdot\nu=0&\text{ on }\partial\Omega_{\delta},\end{cases}

where bδb_{\delta} is the restriction of bb to the domain Ωδ\Omega_{\delta}, which is also well defined. In view of Theorem 3.5, there is a unique Neumann weak solution mδm_{\delta}. Moreover, mδm_{\delta} is in W1,s(Ωδ)W^{1,s}(\Omega_{\delta}) for all 1<s<1<s<\infty and, by Sobelov embedding, is in C0,α(Ωδ)C^{0,\alpha}(\Omega_{\delta}) for all 0<α<10<\alpha<1. Composing the two, we define T=T2T1T=T_{2}\circ T_{1}.

Theorem 5.1.

For FF satisfying (F2)(F2) and α>0\alpha>0, T:C0,α(Ωδ)C0,α(Ωδ)T:C^{0,\alpha}(\Omega_{\delta})\to C^{0,\alpha}(\Omega_{\delta}) has a fixed point.

Proof.

We apply Schaefer’s fixed point theorem as in Theorem 4.1. As we will be working with an HJB equation set in Ω\Omega and a KFP equation set in Ωδ\Omega_{\delta}, we will be exploiting that the method for extending measures from Ωδ\Omega_{\delta} to Ω\Omega defined in (2.1)(\ref{tilde}) is C0,αC^{0,\alpha}-norm preserving.

To show the continuity of TT, we consider a sequence μn\mu_{n} which converges to μ\mu in C0,α(Ωδ)C^{0,\alpha}(\Omega_{\delta}), and we start by showing that T1(μn)T_{1}(\mu_{n}) converges to T1(μ)T_{1}(\mu) locally uniformly. To simplify notation, we let bn:=T1(μn)b_{n}:=T_{1}(\mu_{n}), b:=T1(μ)b:=T_{1}(\mu), and mn:=T2(bn)m_{n}:=T_{2}(b_{n}).

Let (un,ρn)(u_{n},\rho_{n}) be the unique explosive solution of (3.10)(\ref{eq:hj}) that corresponds to F(x;μn)F(x;{\mu}_{n}) and satisfies un(x0)=0u_{n}(x_{0})=0, where x0Ωx_{0}\in\Omega is fixed. As the sequence μn\mu_{n} converges to μ\mu in C0,α(Ωδ)C^{0,\alpha}(\Omega_{\delta}), μn\mu_{n} converges uniformly to μ\mu in Ωδ\Omega_{\delta}, and as a result μ~n\tilde{\mu}_{n} converges uniformly to μ~\tilde{\mu} in Ω\Omega, where μ~n\tilde{\mu}_{n} and μ~\tilde{\mu} are defined as in (2.1)(\ref{tilde}).

Due to the stability properties of viscosity solutions, to prove continuity of T1T_{1} under assumption (F2)(F2), it is sufficient to show that if xnxx_{n}\to x in Ω\Omega, then f(μ~n(xn))f(μ~(x))f(\tilde{\mu}_{n}(x_{n}))\to f(\tilde{\mu}(x)). However, this follows immediately from the continuity of ff and the equicontinuity of μ~n\tilde{\mu}_{n}.

Next we show that T2T_{2} is continuous. Since bnLloc(Ω)||b_{n}||_{L_{loc}^{\infty}(\Omega)} is bounded in nn, mnm_{n} is uniformly bounded in W1,r(Ωδ)W^{1,r}(\Omega_{\delta}) and, hence, bounded in C0,β(Ωδ)C^{0,\beta}(\Omega_{\delta}) for α<β<1n/r\alpha<\beta<1-n/r. In view of the compact embedding of C0,β(Ωδ)C^{0,\beta}(\Omega_{\delta}) in C0,α(Ωδ)C^{0,\alpha}(\Omega_{\delta}), there is a subsequence of mnm_{n} that converges in C0,α(Ωδ)C^{0,\alpha}(\Omega_{\delta}) to some mC0,α(Ωδ)m\in C^{0,\alpha}(\Omega_{\delta}). Moreover, DmnDm_{n} converges weakly to DmDm. Consequently, we can pass to the limit in the weak formulation of mnm_{n} and obtain that mW1,r(Ωδ)m\in W^{1,r}(\Omega_{\delta}) is a positive probability measure that satisfies the KFP equation with drift b=T1(μ)b=T_{1}(\mu). The uniqueness of mm yields convergence of T(μn)T(\mu_{n}) to T(μ)T(\mu) along the full sequence.

Next, to show that TT is precompact, we will use the bounds in Theorem 3.6, Lemma 3.1, and the Sobelov embedding theorem.

For μC0,α(Ωδ)\mu\in C^{0,\alpha}(\Omega_{\delta}) and b=T1(μ)b=T_{1}(\mu), Theorem 3.6 yields |b(x)|Cd(x)1|b(x)|\leq Cd(x)^{-1}, where CC depends on Ω\Omega and F(x;μ)L(Ω;L1(Ω))||F(x;{\mu})||_{L^{\infty}(\Omega;L^{1}(\Omega))}. Note that by assumption (F2)(F2), bL(Ωδ)||b||_{L^{\infty}(\Omega_{\delta})} is bounded by a constant depending on fL()||f||_{{L^{\infty}}(\mathbb{R})}.

Lemma 3.1 yields that T(μ)W1,r(Ωδ)||T(\mu)||_{W^{1,r}(\Omega_{\delta})} is bounded by a constant depending on fL()||f||_{L^{\infty}(\mathbb{R})}. Since W1,r(Ωδ)W^{1,r}(\Omega_{\delta}) embeds compactly in C0,α(Ωδ)C^{0,\alpha}(\Omega_{\delta}) for α<1n/r\alpha<1-n/r, choosing appropriately large rr, TT sends bounded sets of C0,α(Ωδ)C^{0,\alpha}(\Omega_{\delta}) to precompact ones.

It remains to check that if FF is the set defined by

F={mC0,α(Ωδ)|m=λT(m) for some 0λ1},\displaystyle F=\{m\in C^{0,\alpha}(\Omega_{\delta})|m=\lambda T(m)\text{ for some }0\leq\lambda\leq 1\},

then FF is bounded. This follows from Theorem 3.1 as q|Du|q2Duq|Du|^{q-2}Du is bounded in L(Ωδ)L^{\infty}(\Omega_{\delta}).

5.2 Limit of the approximate problem

In this section we prove that in the local setting, the solutions (uδ,ρ,mδ)(u_{\delta},\rho,m_{\delta}) of (1.12)(\ref{eq:deltap}) converge locally uniformly to solutions of (1.5)(\ref{eq:mainintro}). We begin by showing that, along subsequences, uδu_{\delta} has a local uniform limit uu using the Arzelà-Ascoli Theorem. We then construct a Lyapanov function and use Lemma 3.2 and Lemma 3.3 to show that the measures mδm_{\delta} converge locally uniformly to a proper weak solution mm of a continuity equation. Lastly we show that uu is the explosive solution corresponding to f(m(x))f(m(x)).

Theorem 5.2.

Given δ>0\delta>0 and FF satisfying (F2)(F2), let x0Ωx_{0}\in\Omega, and (uδ,ρδ,mδ)Wloc2,r(Ω)××W1,r(Ωδ)(u_{\delta},\rho_{\delta},m_{\delta})\in W_{loc}^{2,r}(\Omega)\times\mathbb{R}\times W^{1,r}(\Omega_{\delta}) be a solution of (1.12)(\ref{eq:deltap}) satisfying uδ(x0)=0u_{\delta}(x_{0})=0. Along subsequences, ρδ\rho_{\delta} converges to a constant ρ\rho and uδu_{\delta} converges in Cloc1(Ω)C^{1}_{loc}(\Omega) to uWloc2,r(Ω)u\in W_{loc}^{2,r}(\Omega).

Proof.

It follows from Theorem 3.6, and the uniform in δ\delta bounds on F(x;mδ)L(Ω;L1(Ω))||F(x;m_{\delta})||_{L^{\infty}(\Omega;L^{1}(\Omega))}, that ρδ\rho_{\delta} and the W1,r(Ωδ)W^{1,r}(\Omega_{\delta}) bounds of uδu_{\delta} are bounded uniformly in δ\delta. The desired convergence along subsequences follows by the Arzelà-Ascoli Theorem. ∎

For now we make no claims about the equation uu solves. We must first make sense of the limit of the measures mδm_{\delta}.

Theorem 5.3.

Given δ>0\delta>0 and FF satisfying (F2)(F2), let (uδ,ρδ,mδ)Wloc2,r(Ω)××W1,r(Ωδ)(u_{\delta},\rho_{\delta},m_{\delta})\in W_{loc}^{2,r}(\Omega)\times\mathbb{R}\times W^{1,r}(\Omega_{\delta}) be a solution of (1.12)(\ref{eq:deltap}). Along subsequences, mδm_{\delta} converges locally uniformly to mWloc1,r(Ω)m\in W^{1,r}_{loc}(\Omega) which is a proper weak solution of

Δm+div(mb)=0 in Ω,\displaystyle\Delta m+div(mb)=0\text{ in }\Omega,

where b=q|Du|q2Dub=q|Du|^{q-2}Du and uu is the limit function from Theorem 5.2.

Proof.

In view of Lemma 3.2 and Lemma 3.3, it suffices to construct a function VV satisfying (3.5)(\ref{v1}) and (3.6)(\ref{v2}) for bδ=q|Duδ|q2Duδb_{\delta}=q|Du_{\delta}|^{q-2}Du_{\delta}.

Consider V(x)=d(x)C+1+ϵV(x)=d(x)^{-C+1+\epsilon}, where C=q/(q1)C=q/(q-1) and 0<ϵ<C10<\epsilon<C-1, which satisfies the desired growth condition (3.5)(\ref{v1}) at the boundary. The L(Ω;L1(Ω))L^{\infty}(\Omega;L^{1}(\Omega))-norm of the coupling term F(x;mδ)F(x;{m}_{\delta}) is uniformly bounded in δ\delta by fL()||f||_{L^{\infty}(\mathbb{R})}. By Lemma 7.1, bδ(x)ν(x)d(x)Cb_{\delta}(x)\cdot\nu(x)d(x)\to C uniformly in δ\delta as d(x)0d(x)\to 0. It follows that the limit ΔVbδDV\Delta V-b_{\delta}\cdot DV\to-\infty at the boundary is likewise uniform in δ\delta, and (3.6)(\ref{v2}) is satisfied. The uniform tightness of the measures mδm_{\delta} ensures that Ωm=1\int_{\Omega}m=1 and mδm_{\delta} converges to mm in L1(Ω)L^{1}(\Omega). ∎

We are now ready to state the equation that uu solves.

Theorem 5.4.

Let FF satisfy (F2)(F2) and (uδ,ρδ,mδ)Wloc2,r(Ω)××W1,r(Ωδ)(u_{\delta},\rho_{\delta},m_{\delta})\in W_{loc}^{2,r}(\Omega)\times\mathbb{R}\times W^{1,r}(\Omega_{\delta}) be a solution of (1.12)(\ref{eq:deltap}). Let uu, mm and ρ\rho be as constructed in Theorems 5.1 and 5.2. Then (u,ρ)(u,\rho) is the explosive solution of

{Δu+|Du|q+ρ=F(x;m) in Ω,limd(x)0u(x)=.\displaystyle\begin{cases}-\Delta u+|Du|^{q}+\rho=F(x;m)\text{ in }\Omega,\\ \lim\limits_{d(x)\to 0}u(x)=\infty.\end{cases}
Proof.

The proof is very similar to the stability argument in Theorem 5.1, and follows from the local uniform Hölder bounds on mδm_{\delta}. Indeed, consider a sequence xδxx_{\delta}\to x contained in a compact set KK, and let M=supδmδC0,α(K)M=\sup_{\delta}||m_{\delta}||_{C^{0,\alpha}(K)}. Then,

|mδ(xδ)m(x)|\displaystyle|m_{\delta}(x_{\delta})-m(x)| |mδ(x)m(x)|+|mδ(xδ)mδ(x)|\displaystyle\leq|m_{\delta}(x)-m(x)|+|m_{\delta}(x_{\delta})-m_{\delta}(x)|
|mδ(x)m(x)|+M|xδx|α.\displaystyle\leq|m_{\delta}(x)-m(x)|+M|x_{\delta}-x|^{\alpha}. (5.1)

By the uniform convergence of mδm_{\delta} to mm on KK, (5.1)(\ref{eq:expression}) goes to 0 as δ0\delta\to 0. The continuity of ff allows us to pass to the limit in the HJB equation, obtaining that uu is a viscosity solution of

Δu+|Du|q+ρ=F(x;m) in Ω.\displaystyle-\Delta u+|Du|^{q}+\rho=F(x;m)\text{ in }\Omega.

The argument that uu blows up at the boundary is the same as in Theorem 3.8. ∎

6 Uniqueness

We are now ready to address the uniqueness claim in Theorem 3.9. Throughout the discussion we take 1<q<21<q<2 and deal with the case q=2q=2 at the end of the section.

Proof of Theorem 3.9.

It only remains to show that uniqueness holds when FF satisfies (F3)(F3). Here we follow the classical Lasry-Lions uniqueness argument. Consider two solutions (u1,ρ1,m1),(u2,ρ2,m2)Wloc2,r(Ω)××Wloc1,r(Ω)(u_{1},\rho_{1},m_{1}),(u_{2},\rho_{2},m_{2})\in W^{2,r}_{loc}(\Omega)\times\mathbb{R}\times W^{1,r}_{loc}(\Omega) of (1.5)(\ref{eq:mainintro}). Without loss of generality, we may assume that u1(x0)=u2(x0)=0u_{1}(x_{0})=u_{2}(x_{0})=0 for a fixed x0Ωx_{0}\in\Omega.

Letting u¯:=u1u2\overline{u}:=u_{1}-u_{2}, ρ¯:=ρ1ρ2\overline{\rho}:=\rho_{1}-\rho_{2}, and m¯:=m1m2\overline{m}:=m_{1}-m_{2}, it follows that (u¯,ρ¯,m¯)(\overline{u},\overline{\rho},\overline{m}) satisfies the system

{Δu¯+|Du1|q|Du2|q+ρ¯=F(x;m1)F(x;m2) in Ω,Δm¯+div(m1b1m2b2)=0 in Ω,\displaystyle\begin{cases}-\Delta\overline{u}+|Du_{1}|^{q}-|Du_{2}|^{q}+\overline{\rho}=F(x;m_{1})-F(x;m_{2})&\text{ in }\Omega,\\ \Delta\overline{m}+div(m_{1}b_{1}-m_{2}b_{2})=0&\text{ in }\Omega,\end{cases}

where b1:=q|Du1|q2Du1b_{1}:=q|Du_{1}|^{q-2}Du_{1} and b2:=q|Du2|q2Du2b_{2}:=q|Du_{2}|^{q-2}Du_{2}.

Consider a non-negative and compactly supported function ϕC(Ω)\phi\in C^{\infty}(\Omega); the particular choice of ϕ\phi will be made later.

Multiplying the first equation by m¯ϕ\overline{m}\phi and integrating by parts, we obtain

Ω[(Du¯)D(m¯ϕ)+(|Du1|q|Du1|q)m¯ϕ+(ρ¯)m¯ϕ]=Ω(F(x;m1)F(x;m2))m¯ϕ.\displaystyle\int_{\Omega}\left[(D\overline{u})\cdot D(\overline{m}\phi)+(|Du_{1}|^{q}-|Du_{1}|^{q})\overline{m}\phi+(\overline{\rho})\overline{m}\phi\right]=\int_{\Omega}(F(x;m_{1})-F(x;m_{2}))\overline{m}\phi.

Similarly, using u¯ϕ\overline{u}\phi as a test function in the KFP equation, we obtain

Ω[D(u¯ϕ)(Dm¯)(m1b1m2b2)D(u¯ϕ)]=0.\displaystyle\int_{\Omega}\left[-D(\overline{u}\phi)\cdot(D\overline{m})-(m_{1}b_{1}-m_{2}b_{2})\cdot D(\overline{u}\phi)\right]=0.

Adding the equations and regrouping gives:

Ω(F(x;m1)F(x;m2))m¯ϕ\displaystyle\int_{\Omega}(F(x;m_{1})-F(x;m_{2}))\overline{m}\phi =Ωϕ[(|Du1|q|Du1|q)m¯(m1b1m2b2)D(u¯)]\displaystyle=\int_{\Omega}\phi\left[(|Du_{1}|^{q}-|Du_{1}|^{q})\overline{m}-(m_{1}b_{1}-m_{2}b_{2})\cdot D(\overline{u})\right]
+Ω(ρ¯)m¯ϕ\displaystyle+\int_{\Omega}(\overline{\rho})\overline{m}\phi
Ω(Dϕ)[(m1b1m2b2)u¯]\displaystyle-\int_{\Omega}(D\phi)\cdot[(m_{1}b_{1}-m_{2}b_{2})\overline{u}]
+Ω(Dϕ)(m¯Du¯u¯Dm¯)\displaystyle+\int_{\Omega}(D\phi)\cdot(\overline{m}D\overline{u}-\overline{u}D\overline{m})
=Ωϕm1(|Du1|q|Du2|qb1D(u1u2))\displaystyle=\int_{\Omega}\phi m_{1}(|Du_{1}|^{q}-|Du_{2}|^{q}-b_{1}\cdot D(u_{1}-u_{2})) (6.1)
+Ωϕm2(|Du2|q|Du1|qb2D(u2u1))\displaystyle+\int_{\Omega}\phi m_{2}(|Du_{2}|^{q}-|Du_{1}|^{q}-b_{2}\cdot D(u_{2}-u_{1})) (6.2)
+Ω(ρ¯)m¯ϕ\displaystyle+\int_{\Omega}(\overline{\rho})\overline{m}\phi
Ω(Dϕ)[(m1b1m2b2)u¯+2m¯Du¯].\displaystyle-\int_{\Omega}(D\phi)\cdot[(m_{1}b_{1}-m_{2}b_{2})\overline{u}+2\overline{m}D\overline{u}].
+Ωm¯Δϕu¯\displaystyle+\int_{\Omega}\overline{m}\Delta\phi\overline{u}

The left-hand side is positive by the monotonicity assumption (F3)(F3), and (6.1)(\ref{con1}) and (6.2)(\ref{con2}) are negative by the convexity of the Hamiltonian. We will show that the remaining terms on the right hand side can be made arbitrarily small for the right choice of ϕ\phi.

We are now ready to define ϕ\phi. Let ψ(s):[0,)[0,1]\psi(s):[0,\infty)\to[0,1] be a smooth function that is 0 in a neighborhood of [0,1][0,1], 11 in a neighborhood of 22, and constant for s2s\geq 2. For δ<min(1,ϵ0)/2\delta<\min(1,\epsilon_{0})/2, let ϕ(x)=ψ(d(x)/δ)\phi(x)=\psi(d(x)/\delta). There exists a constant C1=C1(Ω,ψ)>0C_{1}=C_{1}(\Omega,\psi)>0 such that, in Ωδ\Ω2δ\Omega_{\delta}\backslash\Omega_{2\delta},

|Dϕ|1δψL([0,))DdL(Ωδ\Ω2δ)C11δ.\displaystyle|D\phi|\leq\dfrac{1}{\delta}||\psi^{\prime}||_{L^{\infty}([0,\infty))}||Dd||_{L^{\infty}(\Omega_{\delta}\backslash\Omega_{2\delta})}\leq C_{1}\dfrac{1}{\delta}. (6.3)

and

|Δϕ|1δ2ψ′′L([0,))DdL(Ωδ\Ω2δ)2+1δψL([0,))ΔdL(Ωδ\Ω2δ)C11δ2.\displaystyle|\Delta\phi|\leq\dfrac{1}{\delta^{2}}||\psi^{\prime\prime}||_{L^{\infty}([0,\infty))}||Dd||_{L^{\infty}(\Omega_{\delta}\backslash\Omega_{2\delta})}^{2}+\dfrac{1}{\delta}||\psi^{\prime}||_{L^{\infty}([0,\infty))}||\Delta d||_{L^{\infty}(\Omega_{\delta}\backslash\Omega_{2\delta})}\leq C_{1}\dfrac{1}{\delta^{2}}.

Together with the local bounds on |ui||u_{i}| and |Dui||Du_{i}| from Theorem 3.6, it follows from (6.3)(\ref{eq:dpsi}) that there exists a constant C2>0C_{2}>0, independent of δ\delta, for which

|Ω(Dϕ)[(m1b1m2b2)u¯+2m¯Du¯]|\displaystyle\left|\int_{\Omega}(D\phi)\cdot[(m_{1}b_{1}-m_{2}b_{2})\overline{u}+2\overline{m}D\overline{u}]\right| C2Ωδ\Ω2δ1δ(d1/(q1))q1(m1+m2)d(q2)/(q1)\displaystyle\leq C_{2}\int_{\Omega_{\delta}\backslash\Omega_{2\delta}}\dfrac{1}{\delta}(d^{-1/(q-1)})^{q-1}(m_{1}+m_{2})d^{(q-2)/(q-1)}
+C2Ωδ\Ω2δ1δd1/(q1)|m¯|\displaystyle+C_{2}\int_{\Omega_{\delta}\backslash\Omega_{2\delta}}\dfrac{1}{\delta}d^{-1/(q-1)}|\overline{m}|
2C2Ωδ\Ω2δdq/(q1)(m1+m2+|m¯|).\displaystyle\leq 2C_{2}\int_{\Omega_{\delta}\backslash\Omega_{2\delta}}d^{-q/(q-1)}(m_{1}+m_{2}+|\overline{m}|). (6.4)

Similarly there exists a constant C3>0C_{3}>0, independent of δ\delta, for which

|Ω(Δϕ)(u¯)m¯|\displaystyle\left|\int_{\Omega}(\Delta\phi)(\overline{u})\overline{m}\right| C3Ωδ\Ω2δ1δ2(m1+m2)d(q2)/(q1)\displaystyle\leq C_{3}\int_{\Omega_{\delta}\backslash\Omega_{2\delta}}\dfrac{1}{\delta^{2}}(m_{1}+m_{2})d^{(q-2)/(q-1)} (6.5)
4C3Ωδ\Ω2δdq/(q1)(m1+m2).\displaystyle\leq 4C_{3}\int_{\Omega_{\delta}\backslash\Omega_{2\delta}}d^{-q/(q-1)}(m_{1}+m_{2}). (6.6)

We recall from Lemma 3.4 that, for small ϵ\epsilon, Ωd(x)q/(q1)1+ϵm(x)\int_{\Omega}d(x)^{-q/(q-1)-1+\epsilon}m(x) is finite. It follows that (6.4)(\ref{b2}) and (6.6)(\ref{b3}) can be made arbitrarily small for δ\delta small enough.

To bound the remaining term, we use that m1+m2L1(Ω)m_{1}+m_{2}\in L^{1}(\Omega) and Ω(m1m2)=0\int_{\Omega}(m_{1}-m_{2})=0, obtaining

|Ω(ρ¯)m¯ϕ|\displaystyle\left|\int_{\Omega}(\overline{\rho})\overline{m}\phi\right| |ρ¯|Ωδ\Ω2δ|m1m2|+|ρ¯||Ω2δm1m2|\displaystyle\leq|\overline{\rho}|\int_{\Omega_{\delta}\backslash\Omega_{2\delta}}|m_{1}-m_{2}|+|\overline{\rho}|\left|\int_{\Omega_{2\delta}}m_{1}-m_{2}\right|
|ρ¯|(Ω\Ω2δ(m1+m2)+|0Ω\Ω2δ(m1m2)|)\displaystyle\leq|\overline{\rho}|\left(\int_{\Omega\backslash\Omega_{2\delta}}(m_{1}+m_{2})+\left|0-\int_{\Omega\backslash\Omega_{2\delta}}(m_{1}-m_{2})\right|\right)
2|ρ¯|Ω\Ω2δ(m1+m2),\displaystyle\leq 2|\overline{\rho}|\int_{\Omega\backslash\Omega_{2\delta}}(m_{1}+m_{2}),

which can also be made arbitrarily small for δ\delta small enough.

It follows from the monotonicity of F(x;m)F(x;m), that m1=m2m_{1}=m_{2}. That u1=u2u_{1}=u_{2} and ρ1=ρ2\rho_{1}=\rho_{2} follows from the uniqueness of solutions to (3.10)(\ref{eq:hj}). ∎

Remark.

To prove uniqueness of solutions to (1.5)(\ref{eq:mainintro}) when q=2q=2, (6.4)(\ref{b2}) should be replaced by an estimate of the form

CΩδ\Ω2δd1(log(d))(m1+m2),\displaystyle C\int_{\Omega_{\delta}\backslash\Omega_{2\delta}}d^{-1}(-\log(d))(m_{1}+m_{2}), (6.7)

where C>0C>0 is a constant independent of δ\delta. As d1(log(d))d^{-1}(-\log(d)) has growth slower than d2+ϵd^{-2+\epsilon}, the rest of the proof then proceeds as in the 1<q<21<q<2 case.

7 Appendix

In this section we prove Lemma 7.1, a stronger version of (ii)(ii) in Theorem 3.7, and Lemma 3.4.

Lemma 7.1.

Let mWloc1,r(Ω)m\in W^{1,r}_{loc}(\Omega) and consider the unique solution (um,ρm)Wloc2,r(Ω)×(u_{m},\rho_{m})\in W^{2,r}_{loc}(\Omega)\times\mathbb{R} of

{Δum+|Dum|q+ρm=F(x;m) in Ω,limd(x)0um(x)=.\displaystyle\begin{cases}-\Delta u_{m}+|Du_{m}|^{q}+\rho_{m}=F(x;m)\text{ in }\Omega,\\ \lim\limits_{d(x)\to 0}u_{m}(x)=\infty.\end{cases} (7.1)

Let bm:=q|Dum|q2Dumb_{m}:=q|Du_{m}|^{q-2}Du_{m}. Then, uniformly in F(x;m)L(Ω;L1(Ω))||F(x;m)||_{L^{\infty}(\Omega;L^{1}(\Omega))},

limd(x)0(bm(x)ν(x))d(x)=qq1.\lim_{d(x)\to 0}(b_{m}(x)\cdot\nu(x))d(x)=\dfrac{q}{q-1}.

It will be enough to show uniform convergence of (Dum(x)ν(x))d(x)1/(q1)(Du_{m}(x)\cdot\nu(x))d(x)^{1/(q-1)} as d(x)0d(x)\to 0.
We prove this by following the proof of Theorem 2.3 in [11] and tracking how the convergence depends on F(x;m)L(Ω;L1(Ω))||F(x;m)||_{L^{\infty}(\Omega;L^{1}(\Omega))}.

In the proof that follows, we argue first when 1<q<21<q<2, and discuss the q=2q=2 case at the end of the proof. The same arguments also hold for Hamiltonians satisfying (1.4)(\ref{eq:gen}) locally uniformly in pp as δ0\delta\to 0.

Proof of Lemma 7.1.

Since Ω\Omega has a uniform C2C^{2} boundary, by Theorem 3.7 there exists a δ0>0\delta_{0}>0, depending on F(x;m)L(Ω;L1(Ω))||F(x;m)||_{L^{\infty}(\Omega;L^{1}(\Omega))}, such that for every xΩx\in\partial\Omega, the ball Bδ0(xδ0ν(x))B_{\delta_{0}}(x-\delta_{0}\nu(x)) is contained in Ω\Omega, and, for d(x)<δ0d(x)<\delta_{0} ,

|um(x)d(x)(q2)/(q1)Cq|Cq2,\displaystyle\left|\dfrac{u_{m}(x)}{d(x)^{(q-2)/(q-1)}}-C_{q}\right|\leq\dfrac{C_{q}}{2}, (7.2)

where Cq=(q1)(q2)/(q1)(2q)1C_{q}=(q-1)^{(q-2)/(q-1)}(2-q)^{-1}.

We begin by fixing x0Ωx_{0}\in\partial\Omega and a system of coordinates (η1,,ηn)(\eta_{1},\ldots,\eta_{n}) centered at x0x_{0} whose η1\eta_{1}-axis is the inner normal direction. Although we have fixed an x0x_{0}, any bounds and rates that follow will not depend on x0Ωx_{0}\in\partial\Omega.

Let Oδ=(δ,,0)O_{\delta}=(\delta,\ldots,0) and consider the domain

Dδ=B(Oδ,δ1σ)B((δ0,0,,0),δ0δ) , σ(0,1/2).\displaystyle D_{\delta}=B(O_{\delta},\delta^{1-\sigma})\cap B((\delta_{0},0,\ldots,0),\delta_{0}-\delta)\text{ , }\sigma\in(0,1/2).

Applying the change of variable ζ=(ηOδ)/δ\zeta=(\eta-O_{\delta})/\delta, the domain D~δ:={(ηOδ)/δ:ηDδ}\tilde{D}_{\delta}:=\{(\eta-O_{\delta})/\delta:\eta\in D_{\delta}\} converges to the half space {ζ:ζ1>0}\{\zeta:\zeta_{1}>0\}, where ζ1\zeta_{1} is the component of the vector in the ν(x)-\nu(x) direction. Under this change of variable we have the bounds

(ζ1+1)δd(η)(ζ1+1)δ+O(δ22σ).\displaystyle(\zeta_{1}+1)\delta\leq d(\eta)\leq(\zeta_{1}+1)\delta+O(\delta^{2-2\sigma}). (7.3)

It is now possible to introduce, for δ<δ0\delta<\delta_{0}, the blow up function

vδ,m(ζ)=um(δζ+Oδ)δ(q2)/(q1),\displaystyle v_{\delta,m}(\zeta)=\dfrac{u_{m}(\delta\zeta+O_{\delta})}{\delta^{(q-2)/(q-1)}},

with domain D~δ\tilde{D}_{\delta}, which, in view of (7.2)(\ref{d_0}), is uniformly bounded in δ,F(x;m)L(Ω;L1(Ω)),\delta,||F(x;m)||_{L^{\infty}(\Omega;L^{1}(\Omega))}, and ζ\zeta.

Straightforward computations show that vδ,mv_{\delta,m} describes the blow-up of umu_{m} near the boundary. Indeed, for η1=δζ1+Oδ\eta_{1}=\delta\zeta_{1}+O_{\delta},

Dvδ,m(ζ1)=Dum(δζ1+Oδ)Cqδ1/(q1)=Dum(η1)Cqd(η1)1/(q1)(1+ζ1)1/(q1).\displaystyle Dv_{\delta,m}(\zeta_{1})=\dfrac{Du_{m}(\delta\zeta_{1}+O_{\delta})}{C_{q}\delta^{-1/(q-1)}}=\dfrac{Du_{m}(\eta_{1})}{C_{q}d(\eta_{1})^{-1/(q-1)}}(1+\zeta_{1})^{-1/(q-1)}.

In D~δ\tilde{D}_{\delta}, vδ,mv_{\delta,m} solves, in the viscosity sense,

Δvδ,m+ρmδq/(q1)Cq+Cqq1|Dvδ,m|q=F(δζ+Oδ;m)δq/(q1)Cq.\displaystyle-\Delta v_{\delta,m}+\dfrac{\rho_{m}\delta^{q/(q-1)}}{C_{q}}+C_{q}^{q-1}|Dv_{\delta,m}|^{q}=\dfrac{F(\delta\zeta+O_{\delta};m)\delta^{q/(q-1)}}{C_{q}}. (7.4)

Moreover, Dvδ,m(ζ)Dv_{\delta,m}(\zeta) satisfies the estimate

|Dvδ,m(ζ)|=|Dum(δζ+Oδ)|Cqδ1/(q1)Cmd(η)1/(q1)δ1/(q1)Cm,\displaystyle|Dv_{\delta,m}(\zeta)|=\dfrac{|Du_{m}(\delta\zeta+O_{\delta})|}{C_{q}\delta^{-1/(q-1)}}\leq C_{m}d(\eta)^{-1/(q-1)}\delta^{1/(q-1)}\leq C_{m},

where Cm=Cm(F(x;m)L(Ω;L1(Ω)))C_{m}=C_{m}(||F(x;m)||_{L^{\infty}(\Omega;L^{1}(\Omega))}) comes from the locally uniform bound on DuDu from Theorem 3.6.

Consider a family of {vδ,m}\{v_{\delta,m}\} with δ0\delta\to 0 and measures mm such that F(x;m)L(Ω;L1(Ω))||F(x;m)||_{L^{\infty}(\Omega;L^{1}(\Omega))} is uniformly bounded. This is the case, for example, if (F1)(F1) or (F2)(F2) are satisfied and the measures mm are uniformly bounded in L1(Ω)L^{1}(\Omega). It follows from elliptic regularity that {vδ,m}\{v_{\delta,m}\} has a convergent subsequence converging in Cloc1C^{1}_{loc} to vv defined in the half-plane {ζ:ζ1>0}\{\zeta:\zeta_{1}>0\}.

Since ρmδq/(q1)Cq0\dfrac{\rho_{m}\delta^{q/(q-1)}}{C_{q}}\to 0 and F(δζ+Oδ;m)δq/(q1)Cq0\dfrac{F(\delta\zeta+O_{\delta};m)\delta^{q/(q-1)}}{C_{q}}\to 0 uniformly, vv is a viscosity solution of

Δv+Cqq1|Dv|q=0 in {ζ:ζ1>0}.\displaystyle-\Delta v+C_{q}^{q-1}|Dv|^{q}=0\text{ in }\{\zeta:\zeta_{1}>0\}. (7.5)

In addition, vv satisfies the boundary conditions

limζ10+v(ζ)=Cq and limζ1v(ζ)=0.\displaystyle\lim_{\zeta_{1}\to 0^{+}}v(\zeta)=C_{q}\text{ and }\lim_{\zeta_{1}\to\infty}v(\zeta)=0.

The boundary conditions follow from

vδ,m(ζ)=um(η)d(η)(q2)/(q1)d(η)(q2)/(q1)δ(q2)/(q1),\displaystyle v_{\delta,m}(\zeta)=\dfrac{u_{m}(\eta)}{d(\eta)^{(q-2)/(q-1)}}\dfrac{d(\eta)^{(q-2)/(q-1)}}{\delta^{(q-2)/(q-1)}}, (7.6)

and the estimates in (7.3)(\ref{eta}). As shown in Theorem 4.1 of [11], the equation (7.5)(\ref{v}) paired with boundary conditions (7.6)(\ref{bounds}) has a unique positive solution in the half-plane, v(ζ)=Cq(1+ζ1)(q2)/(q1)v(\zeta)=C_{q}(1+\zeta_{1})^{(q-2)/(q-1)}. From (7.2)(\ref{d_0}), vv is indeed positive. Thus Dvδ,mDv_{\delta,m} converges locally uniformly to Cq(1+ζ1)1/(q1)(2q)/(q1)ν(x0)C_{q}(1+\zeta_{1})^{-1/(q-1)}(2-q)/(q-1)\nu(x_{0}), and (Dum(x)ν(x))d(x)1/(q1)(Du_{m}(x)\cdot\nu(x))d(x)^{1/(q-1)} converges uniformly in F(x;m)L(Ω;L1(Ω))||F(x;m)||_{L^{\infty}(\Omega;L^{1}(\Omega))}, as d(x)0d(x)\to 0.

When q=2q=2, rescaling umu_{m} as before does not yield an equation of the type (7.4)(\ref{vd}). To circumvent this, we consider a function of the form um(η)+log(d(η))u_{m}(\eta)+\log(d(\eta)), which is uniformly bounded in F(x;m)L(Ω;L1(Ω))||F(x;m)||_{L^{\infty}(\Omega;L^{1}(\Omega))} and has a limiting equation of the desired form.

For the case q=2q=2, consider a coordinate system centered at x0Ωx_{0}\in\partial\Omega, denoted by OO, with the η1\eta_{1}-axis as the inner normal direction. By the inner sphere property, there is δ0>0\delta_{0}>0 such that for every xΩx\in\partial\Omega, the ball Bδ0(xδ0ν(x))B_{\delta_{0}}(x-\delta_{0}\nu(x)) is contained in Ω\Omega. Under this new coordinate system define

Dδ=B(O,δ1σ)B((δ0,0,,0),δ0) , σ(0,1/2).\displaystyle D_{\delta}=B(O,\delta^{1-\sigma})\cap B((\delta_{0},0,\ldots,0),\delta_{0})\text{ , }\sigma\in(0,1/2).

Note that we changed the definition of DδD_{\delta} here.

Restricting umu_{m} to DδD_{\delta}, the function

vδ,m(ζ)=um(δζ)+log(δ),\displaystyle v_{\delta,m}(\zeta)=u_{m}(\delta\zeta)+\log(\delta),

is defined on the domain D~δ:={η/δ:ηDδ}\tilde{D}_{\delta}:=\{\eta/\delta:\eta\in D_{\delta}\}. As before, D~δ\tilde{D}_{\delta} converges to the half space {ζ:ζ1>0}\{\zeta:\zeta_{1}>0\}, where ζ1\zeta_{1} is the component of the vector in the ν(x)-\nu(x) direction. Under this change of variable we have the bounds

ζ1δd(η)ζ1δ+O(δ22σ).\displaystyle\zeta_{1}\delta\leq d(\eta)\leq\zeta_{1}\delta+O(\delta^{2-2\sigma}).

In D~δ\tilde{D}_{\delta}, vδ,mv_{\delta,m} satisfies, in the viscosity sense,

Δvδ,m+ρmδ2+|Dvδ,m|2=F(δζ;m)δ2.\displaystyle-\Delta v_{\delta,m}+\rho_{m}\delta^{2}+|Dv_{\delta,m}|^{2}=F(\delta\zeta;m)\delta^{2}.

By Theorem II.3 of [7], um(δζ)+log(d(η))u_{m}(\delta\zeta)+\log(d(\eta)) is bounded by a constant CC which is uniform in F(x;m)L(Ω;L1(Ω))||F(x;m)||_{L^{\infty}(\Omega;L^{1}(\Omega))}. From the bounds (7.3)(\ref{eta}) on d(η)d(\eta) it follows that vδ,mv_{\delta,m} is bounded locally uniformly. From standard elliptic theory, vδv_{\delta} is relatively compact in Cloc1C^{1}_{loc}. Thus, along subsequences vδ,mv_{\delta,m} converges in Cloc1C^{1}_{loc} to a function vv, defined in the upper-half plane, which is a viscosity solution of

Δv+|Dv|2=0.\displaystyle-\Delta v+|Dv|^{2}=0.

Moreover, for constant CC as defined above,

vδ,m\displaystyle v_{\delta,m} =um(δζ)+log(d(η))+log(δd(η))\displaystyle=u_{m}(\delta\zeta)+\log(d(\eta))+\log\left(\dfrac{\delta}{d(\eta)}\right)
C+log(1ζ1+O(δ12σ)).\displaystyle\geq-C+\log\left(\dfrac{1}{\zeta_{1}+O(\delta^{1-2\sigma})}\right).

Thus we have the boundary condition

limζ10+v(ζ)=.\displaystyle\lim\limits_{\zeta_{1}\to 0^{+}}v(\zeta)=\infty.

Solutions of the system

{Δv+|Dv|2=0 in {ζn|ζ1>0},limζ10+v(ζ)=,\displaystyle\begin{cases}-\Delta v+|Dv|^{2}=0\text{ in }\{\zeta\in\mathbb{R}^{n}\,|\,\zeta_{1}>0\},\\ \lim\limits_{\zeta_{1}\to 0^{+}}v(\zeta)=\infty,\end{cases}

are of the form logζ1+κ-\log\zeta_{1}+\kappa for a constant κ\kappa, which follows from the change of variable w=evw=e^{-v}. Thus Dvδ,mDv_{\delta,m} converges locally uniformly to (1/ζ1)ν(x0)(1/\zeta_{1})\nu(x_{0}), and (Dum(x)ν(x))d(x)(Du_{m}(x)\cdot\nu(x))d(x) converges uniformly in F(x;m)L(Ω;L1(Ω))||F(x;m)||_{L^{\infty}(\Omega;L^{1}(\Omega))}, as d(x)0d(x)\to 0.

We now prove Lemma 3.4 from Section 3.1.

Lemma 3.4.

Let b:Ωdb:\Omega\to\mathbb{R}^{d} satisfy (3.9)(\ref{eq:asym}) with C>1C>1. Then there exists a unique proper weak solution mWloc1,r(Ω)m\in W^{1,r}_{loc}(\Omega) of

Δm+div(mb)=0 in Ω,\displaystyle\Delta m+div(mb)=0\text{ in }\Omega, (7.7)

and, for every ϵ(0,C1)\epsilon\in(0,C-1), there exists a δ=δ(C,ϵ)\delta=\delta(C,\epsilon), such that

ΩdC1+ϵmC^=C^(C,bL(Ωδ)).\displaystyle\int_{\Omega}d^{-C-1+\epsilon}m\leq\hat{C}=\hat{C}(C,||b||_{L^{\infty}(\Omega_{\delta})}). (7.8)
Proof.

The existence of a proper weak solution mm was proven in Section 3.1. To prove (7.8), we mimick the proof of Lemma 1.1 in [1].

Let V(x)=d(x)C+1+ϵV(x)=d(x)^{-C+1+\epsilon} for 0<ϵ<C10<\epsilon<C-1. In the neighborhood Ω\Ωϵ0\Omega\backslash\Omega_{\epsilon_{0}} of Ω\partial\Omega,

ΔVbDV\displaystyle\Delta V-b\cdot DV =(C+1+ϵ)dC1+ϵ[(C+ϵ)|Dd|2+dΔd(bDd)d]\displaystyle=(-C+1+\epsilon)d^{-C-1+\epsilon}[(-C+\epsilon)|Dd|^{2}+d\Delta d-(b\cdot Dd)d]
=(C+1+ϵ)dC1+ϵ[C+ϵ+dΔd(bDd)d].\displaystyle=(-C+1+\epsilon)d^{-C-1+\epsilon}[-C+\epsilon+d\Delta d-(b\cdot Dd)d].

Since the last factor approaches ϵ\epsilon as d0d\to 0, in a neighborhood of Ω\partial\Omega, we have

ΔVbDVdC1+ϵ.\displaystyle\Delta V-b\cdot DV\lesssim-d^{-C-1+\epsilon}. (7.9)

Let δ0>0\delta_{0}>0 be such that, for xΩ/Ωδ0x\in\Omega/\Omega_{\delta_{0}},

ΔVbDV1,\displaystyle\Delta V-b\cdot DV\leq-1,

and define S:=Ωδ0|ΔVbDV|mS:=\int_{\Omega_{\delta_{0}}}|\Delta V-b\cdot DV|m.

Note that since mm is a probability measure on Ω\Omega, SS is bounded by the LL^{\infty} norm of |ΔVbDV||\Delta V-b\cdot DV| on Ωδ0\Omega_{\delta_{0}}.

We claim that, for all δ<δ0\delta<\delta_{0},

Ωδ|ΔVbDV|m2S.\int_{\Omega_{\delta}}|\Delta V-b\cdot DV|m\leq 2S. (7.10)

Since SS is independent of δ\delta, it would then follow that

Ω|ΔVbDV|m2S,\displaystyle\int_{\Omega}|\Delta V-b\cdot DV|m\leq 2S,

and the growth of ΔVbDV\Delta V-b\cdot DV in (7.9)(\ref{eq:Vass}) completes the proof of the lemma.

To prove (7.10)(\ref{eq:V}), fix δ\delta, ss, and rr, such that 0<δ0C+1+ϵ<s<δC+1+ϵ<r0<\delta_{0}^{-C+1+\epsilon}<s<\delta^{-C+1+\epsilon}<r. Consider a nondecreasing cut-off function ϕC2()\phi\in C^{2}(\mathbb{R}) such that ϕ′′0\phi^{\prime\prime}\leq 0, ϕ(t)=t\phi(t)=t for 0ts0\leq t\leq s, and ϕ(t)=r\phi(t)=r for trt\geq r. Since ϕ(V(x))\phi(V(x)) is constant in a neighborhood of the boundary Ω\partial\Omega, it follows that it is a permissible test function in (7.7)(\ref{eq:mapp}). The sign of ϕ′′\phi^{\prime\prime} yields,

0\displaystyle 0 =Ω[Δ(ϕ(V))bD(ϕ(V)]\displaystyle=\int_{\Omega}[\Delta(\phi(V))-b\cdot D(\phi(V)]
=Ω[ϕ′′(V)|DV|2+ϕ(V)ΔVϕ(V)bDV]m\displaystyle=\int_{\Omega}[\phi^{\prime\prime}(V)|DV|^{2}+\phi^{\prime}(V)\Delta V-\phi^{\prime}(V)b\cdot DV]m
Ωϕ(V)[ΔVbDV]m.\displaystyle\leq\int_{\Omega}\phi^{\prime}(V)[\Delta V-b\cdot DV]m.

The last integrand is zero outside {Vr}\{V\leq r\}, equals (ΔVbDV)m(\Delta V-b\cdot DV)m on {Vs}\{V\leq s\} and, is non-positive on {sVr}\{s\leq V\leq r\}. Consequently,

{Vs}(ΔVbDV)m0.\displaystyle\int_{\{V\leq s\}}(\Delta V-b\cdot DV)m\geq 0.

Letting sδC+1+ϵs\to\delta^{-C+1+\epsilon} yields

Ωδ(ΔVbDV)m0.\displaystyle\int_{\Omega_{\delta}}(\Delta V-b\cdot DV)m\geq 0.

Since ΔVbDV0\Delta V-b\cdot DV\leq 0 on Ωδ\Ωδ0\Omega_{\delta}\backslash\Omega_{\delta_{0}}, and Ωδ0|ΔVbDV|m=S\int_{\Omega_{\delta_{0}}}|\Delta V-b\cdot DV|m=S, it follows that

Ωδ|ΔVbDV|m2S.\displaystyle\int_{\Omega_{\delta}}|\Delta V-b\cdot DV|m\leq 2S.

References

  • [1] V. Bogachev and M. Röckner. A generalization of Khasminskii’s theorem on the existence of invariant measures for locally integrable drifts. (Russian). Veroyatnost. i Primenen., 45(3):417–436, 2000.
  • [2] P. Cardaliaguet. Notes on mean field games(from p.-l. lions’ lectures at collége de france), 2013.
  • [3] D. Gomes, E. Pimentel, and V. Voskanyan. Regularity Theory for Mean-Field Game Systems. SpringerBriefs in Mathematics. Springer International Publishing, 2016.
  • [4] 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. 2006.
  • [5] W. Huang, M. Ji, Z. Liu, and Y. Yi. Steady states of fokker–planck equations: I. existence. Journal of Dynamics and Differential Equations, 27:721–742, 2015.
  • [6] O. A. Ladyzhenskaya and N. N. Uraltseva. Linear and quasilinear elliptic equations. Academic Press New York, 1968.
  • [7] J.-M. Lasry and P.-L. Lions. Nonlinear elliptic equations with singular boundary conditions and stochastic control with state constraints. Math. Ann., 283:583–630, 1989.
  • [8] J.-M. Lasry and P.-L. Lions. Mean field games. Japanese Journal of Mathematics, 2:229–260, 2007.
  • [9] P.-L. Lions. Cours au collège de france. http://www.college-de-france.fr.
  • [10] A. Porretta and M. Ricciardi. Mean field games under invariance conditions for the state space. Communications in Partial Differential Equations, 45:146 – 190, 2019.
  • [11] A. Porretta and L. Véron. Asymptotic behaviour for the gradient of large solutions to some nonlinear elliptic equations. Adv. Nonlinear Stud., 2006.