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

On Numerical approximations of fractional and nonlocal Mean Field Games

Indranil Chowdhury (I. Chowdhury) Department of Mathematics, Faculty of Science, University of Zagreb,  Croatia indranil.chowdhury@ntnu.no, indranill2011@gmail.com Olav Ersland (O. Ersland) Norwegian University of Science and Technology, Norway olav.ersland@ntnu.no  and  Espen R. Jakobsen (E. R. Jakobsen) Norwegian University of Science and Technology, Norway espen.jakobsen@ntnu.no
Abstract.

We construct numerical approximations for Mean Field Games with fractional or nonlocal diffusions. The schemes are based on semi-Lagrangian approximations of the underlying control problems/games along with dual approximations of the distributions of agents. The methods are monotone, stable, and consistent, and we prove convergence along subsequences for (i) degenerate equations in one space dimension and (ii) nondegenerate equations in arbitrary dimensions. We also give results on full convergence and convergence to classical solutions. Numerical tests are implemented for a range of different nonlocal diffusions and support our analytical findings.

Key words and phrases:
Mean Field Games, jump diffusion, anomalous diffusion, nonlocal operators, fractional PDEs, nonlocal PDEs, degenerate PDEs, semi-Lagrangian scheme, convergence, compactness, Fokker-Planck equations, Hamilton-Jacobi-Bellman equations, duality methods
2020 Mathematics Subject Classification:
35Q89, 47G20, 35Q84, 49L12, 45K05, 35K61, 65M12, 91A16, 65M22, 35R11 , 35R06,

1. Introduction

In this article we study numerical approximations of Mean Field Games (MFGs) with fractional and general non-local diffusions. We consider the mean field game system

(1) {utu+H(x,Du)=F(x,m(t)), in (0,T)×d,mtmdiv(mDpH(x,Du))=0 in (0,T)×d,u(T,x)=G(x,m(T)),m(0)=m0 in d,\displaystyle\begin{cases}-u_{t}-\mathcal{L}u+H(x,Du)=F(x,m(t)),\quad&\text{ in }(0,T)\times\mathbb{R}^{d},\\ m_{t}-\mathcal{L}^{*}m-\text{div}(mD_{p}H(x,Du))=0\quad&\text{ in }(0,T)\times\mathbb{R}^{d},\\ u(T,x)=G(x,m(T)),\ m(0)=m_{0}\quad&\text{ in }\mathbb{R}^{d},\end{cases}

where

(2) ϕ(x)=|z|>0[ϕ(x+z)ϕ(x)𝟙{|z|<1}Dϕ(x)z]𝑑ν(z),\displaystyle\mathcal{L}\phi(x)=\int_{|z|>0}\big{[}\phi(x+z)-\phi(x)-\mathbbm{1}_{\{|z|<1\}}D\phi(x)\cdot z\big{]}d\nu(z),

is a nonlocal diffusion operator (possibly degenerate), ν\nu is a Lévy measure (see assumption (ν\nu0): ), and the adjoint \mathcal{L}^{*} is defined as (ϕ,ψ)L2=(ϕ,ψ)L2(\mathcal{L}^{*}\phi,\psi)_{L^{2}}=(\phi,\mathcal{L}\psi)_{L^{2}} for ϕ,ψCc(d)\phi,\psi\in C_{c}^{\infty}(\mathbb{R}^{d}).

The first equation in (1) is a backward in time Hamilton-Jacobi-Bellman (HJB) equation with terminal data GG, and the second equation is a forward in time Fokker-Planck-Kolmogorov (FPK) equation with initial data m0m_{0}. Here HH is the Hamiltonian, and the system is coupled through the cost functions FF and GG. There are two different types of couplings: (i) Local couplings where FF and GG depend on point values of mm, and (ii) non-local or smoothing couplings where they depend on distributional properties induced from mm through integration or convolution. Here we work with nonlocal couplings.

A mathematical theory of MFGs were introduced by Lasry–Lions [49] and Caines–Huang–Malhame [44], and describes the limiting behavior of NN-player stochastic differential games when the number of players NN tends to \infty [18]. In recent years there has been significant progress on MFG systems with local (or no) diffusion, including e.g. modeling, wellposedness, numerical approximations, long time behavior, convergence of Nash equilibria, and various control and game theoretic questions, see e.g. [5, 27, 18, 13, 39, 43] and references therein. The study of MFGs with ‘non-local diffusion’ is quite recent, and few results exist so far. Stationary problems with fractional Laplacians were studied in [30], and parabolic problems including (1), in [33] and [37]. We refer to [48] and references therein for some development using probabilistic methods.

The difference between problem (1) and standard MFG formulations lies in the type of noise driving the underlying controlled stochastic differential equations (SDEs). Usually Gaussian noise is considered [49, 51, 20, 26, 5], or there is no noise (the first order case) [17, 19]. Here the underlying SDEs are driven by pure jump Lévy processes, which leads to nonlocal operators (2) in the MFG system. In many real world applications, jump processes model the observed noise better than Gaussian processes [9, 50, 34, 54]. Prototypical examples are symmetric σ\sigma-stable processes and their generators, the fractional Laplace operators ()σ2(-\triangle)^{\frac{\sigma}{2}}. In Economy and Finance the observed noise is not symmetric and σ\sigma-stable, but rather non-symmetric and tempered. A typical example is the one-dimensional CGMY process [34] where dνdz(z)=C|z|1+YeGz+Mz\frac{d\nu}{dz}(z)=\frac{C}{|z|^{1+Y}}e^{-Gz^{+}-Mz^{-}} for C,G,M>0C,G,M>0 and Y(0,2)Y\in(0,2). Such models are covered by the results of this article. Our assumptions on the nonlocal operators (cf. (ν\nu1): ) are quite general, allowing for degenerate operators and no restrictions on the tail of the Lévy measure ν\nu.

There has been some development on numerical approximations for MFG systems with local operators. Finite difference schemes for nondegenerate second order equations have been designed and analyzed e.g. by Achdou et al. [1, 2, 3, 4, 7, 8, 6] and Gueant [40, 42, 41]. Semi-Lagrangian (SL) schemes for MFG system have been developed by Carlini–Silva both for first order equations [23] and possibly degenerate second order equations [24]. Other numerical schemes for MFGs include recent machine learning methods [28, 29, 52] for high dimensional problems. We refer to the survey article [6] for recent developments on numerical methods for MFG. We know of no prior schemes or numerical analysis for MFGs with fractional or nonlocal diffusions.

In this paper we will focus on SL schemes. They are monotone, stable, connected to the underlying control problem, easily handles degenerate and arbitrarily directed diffusions, and large time steps are allowed. Although the SL schemes for HJB equations have been studied for some time (see e.g. [38, 16, 14, 35]), there are few results for FPK equations (but see [25]) and the coupled MFG system. For nonlocal problems we only know of the results in [15] for HJB equations.

Our contributions

A. Derivation. We construct fully discrete monotone numerical schemes for the MFG system (1). These dual SL schemes are closely related to the underlying control formulation of the MFG. In our case it is based on the following controlled SDE:

dXt=αtdt+dLt,dX_{t}=-\alpha_{t}\,dt+dL_{t},

where αt\alpha_{t} is the control and LtL_{t} a pure jump Lévy process (cf. (6)). Note that LtL_{t} can be decomposed into small and large jumps, where the small jumps may have infinite intensity. We derive our approximation in several steps:

  1. 1.

    (Approximate small jumps) The small jumps are approximated by Brownian motion (see (7)) following e.g. [10, 15, 36]. This is done to avoid infinitely many jumps per time-interval and singular integrals, and gives a better approximation compared to simply neglecting these terms.

  2. 2.

    (SL scheme for HJB) We discretise the resulting SDE from step 1 in time and approximate the noise by random walks and approximate compound Poisson processes in the spirit of [15] (Section 3.1). From the corresponding discrete time optimal control problem, dynamic programming, and interpolation we construct an SL scheme for the HJB equation (Section 3.2).

  3. 3.

    (Approximate control) We define an approximate optimal feedback control for the SL scheme in step 2 from the continuous optimal feedback control as in [23, 24]: αapprox=DpH(,Dudϵ)\alpha^{*}_{\textup{approx}}=D_{p}H(\cdot,Du_{d}^{\epsilon}), where udϵu_{d}^{\epsilon} is a regularization of the (interpolated) solution from step 2 (Section 3.3).

  4. 4.

    (Dual SL scheme for FPK) The control of step 3 and the scheme in step 2 define a controlled approximate SDE with a corresponding discrete FPK equation for the densities of the solutions. We explicitly derive this FPK equation in weak form, and obtain the final dual SL scheme taking test functions to be linear interpolation basis functions (Section 3.4).

See (18) and (24) in Section 3 for the specific form of our discretizations. These seem to be the first numerical approximations of MFG systems with nonlocal or fractional diffusion and the first SL approximations of nonlocal FPK equations. Our dual SL schemes are extensions to the nonlocal case of the schemes in [23, 24, 25], but a clear derivation of such type of schemes seems to be new. The schemes come in the form of nonlinear coupled systems (27) that need to be resolved numerically. We prove existence of solutions using fixed point arguments, see Proposition 3.4.

B. Analysis. We establish a range of properties for the scheme including monotonicity, consistency, stability, (discrete) regularity, convergence of individual equations, and convergence to the full MFG system.

  1. 1.

    (HJB approximation) For the approximation of the HJB equation we prove pointwise consistency and uniform discrete LL^{\infty}, Lipschitz, and semiconcavity bounds. Convergence to a viscosity solution is obtained via the half relaxed limit method [12].

  2. 2.

    (FKP approximation) We prove consistency in the sense of distributions, preservation of mass and positivity, L1L^{1}-stability, tightness, and equi-continuity in time. In dimension d=1d=1, we also prove uniform LpL^{p}-estimates for all p(1,]p\in(1,\infty]. Convergence is obtained from compactness and stability arguments.

  3. 3.

    (The full MFG approximation) We prove convergence along subsequences to viscosity-very weak solutions of the MFG system in two cases: (i) Degenerate equations in dimension d=1d=1, and (ii) non-degenerate equation in d\mathbb{R}^{d} under the assumption that solutions of the HJB equation are C1C^{1} in space. Full convergence follows for MFGs with unique solutions, and convergence to classical solutions follows under certain regularity and weak uniqueness conditions. Applying the results to the setting of [37], we obtain full convergence to classical solutions in this case.

Because of the nonlocal or smoothing couplings, the HJB approximation can be analysed almost independently of the FKP approximation. The analysis of the FKP scheme on the other hand, strongly depends on boundedness and regularity properties of solutions of the HJB scheme. Compactness in measure is enough in the nondegenerate case when the HJB equation has C1C^{1} solutions, while stronger weak (*) compactness in LpL^{p} for some p(1,]p\in(1,\infty] is needed in the degenerate case. As in [23], we are only able to prove this latter compactness in dimension d=1d=1. A priori estimates and convergence for p(1,)p\in(1,\infty) seems to be new also for local MFGs.

In this paper we study general Lévy jump processes and nonlocal operators. This means that the underlying stochastic processes may not have first moments whatever initial distribution we take (like e.g. σ\sigma-stable processes with σ<1\sigma<1), and then we can no longer work in the commonly used Wasserstein-1 space (P1,d1)(P_{1},d_{1}) for the FKP equations. Instead we work in the space (P,d0)(P,d_{0}) of probability measures under weak convergence metrizised by the Rubinstein-Kantorovich metric d0d_{0} (see Section 2). Surprisingly, a result from [31] (Proposition 6.1) allow us to prove tightness and compactness in this space without any moment assumptions! We refer to section 4.3 for a more detailed discussion along with convergence results in the traditional (P1,d1)(P_{1},d_{1}) topology when first moments are available.

This (P,d0)(P,d_{0}) setting can be adapted to local problems, to give results also there without moment assumptions. Finally, we note that our results for degenerate problems cover the first order equations and improve [23] in the sense that more general initial distributions m0m_{0} are allowed: PLpP\cap L^{p} for some p(1,]p\in(1,\infty] instead of P1+δLP_{1+\delta}\cap L^{\infty} for some δ>0\delta>0.

C. Testing. We provide several numerical simulations. In Example 1 and 2 we use a similar setup as in [24], comparing the effects of a range of different diffusion operators: Fractional Laplacians of different powers, CGMY-diffusions, a degenerate diffusion, a spectrally one-sided diffusion, as well as classical local diffusion and the case of no diffusion. In Example 3 we solve the MFG system on a long time horizon and observe the turnpike property in a nonlocal setting. Finally, in Example 4 we study the convergence of the scheme.

Outline of the paper

In section 2 we list our assumptions and state mostly known results of the MFG system (1) and its individual HJB and FKP equations. In section 3 we construct the discrete schemes for the HJB, FKP, and full MFG equations from the underlying stochastic control problem/game. The convergence results are given in Section 4, along with extensions and a discussion section. In sections 5 and 6 we analyze the discretisations of the HJB and FKP equations respectively, including establishing a priori estimates, stability, and some consistency results. Using these results, we prove the convergence results of section 4 in section 7. In section 8 we provide and discuss numerical simulations of various nonlocal MFG systems. Finally, there are three appendices with proofs of technical results.

2. Assumptions and Preliminaries

We start with some notation. By C,KC,K we mean various constants which may change from line to line. The Euclidean norm on any d\mathbb{R}^{d}-type space is denoted by |||\cdot|. For any subset QdQ\subseteq\mathbb{R}^{d} or Q[0,T]×dQ\subseteq[0,T]\times\mathbb{R}^{d}, and for any bounded, possibly vector valued function on QQ, we will consider LpL^{p}-spaces Lp(Q)L^{p}(Q) and spaces Cb(Q)C_{b}(Q) of bounded continuous functions. Often we use the notation 0\|\cdot\|_{0} as an alternative notation for the norms in CbC_{b} or LL^{\infty}. The space Cbm(Q)C^{m}_{b}(Q) is the subset of Cb(Q)C_{b}(Q) with mm bounded and continuous derivatives, and for Q[0,T]×dQ\subseteq[0,T]\times\mathbb{R}^{d}, Cbl,k(Q)C^{l,k}_{b}(Q) is the subset of Cb(Q)C_{b}(Q) with ll bounded and continuous derivatives in time and kk in space. By P(d)P(\mathbb{R}^{d}) we denote the set of probability measure on d\mathbb{R}^{d}. The Kantorovich-Rubinstein distance d0(μ1,μ2)d_{0}(\mu_{1},\mu_{2}) on the space P(d)P(\mathbb{R}^{d}) is defined as

d0(μ1,μ2):=supfLip1,1(d){df(x)d(μ1μ2)(x)},d_{0}(\mu_{1},\mu_{2}):=\sup_{f\in\mbox{Lip}_{1,1}(\mathbb{R}^{d})}\Big{\{}\int_{\mathbb{R}^{d}}f(x)d(\mu_{1}-\mu_{2})(x)\Big{\}},

where Lip1,1(d)={f:fis Lipschitz continuous andf0,Df01}\mbox{Lip}_{1,1}(\mathbb{R}^{d})=\Big{\{}f:f\,\mbox{is Lipschitz continuous and}\,\|f\|_{0},\|Df\|_{0}\leq 1\Big{\}}. We define the Legendre transform LL of HH as:

L(x,q):=suppd{pqH(x,p)}.\displaystyle L(x,q):=\sup_{p\in\mathbb{R}^{d}}\big{\{}p\cdot q-H(x,p)\big{\}}.

We use the following assumptions for equation (1):

(ν\nu0):

(Lévy condition) ν\nu is a positive Radon measure that satisfies

d1|z|2dν(z)<.\displaystyle\int_{\mathbb{R}^{d}}1\wedge|z|^{2}d\nu(z)<\infty.
(ν\nu1):

(Growth near singularity) There exists constants σ(0,2)\sigma\in(0,2) and C>0C>0 such that the density of ν\nu for |z|<1|z|<1 satisfies

0dνdzC|z|d+σ, for |z|<1.\displaystyle 0\leq\frac{d\nu}{dz}\leq\frac{C}{|z|^{d+\sigma}},\text{ for }|z|<1.
(L0):

(Continuity and local boundedness) The function L:d×dL:\mathbb{R}^{d}\times\mathbb{R}^{d}\to\mathbb{R} is continuous in x,qx,q, and for any K>0K>0, there exists CL(K)>0C_{L}(K)>0 such that

sup|q|K|L(x,q)|CL(K),xd.\displaystyle\sup_{|q|\leq K}|L(x,q)|\leq C_{L}(K),\qquad x\in\mathbb{R}^{d}.
(L1):

(Convexity and growth) The function L(x,q)L(x,q) is convex in qq and satisfies

lim|q|+L(x,q)|q|=+,xd.\displaystyle\lim_{|q|\to+\infty}\frac{L(x,q)}{|q|}=+\infty,\qquad x\in\mathbb{R}^{d}.
(L2):

(Lipschitz regularity) There exists a constant LL>0L_{L}>0 independent of qq, such that

|L(x,q)L(y,q)|LL|xy|.\displaystyle|L(x,q)-L(y,q)|\leq L_{L}|x-y|.
(L3):

(Semi-concavity) There exists a constant cL>0c_{L}>0 independent of qq, such that

L(x+y,q)2L(x,q)+L(xy,q)cL|y|2.\displaystyle L(x+y,q)-2L(x,q)+L(x-y,q)\leq c_{L}|y|^{2}.
(F1):

(Uniform bounds) There exists constants CF,CG>0C_{F},C_{G}>0 such that

|F(x,μ)|CF,|G(x,μ)|CG,xd,μP(d).\displaystyle|F(x,\mu)|\leq C_{F},|G(x,\mu)|\leq C_{G},\qquad\forall x\in\mathbb{R}^{d},\mu\in P(\mathbb{R}^{d}).
(F2):

(Lipschitz assumption) There exists constants LF,LG>0L_{F},L_{G}>0 such that

|F(x,μ1)F(y,μ2)|LF[|xy|+d0(μ1,μ2)],\displaystyle|F(x,\mu_{1})-F(y,\mu_{2})|\leq L_{F}\big{[}|x-y|+d_{0}(\mu_{1},\mu_{2})\big{]},
|G(x,μ1)G(y,μ2)|LG[|xy|+d0(μ1,μ2)].\displaystyle\vskip 6.0pt plus 2.0pt minus 2.0pt|G(x,\mu_{1})-G(y,\mu_{2})|\leq L_{G}\big{[}|x-y|+d_{0}(\mu_{1},\mu_{2})\big{]}.
(F3):

(Semi-concavity) There exists constants cF,cG>0c_{F},c_{G}>0 such that

F(x+y,μ)2F(x,μ)+F(xy,μ)cF\displaystyle F(x+y,\mu)-2F(x,\mu)+F(x-y,\mu)\leq c_{F}
G(x+y,μ)2G(x,μ)+G(xy,μ)cG\displaystyle\vskip 6.0pt plus 2.0pt minus 2.0ptG(x+y,\mu)-2G(x,\mu)+G(x-y,\mu)\leq c_{G}
(M):

(Initial condition) We assume m0P(d)m_{0}\in P(\mathbb{R}^{d}).

(M’):

The dimension d=1d=1, and m0P()Lp()m_{0}\in P(\mathbb{R})\cap L^{p}(\mathbb{R}) for some p(1,]p\in(1,\infty].

By (L1): , the Legendre transform H=LH=L^{*} is welldefined and the optimal qq is q=DpH(x,p)q^{*}=D_{p}H(x,p). To study the convergence of the numerical schemes we further assume local uniform bounds on the derivatives of Hamiltonian:

(H1):

The function DpHC(d×d)D_{p}H\in C(\mathbb{R}^{d}\times\mathbb{R}^{d}), and for every R>0R>0, there is a constant CR>0C_{R}>0 such that for every xdx\in\mathbb{R}^{d} and pBRp\in B_{R} we have |DpH(x,p)|CR|D_{p}H(x,p)|\leq C_{R}.

(H2):

The function DpHC1(d×d)D_{p}H\in C^{1}(\mathbb{R}^{d}\times\mathbb{R}^{d}). For every R>0R>0 there exists a constant CR>0C_{R}>0 such that for every xdx\in\mathbb{R}^{d} and pBRp\in B_{R} we have

|DppH(x,p)|+|DpxH(x,p)|CR.\displaystyle|D_{pp}H(x,p)|+|D_{px}H(x,p)|\leq C_{R}.
Remark 2.1.

We impose most of the conditions on LL, and not on HH, as LL appears in optimal control problem, which would be the basis of our semi-Lagrangian approximation. Assumptions (L1): and (L2): (but, not (L3): !) would immediately carry forward to the corresponding Hamiltonian HH from the definition of Legendre transform. Whereas, we require to assume (H1): (H2): on HH, in contrary to the other assumptions, as it does not follow from the condition on LL in general. However, when the Lagrangian LL behaves like ||r|\cdot|^{r} in qq variable for large qq and r>1r>1, the growth of the corresponding Hamiltonian HH would be ||rr1|\cdot|^{\frac{r}{r-1}} in pp variable for large pp (cf. [32, Proposition 2.1]). The growth of the derivatives of HH for large pp can be computed similarly, which would correspond to similar condition as in (H1): (H2): .

In most of this paper solutions of the HJB equation in (1) are interpreted in the viscosity sense, we refer to [46] and references therein for general definition and wellposedness results, while solutions of FPK equation in (1) are considered in the very weak sense defined as follows:

Definition 2.2.

(a) If uCb0,1((0,T)×d)u\in C^{0,1}_{b}((0,T)\times\mathbb{R}^{d}), a measure mC([0,T],P(d))m\in C([0,T],P(\mathbb{R}^{d})) is a very weak solution of the FPK equation in (1), if for every ϕCc(d)\phi\in C_{c}^{\infty}(\mathbb{R}^{d}) and t[0,T]t\in[0,T]

(3) dϕ(x)𝑑m(t)(x)dϕ(x)𝑑m0(x)=0td([ϕ](x)DpH(x,Du)Dϕ(x))𝑑m(s)(x)𝑑s.\displaystyle\begin{split}&\int_{\mathbb{R}^{d}}\phi(x)\,dm(t)(x)-\int_{\mathbb{R}^{d}}\phi(x)\,dm_{0}(x)\\ &=\int_{0}^{t}\int_{\mathbb{R}^{d}}\Big{(}\mathcal{L}[\phi](x)-D_{p}H(x,Du)\cdot D\phi(x)\Big{)}dm(s)(x)ds.\end{split}

(b) If uL(0,T;W1,(d))u\in L^{\infty}(0,T;W^{1,\infty}(\mathbb{R}^{d})) and p[1,]p\in[1,\infty], a function mC([0,T],P(d))Lp([0,T]×d)m\in C([0,T],P(\mathbb{R}^{d}))\cap L^{p}([0,T]\times\mathbb{R}^{d}), is a very weak solution of the FPK equation in (1), if (3) holds for every ϕCc(d)\phi\in C_{c}^{\infty}(\mathbb{R}^{d}) and t[0,T]t\in[0,T].

Remark 2.3.

Inequality (3) holding for every ϕCc(d)\phi\in C_{c}^{\infty}(\mathbb{R}^{d}) and t[0,T]t\in[0,T] is equivalent to

dϕ(T,x)d(m(T))(x)dϕ(0,x)𝑑m0(x)=0Td(ϕt(s,x)+[ϕ](s,x)DpH(x,Du)Dϕ(s,x))𝑑m(s)(x)𝑑s,\displaystyle\begin{split}&\int_{\mathbb{R}^{d}}\phi(T,x)\,d(m(T))(x)-\int_{\mathbb{R}^{d}}\phi(0,x)\,dm_{0}(x)\\ &=\int_{0}^{T}\int_{\mathbb{R}^{d}}\Big{(}\phi_{t}(s,x)+\mathcal{L}[\phi](s,x)-D_{p}H(x,Du)\cdot D\phi(s,x)\Big{)}dm(s)(x)ds,\end{split}

holding for every ϕCb1,2([0,T]×d)\phi\in C^{1,2}_{b}([0,T]\times\mathbb{R}^{d}) (cf. [31, Lemma 6.1]).

Definition 2.4.

A pair (u,m)(u,m) is a viscosity-very weak solution of the MFG system (1), if uu is a viscosity solution of the HJB equation, and mm is a very weak solution of the FPK equation (see, Definition 2.2).

Proposition 2.5.

Fix, μC([0,T],P(d))\mu\in C([0,T],P(\mathbb{R}^{d})). Let (ν\nu0): , (L2): and (F1): hold.

(a) (Comparison principle) If uu is a viscosity subsolution and vv is a viscosity supersolution of the HJB equation in (1) with u(T,)v(T,)u(T,\cdot)\leq v(T,\cdot), then uvu\leq v.

(b) There exists a unique bounded viscosity solution uCb([0,T]×d)u\in C_{b}([0,T]\times\mathbb{R}^{d}) of the HJB equation in (1), and for any t[0,T]t\in[0,T] we have u(t)0CFT+CG\|u(t)\|_{0}\leq C_{F}T+C_{G}.

(c) If (L2): and (F2): hold, then the viscosity solution uu is Lipschitz continuous in space variable and for every t[0,T]t\in[0,T] and x,ydx,y\in\mathbb{R}^{d} we have

|u(t,x)u(t,x+y)|(T(LL+LF)+LG)|y|.\displaystyle|u(t,x)-u(t,x+y)|\leq\big{(}T(L_{L}+L_{F})+L_{G}\big{)}\,|y|.

In addition, if (L3): and (F3): hold, then uu is semiconcave in space variable and for every t[0,T]t\in[0,T] and x,ydx,y\in\mathbb{R}^{d} we have

u(t,x+y)+u(t,xy)2u(t,x)(T(cL+cF)+cG)|y|2.\displaystyle u(t,x+y)+u(t,x-y)-2u(t,x)\leq\big{(}T(c_{L}+c_{F})+c_{G}\big{)}\,|y|^{2}.
Proof.

These results are by now standard: (a) follows by a similar argument as for [46, Theorem 3.1], (b) follows by e.g. Perron’s method, and (c) by adapting the comparison arguments of [46] in a standard way. We omit the details. Under some extra assumptions, (b) and (c) also follows from Theorem 5.4 and Lemma 5.3 below. ∎

Proposition 2.6.

(a) If uC([0,T];Cb1(d))u\in C([0,T];C^{1}_{b}(\mathbb{R}^{d})), then there exists a very weak solution mC([0,T];P(d))m\in C([0,T];P(\mathbb{R}^{d})) of the FPK equation in (1).

(b) If d=1d=1, uC([0,T];W1,())u\in C([0,T];W^{1,\infty}(\mathbb{R})), uu semi-concave, and (M’): holds, then there exists a very weak solution mC([0,T];P())Lp([0,T]×)m\in C([0,T];P(\mathbb{R}))\cap L^{p}([0,T]\times\mathbb{R}) of the FPK equation in (1). Moreover, m(t)Lp()eCTm0Lp()\|m(t)\|_{L^{p}(\mathbb{R})}\leq e^{CT}\|m_{0}\|_{L^{p}(\mathbb{R})} for some constant C>0C>0 and t[0,T]t\in[0,T].

Proof.

The results follow from the convergence of the discrete scheme in this article. The proof of (a) follows the proof of Theorem 4.3, setting Duρ,h=DuDu_{\rho,h}=Du. The proof of (b) follows the proof of Theorem 4.1 and Theorem 6.7, setting Duρ,h=DuDu_{\rho,h}=Du. Note that semi-concavity of uu is crucial for the the LpL^{p}-bound of Theorem 6.7. ∎

Existence and uniqueness results are given in [37] for classical solutions of MFGs with nonlocal diffusions under additional assumptions:

(ν\nu2):

(Growth near singularity) There exists constants σ(1,2)\sigma\in(1,2) and c>0c>0 such that the density of ν\nu for |z|<1|z|<1 satisfies

c|z|d+σdνdz, for |z|<1.\displaystyle\frac{c}{|z|^{d+\sigma}}\leq\frac{d\nu}{dz},\text{ for }|z|<1.
(F4):

There exists constants CF,CG>0C_{F},C_{G}>0, such that F(,m)Cb2CF\|F(\cdot,m)\|_{C_{b}^{2}}\leq C_{F} and G(,m~)Cb3CG\|G(\cdot,\tilde{m})\|_{C_{b}^{3}}\leq C_{G} for all m,m~P(d)m,\tilde{m}\in P(\mathbb{R}^{d}).

(F5):

FF and GG satisfy monotonicity conditions:

d(F(x,m1)F(x,m2))d(m1m2)(x)\displaystyle\int_{\mathbb{R}^{d}}\left(F\left(x,m_{1}\right)-F\left(x,m_{2}\right)\right)d\left(m_{1}-m_{2}\right)\left(x\right) 0m1,m2P(d),\displaystyle\geq 0\qquad\forall m_{1},m_{2}\in P(\mathbb{R}^{d}),
d(G(x,m1)G(x,m2))d(m1m2)(x)\displaystyle\int_{\mathbb{R}^{d}}\left(G\left(x,m_{1}\right)-G\left(x,m_{2}\right)\right)d\left(m_{1}-m_{2}\right)\left(x\right) 0m1,m2P(d).\displaystyle\geq 0\qquad\forall m_{1},m_{2}\in P(\mathbb{R}^{d}).
(H3):

The Hamiltonian HC3(d×d)H\in C^{3}(\mathbb{R}^{d}\times\mathbb{R}^{d}), and for every R>0R>0 there is CR>0C_{R}>0 such that for xdx\in\mathbb{R}^{d}, pBRp\in B_{R}, α0N\alpha\in\mathbb{N}_{0}^{N}, |α|3|\alpha|\leq 3, then |DαH(x,p)|CR|D^{\alpha}H(x,p)|\leq C_{R}.

(H4):

For every R>0R>0 there is CR>0C_{R}>0 such that for x,yd,u[R,R],pdx,y\in\mathbb{R}^{d},u\in\left[-R,R\right],p\in\mathbb{R}^{d}: |H(x,u,p)H(y,u,p)|CR(|p|+1)|xy||H\left(x,u,p\right)-H\left(y,u,p\right)|\leq C_{R}\left(|p|+1\right)|x-y|.

(H5):

(Uniform convexity) There exists a constant C>0C>0 such that 1CIdDpp2H(x,p)CId\frac{1}{C}I_{d}\leq D_{pp}^{2}H\left(x,p\right)\leq CI_{d}.

(M”):

The probability measure m0m_{0} has a density (also denoted by m0m_{0}) m0Cb2m_{0}\in C_{b}^{2}.

Theorem 2.7.

(a) There exists a classical solution (u,m)(u,m) of (1) such that uCb1,3((0,T)×d)u\in C^{1,3}_{b}((0,T)\times\mathbb{R}^{d}) and mCb1,2((0,T)×d)C(0,T;P(d))m\in C^{1,2}_{b}((0,T)\times\mathbb{R}^{d})\cap C(0,T;P(\mathbb{R}^{d})).

(b) If in addition (F5): and (H5): hold, then the classical solution is unique.

This is a consequence of [37, Theorem 2.5 and Theorem 2.6]. We refer to [37] for more general results, where in particular assumptions (ν\nu1): and (ν\nu2): can be relaxed to allow for a much larger class of nonlocal operators \mathcal{L}. In the nondegenerate case, for the individual equations in (1) we also have uniqueness of viscosity-very weak solutions and existence of classical solutions. Uniqueness for HJB equations and existence for HJB and FPK equations follows by Theorem 5.3, Theorem 5.5, and Proposition 6.8 in [37]. We prove uniqueness for very weak solutions of FPK equations here.

Proposition 2.8 (Uniqueness for the FPK equation).

Assume (ν\nu0): , (ν\nu1): , (ν\nu2): , and DpH(x,Du(t))Cb0,2((0,T)×d)D_{p}H(x,Du(t))\in C_{b}^{0,2}((0,T)\times\mathbb{R}^{d}). Then there is at most one very weak solution of the FPK equation in (1).

Proof.

Let m1,m2m_{1},m_{2} be two very weak solutions, define m~:=m1m2\tilde{m}:=m_{1}-m_{2} and take any ψCc(d)\psi\in C_{c}^{\infty}\left(\mathbb{R}^{d}\right). For any τ(0,T)\tau\in(0,T), the terminal value problem

tϕ+ϕDϕDpH(x,Du)=0ind×(0,τ)andϕ(x,τ)=ψ(x)ind,\displaystyle\partial_{t}\phi+\mathcal{L}\phi-D\phi\cdot D_{p}H(x,Du)=0\quad\text{in}\quad\mathbb{R}^{d}\times(0,\tau)\quad\text{and}\quad\phi(x,\tau)=\psi(x)\quad\text{in}\quad\mathbb{R}^{d},

has a unique classical solution ϕCb1,2((0,τ)×d)\phi\in C_{b}^{1,2}((0,\tau)\times\mathbb{R}^{d}) essentially by [37, Theorem 5.5] (the result follows from Proposition 5.8 with k=2k=2 and the observation that the proof of Theorem 5.5 also holds for k=2k=2). Using the definition of very weak solution (see Remark 2.3) we get

dψ(x)𝑑m~(τ)(x)=0τd(tϕ+ϕDϕDpH(x,Du))𝑑m~(t)(x)𝑑t=0,\displaystyle\int_{\mathbb{R}^{d}}\psi(x)\,d\tilde{m}(\tau)(x)=\int_{0}^{\tau}\int_{\mathbb{R}^{d}}\big{(}\partial_{t}\phi+\mathcal{L}\phi-D\phi\cdot D_{p}H(x,Du)\big{)}\,d\tilde{m}(t)(x)\,dt=0,

for any τ[0,T]\tau\in[0,T]. Since ψ\psi was arbitrary, it follows that m~(τ)=0\tilde{m}(\tau)=0 in P(d)P(\mathbb{R}^{d}) for every τ[0,T]\tau\in[0,T], and uniqueness follows. ∎

3. Discretisation of the MFG system

To discretise the MFG system (1), we first follow [15] and derive a Semi-Lagrange approximation of the HJB equation in (1). Using this approximation and the optimal control of the original problem, we derive an approximation of the FPK equation in (1) which is in (approximate) duality with the approximation of the HJB-equation.

This derivation is based on the following control interpretation of the HJB equation. For a fixed given density m=μm=\mu, the solution uu of the HJB equation in (1) is the value function of the optimal stochastic control problem:

(4) u(t,x)=infαJ(x,t,α),\displaystyle u(t,x)=\inf_{\alpha}J\big{(}x,t,\alpha\big{)},

where αt\alpha_{t} is an admissible control, JJ is the total cost to be minimized,

(5) J(x,t,α)=𝔼[tT(L(X~s,αs)+F(X~s,μs)ds+G(X~T,μT)],\displaystyle J\big{(}x,t,\alpha\big{)}=\mathbb{E}\bigg{[}\int_{t}^{T}\Big{(}L(\tilde{X}_{s},\alpha_{s})+F(\tilde{X}_{s},\mu_{s}\Big{)}ds+G(\tilde{X}_{T},\mu_{T})\bigg{]},

and X~s=X~sx,t\tilde{X}_{s}=\tilde{X}_{s}^{x,t} solves the controlled stochastic differential equation (SDE)

(6) {dX~s=αsds+|z|<1zN~(dz,ds)+|z|1zN(dz,ds),s>t,X~t=x,\displaystyle\begin{cases}&d\tilde{X}_{s}=-\alpha_{s}\,ds+\int_{|z|<1}z\tilde{N}(dz,ds)+\int_{|z|\geq 1}zN(dz,ds),\quad s>t,\\ &\tilde{X}_{t}=x,\end{cases}

where NN a Poisson random measure with intensity/Lévy measure ν(dz)ds\nu(dz)ds, and N~=N(dz,ds)ν(dz)ds\tilde{N}=N(dz,ds)-\nu(dz)ds is the compensated Poisson measure.111The NN-integral is just a (difficult way of writing a) compound Poisson jump-process, while the N~\tilde{N}-integral is a centered jump process with an infinite number of (small) jumps per time interval a.s. [9].

3.1. Approximation of the underlying controlled SDE

A. Approximate small jumps by Brownian motion.

First we approximate small jumps in (6) by (vanishing) Brownian motion222To avoid singular integrals and infinite number of jumps per time interval. (cf. [10]): For r(0,1)r\in(0,1), let Xs=Xsx,tX_{s}=X_{s}^{x,t} solve

(7) {dXs=b¯(αs)ds+σrdWs+|z|rzN(dz,ds),s>tXt=x,\displaystyle\begin{cases}dX_{s}=\bar{b}(\alpha_{s})ds+\sigma_{r}\,dW_{s}+\int_{|z|\geq r}zN(dz,ds),\quad s>t\\ X_{t}=x,\end{cases}

where WsW_{s} is a standard Brownian motion, b¯(αs)=αsbrσ\bar{b}(\alpha_{s})=-\,\alpha_{s}-b_{r}^{\sigma}, and

(8) brσ:=r<|z|<1zν(dz),\displaystyle b_{r}^{\sigma}:=\int_{r<|z|<1}z\,\nu(dz),
(9) σr:=(12|z|<rzzTν(dz))1/2.\displaystyle\sigma_{r}:=\bigg{(}\frac{1}{2}\int_{|z|<r}zz^{T}\nu(dz)\bigg{)}^{1/2}.

The last integral in (7) is a compound Poisson process (cf. e.g. [9]): For any t0t\geq 0,

(10) 0t|z|rzN(dz,dt)=j=1N^tJj\displaystyle\int_{0}^{t}\int_{|z|\geq r}zN(dz,dt)=\sum_{j=1}^{\hat{N}_{t}}J_{j}

where the number of jumps up to time tt is N^tPoisson(tλr)\hat{N}_{t}\sim\textup{Poisson}(t\lambda_{r}), the jumps {Jj}j\{J_{j}\}_{j} are iid rv’s in d\mathbb{R}^{d} with distribution νr\nu_{r} and J0=0J_{0}=0, and for r(0,1]r\in(0,1],

(11) νr:=ν𝟙|z|>randλr:=dνr(dz).\displaystyle\nu_{r}:=\nu\mathbbm{1}_{|z|>r}\qquad\text{and}\qquad\lambda_{r}:=\int_{\mathbb{R}^{d}}\nu_{r}(dz).

The infinitesimal generators α\mathcal{L}^{\alpha} and ^α\hat{\mathcal{L}}^{\alpha} of the SDEs (6) and (7) are (cf. [9])

αϕ(x)=αtϕ+1ϕ(x)+1ϕ(x),\displaystyle\mathcal{L}^{\alpha}\phi(x)=-\alpha_{t}\cdot\nabla\phi+\mathcal{L}_{1}\phi(x)+\mathcal{L}^{1}\phi(x),
^αϕ(x)=b¯(αt)ϕ(x)+tr(σrTD2ϕ(x)σr)+r[ϕ](x)\displaystyle\hat{\mathcal{L}}^{\alpha}\phi(x)=\,\bar{b}(\alpha_{t})\cdot\nabla\phi(x)+tr\big{(}\sigma_{r}^{T}\cdot D^{2}\phi(x)\cdot\sigma_{r}\big{)}+\mathcal{L}^{r}[\phi](x)

for ϕCb2(d)\phi\in C^{2}_{b}(\mathbb{R}^{d}), where

(12) ϕ(x)=rϕ(x)+rϕ(x):=(|z|<r+|z|>r)(ϕ(x+z)ϕ(x)𝟙{|z|<1}Dϕ(x)z)dν(z).\displaystyle\begin{split}&\mathcal{L}\phi(x)=\mathcal{L}_{r}\phi(x)+\mathcal{L}^{r}\phi(x)\\ &:=\bigg{(}\int_{|z|<r}+\int_{|z|>r}\bigg{)}\Big{(}\phi(x+z)-\phi(x)-\mathbbm{1}_{\{|z|<1\}}D\phi(x)\cdot z\Big{)}d\nu(z).\end{split}

The operator ^α\hat{\mathcal{L}}^{\alpha} is an approximation of α\mathcal{L}^{\alpha}.

Lemma 3.1 ([47]).

If (ν\nu1): holds and ϕCb3(d)\phi\in C_{b}^{3}(\mathbb{R}^{d}), then for r\mathcal{L}_{r} and σr\sigma_{r} defined in (12) and (9) respectively, we have

|rϕ(x)tr(σrTD2ϕ(x)σr)|Cr3σD3ϕ0.\displaystyle|\mathcal{L}_{r}\phi(x)-tr\big{(}\sigma_{r}^{T}\cdot D^{2}\phi(x)\cdot\sigma_{r}\big{)}|\leq Cr^{3-\sigma}\|D^{3}\phi\|_{0}.

If in addition, ϕCb4(d)\phi\in C_{b}^{4}(\mathbb{R}^{d}) and the Lévy measure ν\nu is symmetric, then

|rϕ(x)tr(σrTD2ϕ(x)σr)|Cr4σD4ϕ0.\displaystyle|\mathcal{L}_{r}\phi(x)-tr\big{(}\sigma_{r}^{T}\cdot D^{2}\phi(x)\cdot\sigma_{r}\big{)}|\leq Cr^{4-\sigma}\|D^{4}\phi\|_{0}.

B. Time discretization of the approximate SDE

Fix a time step h=TN(0,1)h=\frac{T}{N}\in(0,1) for some NN\in\mathbb{N} and discrete times tk=kht_{k}=kh for k{0,1,,N}k\in\{0,1,\dots,N\}. Following [15], we propose the following Euler-Maruyama discretization of the SDE (7): Let Xntl,xXtntl,xX_{n}^{t_{l},x}\approx X^{t_{l},x}_{t_{n}}, where Xn=Xntl,xX_{n}=X^{t_{l},x}_{n}solves

(13) {Xl=xXn=Xn1+hb¯(αn1)+hm=1dσrmξn1m,n=l+Ni+1,,l+Ni+11,Xl+Ni+1=Xl+Ni+11+Ji.\displaystyle\begin{cases}X_{l}=x\\ X_{n}=X_{n-1}+h\bar{b}(\alpha_{n-1})+\sqrt{h}\displaystyle\sum_{m=1}^{d}\sigma_{r}^{m}\xi_{n-1}^{m},\ \ n=l+N_{i}+1,\dots,l+N_{i+1}-1,\\ X_{l+N_{i+1}}=X_{l+N_{i+1}-1}+J_{i}.\end{cases}

Here the control αn\alpha_{n} is constant on each time interval, σrm\sigma_{r}^{m} is the mmth-column of σr\sigma_{r}, and ξn=(ξn1,,ξnd)\xi_{n}=(\xi_{n}^{1},\ldots,\xi_{n}^{d}) is a random walk in d\mathbb{R}^{d} with

𝐏(ξni=±1)=12d.\displaystyle\mathbf{P}\big{(}\xi^{i}_{n}=\pm 1\big{)}=\frac{1}{2d}.

The processes JkJ_{k} and NkN_{k} defines an approximation of the compound Poisson part of (7) through equation (10) where N^t\hat{N}_{t} is replaced by an approximation

N~t=max{k:ΔT1+ΔT2++ΔTkt},\tilde{N}_{t}=\max\{k:\Delta T_{1}+\Delta T_{2}+\dots+\Delta T_{k}\leq t\},

where exponentially distributed waiting times (time between jumps) are replaced by approximations {ΔTk}k\{\Delta T_{k}\}_{k\in\mathbb{N}}333In the new model, N~t\tilde{N}_{t} still gives the number of jumps up to time tt.: ΔTk=hΔNk=h(NkNk1)\Delta T_{k}=h\Delta N_{k}=h(N_{k}-N_{k-1}) where Nk:Ω{0}N_{k}:\Omega\to\mathbb{N}\cup\{0\}, N0=0N_{0}=0, and ΔNk\Delta N_{k} iid with approximate hλrh\lambda_{r}-exponential distribution given by

𝐏[ΔNk>j]=ehλrjforj=0,1,2,.\displaystyle\mathbf{P}[\Delta N_{k}>j]=e^{-h\lambda_{r}j}\quad\mbox{for}\quad j=0,1,2,\dots.

Then for pj:=P[ΔNk=j]p_{j}:=P[\Delta N_{k}=j], p0=0p_{0}=0 and pj=P[ΔNk>j1]P[ΔNk>j]=ejhλr(ehλr1)p_{j}=P[\Delta N_{k}>j-1]-P[\Delta N_{k}>j]=e^{-jh\lambda_{r}}(e^{h\lambda_{r}}-1) for j>0j>0. We find that j=0pj=1\sum_{j=0}^{\infty}p_{j}=1 and E(ΔNk)=j=0ejhλr=ehλrehλr1E(\Delta N_{k})=\sum_{j=0}^{\infty}e^{-jh\lambda_{r}}=\frac{e^{h\lambda_{r}}}{e^{h\lambda_{r}}-1}. Note that in each time interval, approximation (13) either diffuses (the second equation) or jumps (the third equation), and that we have ignored the unlikely event of more than one jump per time interval. For the scheme to converge, we will see that we need to send both h0h\to 0 and hλr0h\lambda_{r}\to 0. In this case E(ΔNk)E(\Delta N_{k})\to\infty and the jumps become less and less frequent and the random walk dominates the evolution of XkX_{k} (which is to be expected).

3.2. Semi-Lagrangian approximation of the HJB equation

A. Control approximation of the HJB equation

We approximate the control problem (4) – (6) by a discrete time control problem: Define the value function

(14) u~h(tl,x)=inf{αn}Jh(x,tl,{αn}),\displaystyle\tilde{u}_{h}(t_{l},x)=\inf_{\{\alpha_{n}\}}J_{h}\big{(}x,t_{l},\{\alpha_{n}\}\big{)},

where the controls {αn}\{\alpha_{n}\} are piecewise constant in time, the cost function JhJ_{h} is given by

(15) Jh(x,tl,{αn})=𝔼[n=lN1(L(Xn,αn)+F(Xn,μ(tn)))h+G(XN,μ(tN))],\displaystyle J_{h}\big{(}x,t_{l},\{\alpha_{n}\}\big{)}=\mathbb{E}\bigg{[}\sum_{n=l}^{N-1}\Big{(}L(X_{n},\alpha_{n})+F(X_{n},\mu(t_{n}))\Big{)}h+G(X_{N},\mu(t_{N}))\bigg{]},

and the controlled discrete time process Xn=Xntl,xX_{n}=X_{n}^{t_{l},x} is the solution of (13). By the (discrete time) Dynamic Programming principle it follows that

u~h(tl,x)=infαn𝔼[n=ll+p(L(Xntl,x,αn)+F(Xntl,x,μ(tn)))h+u~h(tl+p+1,Xl+p+1tl,x)],\displaystyle\tilde{u}_{h}(t_{l},x)=\inf_{\alpha_{n}}\mathbb{E}\bigg{[}\sum_{n=l}^{l+p}\Big{(}L(X_{n}^{t_{l},x},\alpha_{n})+F(X_{n}^{t_{l},x},\mu(t_{n}))\Big{)}h+\tilde{u}_{h}(t_{l+p+1},X_{l+p+1}^{t_{l},x})\bigg{]},

for l+p+1Nl+p+1\leq N. Taking p=0p=0 and computing the expectation using conditional probabilities (the probability to jump in a time interval is p1=1ehλrp_{1}=1-e^{-h\lambda_{r}}), we find a (discrete time) HJB equation

u~h(tl,x\displaystyle\tilde{u}_{h}(t_{l},x )=infα{hF(x,μ(tl))+hL(x,α)+[ehλr2dm=1d(u~h(tl+h,x+hb¯(α)+hdσrm)\displaystyle)=\inf_{\alpha}\bigg{\{}hF(x,\mu(t_{l}))+hL(x,\alpha)+\Big{[}\frac{e^{-h\lambda_{r}}}{2d}\sum_{m=1}^{d}\big{(}\tilde{u}_{h}(t_{l}+h,x+h\bar{b}(\alpha)+\sqrt{hd}\sigma_{r}^{m})
(16) +u~h(tl+h,x+hb¯(α)hdσrm))+1ehλrλr|z|ru~h(tl+h,x+z)ν(dz)]}.\displaystyle+\tilde{u}_{h}(t_{l}+h,x+h\bar{b}(\alpha)-\sqrt{hd}\sigma_{r}^{m})\big{)}+\frac{1-e^{-h\lambda_{r}}}{\lambda_{r}}\int_{|z|\geq r}\tilde{u}_{h}(t_{l}+h,x+z)\nu(dz)\Big{]}\bigg{\}}.

B. Interpolation and the fully discrete scheme

For ρ>0\rho>0 we fix a grid 𝒢ρ={iρ:id}\mathcal{G}_{\rho}=\{i\rho:i\in\mathbb{Z}^{d}\} and a linear/multilinear 𝒢ρ\mathcal{G}_{\rho}-interpolation II. For functions f:𝒢ρf:\mathcal{G}_{\rho}\rightarrow\mathbb{R},

(17) I[f](x):=idf(xi)βi(x),xd,\displaystyle I[f](x):=\sum_{i\in\mathbb{Z}^{d}}f(x_{i})\beta_{i}(x),\qquad x\in\mathbb{R}^{d},

where the βj\beta_{j}’s are piecewise linear/multilinear basis functions satisfying

βj0,βj(xi)=δj,i,jβj(x)=1,andI[ϕ]ϕ0=D2ϕ0ρ2\displaystyle\beta_{j}\geq 0,\quad\beta_{j}(x_{i})=\delta_{j,i},\quad\sum_{j}\beta_{j}(x)=1,\quad\text{and}\quad\|I[\phi]-\phi\|_{0}=\|D^{2}\phi\|_{0}\rho^{2}

for any ϕCb2(d)\phi\in C^{2}_{b}(\mathbb{R}^{d}). A fully discrete scheme is then obtained from (16) as follows:

(18) u~i,k[μ]=Sρ,h,r[μ](u~,k+1,i,k),k<N,andu~i,N[μ]=G(xi,μ(tN)),\displaystyle\tilde{u}_{i,k}[\mu]=S_{\rho,h,r}[\mu](\tilde{u}_{\cdot,k+1},i,k),\ k<N,\quad\text{and}\quad\tilde{u}_{i,N}[\mu]=G(x_{i},\mu(t_{N})),

where

Sρ,h,r[μ]\displaystyle S_{\rho,h,r}[\mu] (v,i,k)=infα{hF(xi,μ(tk))+hL(xi,α)+1ehλrλr|z|rI[v](xi+z)ν(dz)\displaystyle(v,i,k)=\inf_{\alpha}\Bigg{\{}hF(x_{i},\mu(t_{k}))+hL(x_{i},\alpha)+\frac{1-e^{-h\lambda_{r}}}{\lambda_{r}}\int_{|z|\geq r}I[v](x_{i}+z)\nu(dz)
(19) +ehλr2dm=1d(I[v](xi+hb¯(α)+hdσrm)+I[v](xi+hb¯(α)hdσrm))}.\displaystyle+\frac{e^{-h\lambda_{r}}}{2d}\sum_{m=1}^{d}\Big{(}I[v](x_{i}+h\bar{b}(\alpha)+\sqrt{hd}\sigma_{r}^{m})+I[v](x_{i}+h\bar{b}(\alpha)-\sqrt{hd}\sigma_{r}^{m})\Big{)}\Bigg{\}}.

Finally, we extend the solution of the discrete scheme u~i,k[μ]\tilde{u}_{i,k}[\mu] to the whole d×[0,T]\mathbb{R}^{d}\times[0,T] by linear interpolation in xx and piecewise constant interpolation in tt:

(20) u~ρ,h[μ](t,x)=I(u~,[th][μ])(x)=idβi(x)u~i,[th][μ]for any(t,x)[0,T)×d.\displaystyle\tilde{u}_{\rho,h}[\mu](t,x)=I\big{(}\tilde{u}_{\cdot,[\frac{t}{h}]}[\mu]\big{)}(x)=\sum_{i\in\mathbb{Z}^{d}}\beta_{i}(x)\,\tilde{u}_{i,[\frac{t}{h}]}[\mu]\quad\mbox{for any}\quad(t,x)\in[0,T)\times\mathbb{R}^{d}.

3.3. Approximate optimal feedback control

For the HJB equation in (1), satisfied by the value function (4), it easily follows that the optimal feedback control is

α(t,x)=DpH(x,Du[μ](t,x)).\alpha(t,x)=D_{p}H(x,Du[\mu](t,x)).

Based on this feedback law, we define an approximate feedback control for the discrete time optimal control problem (13)–(15) in the following way: For h,ρ,ϵ>0h,\rho,\epsilon>0 and (t,x)d×[0,T](t,x)\in\mathbb{R}^{d}\times[0,T],

(21) αnum(t,x):=DpH(x,Du~ρ,hϵ[μ](t,x)),\displaystyle\alpha_{\text{num}}(t,x):=D_{p}H(x,D\tilde{u}^{\epsilon}_{\rho,h}[\mu](t,x)),

where u~ρ,h[μ]\tilde{u}_{\rho,h}[\mu] is given by (20),

(22) u~ρ,hϵ[μ](t,x)=u~ρ,h[μ](t,)ρϵ(x),\displaystyle\tilde{u}_{\rho,h}^{\epsilon}[\mu](t,x)=\tilde{u}_{\rho,h}[\mu](t,\cdot)*\rho_{\epsilon}(x),

and the mollifier ρϵ(x)=1ϵdρ(xϵ)\rho_{\epsilon}(x)=\frac{1}{\epsilon^{d}}\rho\big{(}\frac{x}{\epsilon}\big{)} for 0ρCc(d)0\leq\rho\in C_{c}^{\infty}(\mathbb{R}^{d}) with dρ(x)𝑑x=1\int_{\mathbb{R}^{d}}\rho(x)dx=1. We state a standard result on mollification.

Lemma 3.2.

If uW1,(d)u\in W^{1,\infty}(\mathbb{R}^{d}), ϵ>0\epsilon>0, and uϵ=uρϵu^{\epsilon}=u*\rho_{\epsilon}. Then uϵCb(d)u^{\epsilon}\in C_{b}^{\infty}(\mathbb{R}^{d}), and there exists a constant cρ>0,c_{\rho}>0, such that for all ϵ>0\epsilon>0,

uϵu0Du0ϵandDpuϵ0cρDu0ϵ1pfor anyp.\displaystyle\|u^{\epsilon}-u\|_{0}\leq\|Du\|_{0}\,\epsilon\qquad\mbox{and}\qquad\|D^{p}u^{\epsilon}\|_{0}\leq c_{\rho}\|Du\|_{0}\,\epsilon^{1-p}\ \ \text{for any}\ \ p\in\mathbb{N}.

By construction, we expect αnum\alpha_{\text{num}} to be an approximation of the optimal feedback control for the approximate control problem with value function (14) when h,ρ,ϵh,\rho,\epsilon are small and u~ρ,hϵ\tilde{u}^{\epsilon}_{\rho,h} is close to uu.

3.4. Dual SL discretization of the FPK equation

A. Dual approximation of the FPK equation

First note that if X~s=X~s0,Z0\tilde{X}_{s}=\tilde{X}^{0,Z_{0}}_{s} solves (6) with t=0t=0 and X0=Z0X_{0}=Z_{0}, a rv with distribution m0m_{0}, then the FPK equation for m~:=Law(X~s)\tilde{m}:=Law(\tilde{X}_{s}) is

{m~tm~div(m~α)=0,m~(0)=m0.\displaystyle\begin{cases}\tilde{m}_{t}-\mathcal{L}^{*}\tilde{m}-\text{div}(\tilde{m}\alpha)=0,\\ \tilde{m}(0)=m_{0}.\end{cases}

Setting α=αnum\alpha=\alpha_{\text{num}}, this equation becomes an approximation of the FPK equation in (1). With this choice of α\alpha, we further approximate m~\tilde{m} by the density m~k:=Law(Xk)\tilde{m}_{k}:=Law(X_{k}), of the approximate process Xk=Xk0,Z0X_{k}=X_{k}^{0,Z_{0}} solving (13) with l=0l=0 and X0=Z0X_{0}=Z_{0}.

We now derive a FPK equation for m~k\tilde{m}_{k} which in discretised form will serve as our approximation of the FPK equation in (1). To simplify we consider dimension d=1d=1. By definition of m~k\tilde{m}_{k},

𝔼[ϕ(Xk+1)]=ϕ(x)𝑑m~k+1(x),\displaystyle\mathbb{E}[\phi(X_{k+1})]=\int_{\mathbb{R}}\phi(x)\,d\tilde{m}_{k+1}(x),

for ϕCb(d)\phi\in C_{b}(\mathbb{R}^{d}) and k{0}k\in\mathbb{N}\cup\{0\}. Let AkA_{k} be the event of at least one jump in [tk,tk+1)[t_{k},t_{k+1}), i.e. Ak={ω:Nk+1(ω)Nk(ω)1}A_{k}=\{\omega:N_{k+1}(\omega)-N_{k}(\omega)\geq 1\} where NkN_{k} is the random jump time defined in Section 3.1 B. Then by the definition of XkX_{k} in (13), the fact that NkN_{k}, JkJ_{k}, and ξk\xi_{k} are i.i.d. and hence independent of XkX_{k}, and conditional expectations, we find that

ϕ(x)𝑑m~k+1(x)=𝔼[ϕ(Xk+1)]\displaystyle\int_{\mathbb{R}}\phi(x)\,d\tilde{m}_{k+1}(x)=\mathbb{E}[\phi(X_{k+1})]
=𝔼[ϕ(Xk+1)|Akc]P(Akc)+𝔼[ϕ(Xk+1)|Ak]P(Ak)\displaystyle=\mathbb{E}[\phi(X_{k+1})|A_{k}^{c}]\,P(A_{k}^{c})+\mathbb{E}[\phi(X_{k+1})|A_{k}]\,P(A_{k})
=ehλr𝔼(ϕ(Xk+hb¯(αnum)+hσrξk))+(1ehλr)𝔼(ϕ(Xk+Ji))\displaystyle=e^{-h\lambda_{r}}\mathbb{E}(\phi(X_{k}+h\bar{b}(\alpha_{\text{num}})+\sqrt{h}\sigma_{r}\xi_{k}))+(1-e^{-h\lambda_{r}})\mathbb{E}(\phi(X_{k}+J_{i}))
=ehλr2(ϕ(x+hb¯(αnum)+hσr)+ϕ(x+hb¯(αnum)hσr))m~k(dx)\displaystyle=\frac{e^{-h\lambda_{r}}}{2}\int_{\mathbb{R}}\big{(}\phi(x+h\bar{b}(\alpha_{\text{num}})+\sqrt{h}\sigma_{r})+\phi(x+h\bar{b}(\alpha_{\text{num}})-\sqrt{h}\sigma_{r})\big{)}\tilde{m}_{k}(dx)
+(1ehλr)|z|>rϕ(x+z)ν(dz)λrm~k(dx).\displaystyle\qquad+(1-e^{-h\lambda_{r}})\int_{\mathbb{R}}\int_{|z|>r}\phi(x+z)\frac{\nu(dz)}{\lambda_{r}}\tilde{m}_{k}(dx).

Let Ei:=(xiρ2,xi+ρ2)E_{i}:=\big{(}x_{i}-\frac{\rho}{2},x_{i}+\frac{\rho}{2}\big{)}, m~j,k=Ejm~k(dx)\tilde{m}_{j,k}=\int_{E_{j}}\tilde{m}_{k}(dx). We approximate the above expression by a midpoint (quadrature) approximation, i.e. Ejf(x)m~k(dx)f(xj)m~j,k\int_{E_{j}}f(x)\tilde{m}_{k}(dx)\approx f(x_{j})\tilde{m}_{j,k}, then by choosing ϕ(x)=βj(x)\phi(x)=\beta_{j}(x) (linear interpolant) for jj\in\mathbb{Z} and using βj(xi)=δj,i\beta_{j}(x_{i})=\delta_{j,i} we get a fully discrete approximation

m~j,k+1i\displaystyle\tilde{m}_{j,k+1}\approx\sum_{i\in\mathbb{Z}} m~i,k[ehλr2(βj(xi+hb¯(αnum)+hσr)+βj(xi+hb¯(αnum)hσr))\displaystyle\tilde{m}_{i,k}\Big{[}\frac{e^{-h\lambda_{r}}}{2}\Big{(}\beta_{j}(x_{i}+h\bar{b}(\alpha_{\text{num}})+\sqrt{h}\sigma_{r})+\beta_{j}(x_{i}+h\bar{b}(\alpha_{\text{num}})-\sqrt{h}\sigma_{r})\Big{)}
+1ehλrλr|z|>rβj(xi+z)ν(dz)].\displaystyle\qquad+\frac{1-e^{-h\lambda_{r}}}{\lambda_{r}}\int_{|z|>r}\beta_{j}(x_{i}+z)\nu(dz)\Big{]}.

In arbitrary dimension dd, we denote

(23) Φj,k,pϵ,±:=xjh(Hp(xj,Du~ρ,hϵ[μ](tk,xj))+Brσ)±hdσrp.\displaystyle\Phi^{\epsilon,\pm}_{j,k,p}:=x_{j}-h\,\big{(}H_{p}(x_{j},D\tilde{u}_{\rho,h}^{\epsilon}[\mu](t_{k},x_{j}))+B_{r}^{\sigma}\big{)}\pm\sqrt{hd}\sigma_{r}^{p}.

for jdj\in\mathbb{Z}^{d}, k=0,,Nk=0,\ldots,N, p=1,,dp=1,\ldots,d. Redefining Ei:=xi+ρ2(1,1)dE_{i}:=x_{i}+\frac{\rho}{2}(-1,1)^{d} and reasoning as for d=1d=1 above, we get the following discrete FPK equation

(24) {m~i,k+1[μ]:=jdm~j,k[μ]𝐁ρ,h,r[Hp(,Du~ρ,hϵ[μ])(i,j,k),m~i,0=Ei𝑑m0(x),\displaystyle\begin{cases}\tilde{m}_{i,k+1}[\mu]&:=\displaystyle\sum_{j\in\mathbb{Z}^{d}}\tilde{m}_{j,k}[\mu]\,\mathbf{B}_{\rho,h,r}[H_{p}(\cdot,D\tilde{u}_{\rho,h}^{\epsilon}[\mu])(i,j,k),\\ \tilde{m}_{i,0}&=\displaystyle\int_{E_{i}}dm_{0}(x),\end{cases}

where

(25) 𝐁ρ,h,r[Hp(,Du~ρ,hϵ[μ]](i,j,k):=[eλrh2dp=1d(βi(Φj,k,pϵ,+)+βi(Φj,k,pϵ,))+1eλrhλr|z|>rβi(xj+z)ν(dz)].\displaystyle\begin{split}\mathbf{B}_{\rho,h,r}[H_{p}(\cdot,D\tilde{u}_{\rho,h}^{\epsilon}[\mu]](i,j,k)&:=\bigg{[}\frac{e^{-\lambda_{r}h}}{2d}\sum_{p=1}^{d}\Big{(}\beta_{i}\big{(}\Phi^{\epsilon,+}_{j,k,p}\big{)}+\beta_{i}\big{(}\Phi^{\epsilon,-}_{j,k,p}\big{)}\Big{)}\\ &\hskip 56.9055pt+\frac{1-e^{-\lambda_{r}h}}{\lambda_{r}}\int_{|z|>r}\beta_{i}(x_{j}+z)\nu(dz)\bigg{]}.\end{split}

The solution is a probability distribution on 𝒢ρ×h𝒩h\mathcal{G}_{\rho}\times h\mathcal{N}_{h}, where 𝒩h:={0,,N}\mathcal{N}_{h}:=\{0,\dots,N\}:

Lemma 3.3.

Let (m~i,k)(\tilde{m}_{i,k}) be the solution of (24). If m0P(d)m_{0}\in P(\mathbb{R}^{d}), then (m~i,k)iP(d)(\tilde{m}_{i,k})_{i}\in P(\mathbb{Z}^{d}), i.e. m~i,k0\tilde{m}_{i,k}\geq 0, idi\in\mathbb{Z}^{d}, and jdm~j,k=1\sum_{j\in\mathbb{Z}^{d}}\tilde{m}_{j,k}=1 for all k𝒩hk\in\mathcal{N}_{h}.

Proof.

First note that m~i,k0\tilde{m}_{i,k}\geq 0 follows directly from the definition of the scheme and mi,00m_{i,0}\geq 0. Changing the order of summation and as i𝐁ρ,h,r[Hp(,Du~ρ,hϵ[μ]](i,j,k)=1\sum_{i}\mathbf{B}_{\rho,h,r}[H_{p}(\cdot,D\tilde{u}_{\rho,h}^{\epsilon}[\mu]](i,j,k)=1, we find that

im~i,k+1=ijm~j,k𝐁ρ,h,r[Hp(,Du~ρ,hϵ[μ]](i,j,k)=jm~j,k.\displaystyle\sum_{i}\tilde{m}_{i,k+1}=\sum_{i}\sum_{j}\tilde{m}_{j,k}\mathbf{B}_{\rho,h,r}[H_{p}(\cdot,D\tilde{u}_{\rho,h}^{\epsilon}[\mu]](i,j,k)=\sum_{j}\tilde{m}_{j,k}.

The result follows by iteration since jm~j,0=1\sum_{j}\tilde{m}_{j,0}=1. ∎

We extend (m~i,k[μ])(\tilde{m}_{i,k}[\mu]) to d\mathbb{R}^{d} by piecewise constant interpolation in xx and then to [0,T][0,T] by linear interpolation in tt: For t[tk,tk+1]t\in[t_{k},t_{k+1}] and k𝒩hk\in\mathcal{N}_{h},

(26) m~ρ,hϵ[μ](t,x)\displaystyle\tilde{m}_{\rho,h}^{\epsilon}[\mu](t,x) :=ttkhm~ρ,hϵ[μ](tk+1,x)+tk+1thm~ρ,hϵ[μ](tk,x),\displaystyle:=\frac{t-t_{k}}{h}\tilde{m}_{\rho,h}^{\epsilon}[\mu](t_{k+1},x)+\frac{t_{k+1}-t}{h}\tilde{m}_{\rho,h}^{\epsilon}[\mu](t_{k},x),

where, m~ρ,hϵ[μ](tk,x):=1ρdidm~i,k[μ] 1Ei(x)\tilde{m}_{\rho,h}^{\epsilon}[\mu](t_{k},x):=\frac{1}{\rho^{d}}\sum_{i\in\mathbb{Z}^{d}}\tilde{m}_{i,k}[\mu]\,\mathbbm{1}_{E_{i}}(x). Note that m~ρ,hϵ[μ]C([0,T],P(d))\tilde{m}_{\rho,h}^{\epsilon}[\mu]\in C([0,T],P(\mathbb{R}^{d})) and the duality with the linear in xx/constant in tt interpolation used for u~ρ,h\tilde{u}_{\rho,h} in (20).

3.5. Discretisation of the coupled MFG system

The discretisation of the MFG system is obtained by coupling the two discretisations above by setting μ=m~ρ,hϵ[μ]\mu=\tilde{m}^{\epsilon}_{\rho,h}[\mu]. With this choice and u=u~[μ]u=\tilde{u}[\mu] and m=m~[μ]m=\tilde{m}[\mu] we get the following discretisation of (1):

(27) {ui,k=Sρ,h,r[mρ,hϵ](u,k+1,i,k),ui,N=G(xi,mρ,hϵ(tN)),mi,k+1=jdmj,k𝐁ρ,h,r[Hp(,Duρ,hϵ)](i,j,k),mi,0=Ei𝑑m0(x),\displaystyle\begin{cases}u_{i,k}=S_{\rho,h,r}[m^{\epsilon}_{\rho,h}](u_{\cdot,k+1},i,k),\\[5.69046pt] u_{i,N}=G(x_{i},m^{\epsilon}_{\rho,h}(t_{N})),\\[5.69046pt] m_{i,k+1}=\sum_{j\in\mathbb{Z}^{d}}m_{j,k}\,\mathbf{B}_{\rho,h,r}[H_{p}(\cdot,Du_{\rho,h}^{\epsilon})](i,j,k),\\[5.69046pt] m_{i,0}=\int_{E_{i}}dm_{0}(x),\end{cases}

where Sρ,h,r,𝐁ρ,h,r,uρ,hϵ,mρ,hϵS_{\rho,h,r},\mathbf{B}_{\rho,h,r},u_{\rho,h}^{\epsilon},m^{\epsilon}_{\rho,h} are defined above.

The individual discretisations are explicit, but due to the forward-backward nature of the coupling, the total discretisation is not explicit. It yields a nonlinear system that must be solved by some method like e.g. a fixed point iteration or a Newton type method.

The approximation scheme (27) has a least one solution:

Proposition 3.4.

(Existence for the discrete MFG system) Assume (ν\nu0): , (ν\nu1): , (L1): (L2): , (F1): (F2): , (H1): , and (M): . Then there exist a pair (uρ,h,mρ,hϵ)(u_{\rho,h},\ m_{\rho,h}^{\epsilon}) solving (27).

The proof of this result is non-constructive and given in Appendix A.

4. Convergence to the MFG system

In this section we give the main theoretical results of this paper, various convergence results as h,ρ,ϵ,r0h,\rho,\epsilon,r\to 0 under CFL-conditions. The proofs will be given in Section 7 and require results for the individual schemes given in Sections 5 and 6.

4.1. Convergence to viscosity-very weak solutions

We consider degenerate and non-degenerate cases separately. For the degenerate case, the convergence holds only in dimension d=1d=1.

Theorem 4.1 (Degenerate case, d=1d=1).

Assume (ν\nu0): , (ν\nu1): , (L1): (L3): , (F1): (F3): , (H1): (H2): , (M’): , {(uρ,h,mρ,hϵ)}ρ,h,ϵ>0\{(u_{\rho,h},m^{\epsilon}_{\rho,h})\}_{\rho,h,\epsilon>0} are solutions of the discrete MFG system (27). If ρn,hn,ϵn,rn0\rho_{n},h_{n},\epsilon_{n},r_{n}\to 0 under the CFL conditions ρn2hn,hnrnσ,hnϵn=o(1)\frac{\rho_{n}^{2}}{h_{n}},\frac{h_{n}}{r_{n}^{\sigma}},\frac{\sqrt{h_{n}}}{\epsilon_{n}}=o(1), then:

  • (i)

    {uρn,hn}n\{u_{\rho_{n},h_{n}}\}_{n} is precompact in Cb([0,T]×K)C_{b}([0,T]\times K) for every compact set KK\subset\mathbb{R}.

  • (ii)

    {mρn,hnϵn}n\{m^{\epsilon_{n}}_{\rho_{n},h_{n}}\}_{n} is sequentially precompact in C([0,T],P())C([0,T],P(\mathbb{R})), and (a) in L1L^{1} weak if p(1,)p\in(1,\infty) in (M’): , or (b) in LL^{\infty} weak * if p=p=\infty in (M’): .

  • (iii)

    If (u,m)(u,m) is a limit point of {(uρn,hn,mρn,hnϵn)}n\{(u_{\rho_{n},h_{n}},m^{\epsilon_{n}}_{\rho_{n},h_{n}})\}_{n}, then (u,m)(u,m) is a viscosity-very weak solution of the MFG system (1).

Note that {mρ,hϵ}\{m^{\epsilon}_{\rho,h}\} is precompact in C([0,T],P(d))C([0,T],P(\mathbb{R}^{d})), just by assuming (M): for the initial distribution. But in the degenerate case this is not enough for convergence of the MFG system, due to lower regularity of the solutions of the HJB equation (no longer C1C^{1}). Therefore we need assumption (M’): and the stronger compactness given by Theorem 4.1(ii) part (a) or (b). This latter result we are only able to show in d=1d=1.

In arbitrary dimensions we assume more regularity on solutions of the HJB equation in (1):

(U):

Let u[m]u[m] be a viscosity solution of the HJB equation in (1). For any mC([0,T],P(d))m\in C([0,T],P(\mathbb{R}^{d})) and t(0,T)t\in(0,T), u[m](t)C1(d)u[m](t)\in C^{1}(\mathbb{R}^{d}).

Remark 4.2.

Assumption (U): holds in non-degenerate cases, e.g. under assumption (ν\nu2): , see Theorem 2.7 and the discussion below.

We have the following convergence result in arbitrary dimensions.

Theorem 4.3 (Non-degenerate case).

Assume (ν\nu0): , (ν\nu1): , (L1): (L3): , (F1): (F3): , (H1): (H2): , (U): , (M): , {(uρ,h,mρ,hϵ)}ρ,h,ϵ>0\{(u_{\rho,h},m^{\epsilon}_{\rho,h})\}_{\rho,h,\epsilon>0} are solutions of the discrete MFG system (27). If ρn,hn,ϵn,rn0\rho_{n},h_{n},\epsilon_{n},r_{n}\to 0 under the CFL conditions ρn2hn,hnrnσ,hnϵn=o(1)\frac{\rho_{n}^{2}}{h_{n}},\frac{h_{n}}{r_{n}^{\sigma}},\frac{\sqrt{h_{n}}}{\epsilon_{n}}=o(1), then:

  • (i)

    {uρn,hn}n\{u_{\rho_{n},h_{n}}\}_{n} is precompact in Cb([0,T]×K)C_{b}([0,T]\times K) for every compact set KdK\subset\mathbb{R}^{d}.

  • (ii)

    {mρn,hnϵn}n\{m^{\epsilon_{n}}_{\rho_{n},h_{n}}\}_{n} is precompact in C([0,T],P(d))C([0,T],P(\mathbb{R}^{d})).

  • (iii)

    If (u,m)(u,m) is a limit point of {(uρn,hn,mρn,hnϵn)}n\{(u_{\rho_{n},h_{n}},m^{\epsilon_{n}}_{\rho_{n},h_{n}})\}_{n}, then (u,m)(u,m) is a viscosity-very weak solution of the MFG system (1).

These results give compactness of the approximations and convergence along subsequences. To be precise, by part (i) and (ii) there are convergent subsequences, and by part (iii) the corresponding limits are solutions of the MFG system (1).

We immediately have existence for (1).

Corollary 4.4 (Existence of solutions of (1)).

Under the assumptions of either Theorem 4.1 or 4.3, there exists a viscosity-very weak solution (u,m)(u,m) of the MFG system (1).

If in addition we have uniqueness for the MFG system (1), then we have full convergence of the sequence of approximations.

Corollary 4.5.

Under the assumption of either Theorem 4.1 or Theorem 4.3, if the MFG system (1) has at most one viscosity-very weak solution, then the whole sequence {(uρn,hn,mρn,hnϵn)}n\{(u_{\rho_{n},h_{n}},m^{\epsilon_{n}}_{\rho_{n},h_{n}})\}_{n} converges to a limit (u,m)(u,m) which is the (unique) viscosity-very weak solution of the MFG system (1).

4.2. Convergence to classical solutions

In the case the individual equations are regularising, we can get convergence to classical solutions of the MFG system. To be precise we need:

  • 1.

    (“Weak” uniqueness of individual PDEs) The HJB equation have unique viscosity solutions, and the FPK equation have unique very weak solutions.

  • 2.

    (Smoothness of individual PDEs) Both equations have classical solutions.

This means that viscosity-very weak solutions of the MFG system automatically (by uniqueness for individual equations) are classical solutions. If in addition

  • 3.

    (Classical uniqueness for MFG) classical solutions of the MFG system are unique,

we get full convergence of the approximate solutions to the solution of the MFG system.

We now give a precise result in the setting of [37], see Theorem 2.7 in Section 2 for existence and uniqueness of classical solutions of (1).

Corollary 4.6.

Assume (ν\nu0): (ν\nu2): , (L1): (L3): , (F1): (F4): , (H3): (H4): , and (M”): . Let (uρ,h,mρ,hϵ)(u_{\rho,h},m^{\epsilon}_{\rho,h}) be solutions of the discrete MFG system (27). If ρn,hn,ϵn,rn0\rho_{n},h_{n},\epsilon_{n},r_{n}\to 0 under the CFL conditions ρn2hn,hnrnσ,hnϵn=o(1)\frac{\rho_{n}^{2}}{h_{n}},\frac{h_{n}}{r_{n}^{\sigma}},\frac{\sqrt{h_{n}}}{\epsilon_{n}}=o(1), then:

(a) {(uρn,hn,mρn,hnϵn)}n\{(u_{\rho_{n},h_{n}},m^{\epsilon_{n}}_{\rho_{n},h_{n}})\}_{n} has a convergent subsequence in the space Cb,loc([0,T]×d)×C([0,T],P(d))C_{b,\textup{loc}}([0,T]\times\mathbb{R}^{d})\times C([0,T],P(\mathbb{R}^{d})), and any limit point is a classical-classical solution of (1).

(b) If in addition (F5): and (H5): hold, then the whole sequence in (a) converges to the unique classical-classical solution (u,m)(u,m) of (1).

Proof.

1. Assumption (U): holds by Theorem 2.7, and then by Theorem 4.3, there is a convergent subsequence {(uρn,hn,mρn,hnϵn)}n\{(u_{\rho_{n},h_{n}},m_{\rho_{n},h_{n}}^{\epsilon_{n}})\}_{n} such that (uρn,hn,mρn,hnϵn)(u,m)(u_{\rho_{n},h_{n}},m_{\rho_{n},h_{n}}^{\epsilon_{n}})\to(u,m) and (u,m)(u,m) is a viscosity-very weak solution of (1).

2. Since mC([0,T],P(d))m\in C([0,T],P(\mathbb{R}^{d})), the viscosity solution uu is unique by Proposition 2.5 (b) (see also [37, Theorem 5.35.3]). Hence it coincides with the classical Cb1,3((0,T)×d)C_{b}^{1,3}((0,T)\times\mathbb{R}^{d}) solution given by [37, Theorem 5.55.5].

3. Now DpH(x,Du(t))Cb2(d)D_{p}H(x,Du(t))\in C_{b}^{2}(\mathbb{R}^{d}) by part 2 and (H3): , and then by Proposition 2.8 there is at most one very weak solution of the FPK equation. Hence it coincides with the classical Cb1,2((0,T)×d)C_{b}^{1,2}((0,T)\times\mathbb{R}^{d}) solution given by [37, Proposition 6.86.8].

4. In addition if (F5): and (H5): hold, there is a most one classical solution (u,m)(u,m) by Theorem 2.7 (b).

5. This shows (compactness, smoothness, and uniqueness) that all convergent subsequences of {(uρn,hn,mρn,hnϵn)}n\{(u_{\rho_{n},h_{n}},m_{\rho_{n},h_{n}}^{\epsilon_{n}})\}_{n} have the same limit, and thus the whole sequence converges to (u,m)(u,m), the unique classical solution of (1). ∎

4.3. Extension and discussion

Extension to more general Lévy operators

The results of Theorem 4.1 and 4.3 hold under much more general assumptions on the Lévy operator \mathcal{L}. In [37] they use (ν\nu0): together with the assumptions,

(ν\nu1):

r2+σ|z|<r|z|2𝑑ν+r1+σr<|z|<1|z|𝑑ν+rσr<|z|<1𝑑νc,r(0,1)\displaystyle r^{-2+\sigma}\int_{|z|<r}|z|^{2}d\nu+r^{-1+\sigma}\int_{r<|z|<1}|z|d\nu+r^{\sigma}\int_{r<|z|<1}d\nu\leq c,\ r\in(0,1).

(ν\nu2):

There are σ(1,2)\sigma\in(1,2) and 𝒦>0\mathcal{K}>0 such that the heat kernels KσK_{\sigma} and KσK_{\sigma}^{*} of \mathcal{L} and \mathcal{L}^{*} satisfy for K=Kσ,KσK=K_{\sigma},K_{\sigma}^{*} : K0K\geq 0, K(t,)L1(d)=1\|K(t,\cdot)\|_{L^{1}(\mathbb{R}^{d})}=1, and

DβK(t,)Lp(d)𝒦t1σ(|β|+(11p)d)for t(0,T)\displaystyle\|D^{\beta}K(t,\cdot)\|_{L^{p}(\mathbb{R}^{d})}\leq\mathcal{K}t^{-\frac{1}{\sigma}\big{(}|\beta|+(1-\frac{1}{p})d\big{)}}\quad\text{for $t\in(0,T)$}

and any p[1,)p\in[1,\infty) and multi-index βd{0}\beta\in\mathbb{N}^{d}\cup\{0\}.

where the heat kernel of the operator \mathcal{L} is defined as the fundamental solution of the heat equation tuu=0\partial_{t}u-\mathcal{L}u=0. These assumptions cover lots of new cases compared to (ν\nu0): , (ν\nu1): , and (ν\nu2): . New cases include (i) sums of operators satisfying (ν\nu1): on subspaces spanning d\mathbb{R}^{d}, having possibly different orders, (ii) more general non-absolutely continuous Lévy measures, and (iii) Lévy measures supported on positive cones. An example of (i) (cf. [37]) is

=(2x12)σ1/2(2xd2)σd/2,σ1,,σd(1,2),\mathcal{L}=-\Big{(}\!-\frac{\partial^{2}}{\partial x_{1}^{2}}\Big{)}^{\sigma_{1}/2}-\dots-\Big{(}\!-\frac{\partial^{2}}{\partial x_{d}^{2}}\Big{)}^{\sigma_{d}/2},\qquad\sigma_{1},\dots,\sigma_{d}\in(1,2),

which satisfies (ν\nu1): with σ=miniσi\sigma=\min_{i}\sigma_{i} and dν(z)=i=1ddzi|zi|1+σiΠjiδ0(dzj)d\nu(z)=\sum_{i=1}^{d}\frac{dz_{i}}{|z_{i}|^{1+\sigma_{i}}}\Pi_{j\neq i}\delta_{0}(dz_{j}). This is a sum of one-dimensional fractional Laplacians of different orders. An example of (iii) is given by the spectrally positive “fractional Laplacian” in one space dimension: u=cσ0(u(x+z)u(x)Du(x)z𝟙{z<1})z1σ𝑑z\mathcal{L}u=c_{\sigma}\int_{0}^{\infty}(u(x+z)-u(x)-Du(x)\cdot z\mathbbm{1}_{\{z<1\}})z^{-1-\sigma}dz.

We have the following generalization of the wellposedness result for classical solutions given in Theorem 2.7.

Theorem 4.7 ([37]).

Theorem 2.7 holds when you replace (ν\nu1): (ν\nu2): by (ν\nu1): (ν\nu2): .

It follows that (U): holds whenever Theorem 4.7 holds. Since (ν\nu1): implies (ν\nu1): and the integrals in (ν\nu1): are what appear in the different proofs, it is easy to check that all estimates in this paper are true for Lévy measures satisfying (ν\nu1): instead of (ν\nu1): . This means that under assumption (ν\nu1): and (ν\nu2): we have the following extensions of Theorems 4.1 and 4.3 and Corollary 4.6.

Theorem 4.8.

Theorem 4.1 holds when you replace (ν\nu1): with (ν\nu1): .

Theorem 4.9.

Theorem 4.3 holds when you replace (ν\nu1): (ν\nu2): by (ν\nu1): (ν\nu2): .

Corollary 4.10.

Corollary 4.6 holds when you replace (ν\nu1): (ν\nu2): by (ν\nu1): (ν\nu2): .

The Wasserstein metric d1d_{1} versus our metric d0d_{0}

The typical setting for the FPK equations in the MFG literature seems to be the metric space (P1(d),d1)(P_{1}(\mathbb{R}^{d}),d_{1}), that is the 11-Wasserstein space 𝒲1\mathcal{W}_{1} of probability measures with finite first moment. This is also the case in [25] where convergence results are given for SL schemes for local nondegenerate MFGs in d\mathbb{R}^{d}. In this paper we can not assume finite first moments if we want to cover general non-local operators. An example is the fractional Laplacian (Δ)σ2-(-\Delta)^{\frac{\sigma}{2}} for σ<1\sigma<1, where the underlying σ\sigma-stable process only has finite moments of order less than σ\sigma. Instead we consider the weaker metric space (P(d),d0)(P(\mathbb{R}^{d}),d_{0}), which is just a metrization of the weak (weak-* in CbC_{b}) convergence of probability measures. In this topology we can consider processes, probability measures and solutions of the FPK equations that do not have any finite moments or any restrictions on the tail behaviour of the corresponding Lévy measures.

Of course, under additional assumptions convergence in d0d_{0} implies convergence in d1d_{1}.

Lemma 4.11.

If mnm_{n} converges to mm in (P(d),d0)(P(\mathbb{R}^{d}),d_{0}) and mnm_{n} and mm has uniformly bounded (1+δ)(1+\delta)-moments for δ>0\delta>0, then mnmm_{n}\to m in (P1(d),d1)(P_{1}(\mathbb{R}^{d}),d_{1}).

Convergence in P1(d)P_{1}(\mathbb{R}^{d}) [53, Definition 6.8] is by definition equivalent to weak convergence plus convergence of first moments, and the result follows from e.g. Proposition 1.1 and Lemma 1.5 in [5].

We then have the following version of Theorem 4.1 and Theorem 4.3.

Corollary 4.12.

Assume m0P1+δ()m_{0}\in P_{1+\delta}(\mathbb{R}), dB1|z|1+δ𝑑ν(z)<\int_{\mathbb{R}^{d}\setminus B_{1}}|z|^{1+\delta}d\nu(z)<\infty for some δ>0\delta>0, and the assumptions of Theorem 4.1 and Theorem 4.3. Then the statements of Theorem 4.1 and Theorem 4.3 hold if we replace (P,d0)(P,d_{0}) by (P1,d1)(P_{1},d_{1}) in part (ii).

Note that the number of moments of mm is determined by the number of moments of 1|z|>1ν1_{|z|>1}\nu (and m0m_{0}), see e.g. the discussion in section 2.3 in [37]. Moreover, if 1|z|>1ν1_{|z|>1}\nu has at most α\alpha finite moments, then u\mathcal{L}u is well-defined only if uu has at most order α\alpha growth at infinity. Hence in the nonlocal case there is ”duality” between the moments of mm and the growth of uu. Note that umum will always be integrable which is natural since then e.g. Eu(Xt,t)=u(x,t)m(dx,t)Eu(X_{t},t)=\int u(x,t)m(dx,t) is finite.

In our case we assume no moments and have to work with bounded solutions uu.

On moments and weak compactness in LpL^{p} in the degenerate case

Previous results for Semi-Lagrangian schemes in the first order and the degenerate second order case [23, 24] cover the case m0P1()L()m_{0}\in P_{1}(\mathbb{R})\cap L^{\infty}(\mathbb{R}), which means that m0m_{0} has finite first-moments. Our results assume m0P()Lp()m_{0}\in P(\mathbb{R})\cap L^{p}(\mathbb{R}), for p(1,]p\in(1,\infty], and hence no moment bounds and possibly unbounded m0m_{0}. When p<p<\infty we have weak compactness in L1L^{1} instead of weak-* compactness in LL^{\infty}.

Since our results in the degenerate case allows for =0\mathcal{L}=0, they immediately give an extension to this PLpP\cap L^{p} setting for the convergence results for first order problems of [23]. Moreover, the same conditions, arguments, and results easily also holds in the local diffusive case considered in [24].

5. On the SL scheme for the HJB equation

We prove results for the numerical approximation of the HJB equation, including monotonicity, consistency, and different uniform a priori stability and regularity estimates. Using the “half-relaxed” limit method [12], we then show convergence in the form of vρn,hn[μn](tn,xn)v[μ](t,x)v_{\rho_{n},h_{n}}[\mu_{n}](t_{n},x_{n})\rightarrow v[\mu](t,x), where v[μ]v[\mu] is the (viscosity) solution of the continuous HJB equation. Let B(𝒢ρ)B(\mathcal{G}_{\rho}) be the set of all bounded functions defined on 𝒢ρ\mathcal{G}_{\rho}.

Theorem 5.1.

Assume (ν\nu0): , (L1): , ρ,h,r>0\rho,h,r>0, μC([0,T],P(d))\mu\in C([0,T],P(\mathbb{R}^{d})), and let Sρ,h,r[μ]S_{\rho,h,r}[\mu] denote the scheme defined in (18).

  • (i)

    (Bounded control) If ϕLip(d)\phi\in\text{Lip}(\mathbb{R}^{d}), Sρ,h,r[μ](ϕ,i,k)S_{\rho,h,r}[\mu](\phi,i,k) has a minimal control and |α|K|\alpha|\leq K where KK only depends on Dxϕ0\|D_{x}\phi\|_{0} and the growth of LL as |x||x|\rightarrow\infty.

  • (ii)

    (Monotonicity) For all v,wB(𝒢ρ)v,w\in B(\mathcal{G}_{\rho}) with vwv\leq w we have,

    Sρ,h,r[μ](v,i,k)Sρ,h,r[μ](w,i,k) for all i𝒢ρ,k=0,,N1.S_{\rho,h,r}[\mu](v,i,k)\leq S_{\rho,h,r}[\mu](w,i,k)\text{ for all }i\in\mathcal{G}_{\rho},\ k=0,\ldots,N-1.
  • (iii)

    (Commutation by constant) For every cc\in\mathbb{R} and wB(𝒢ρ)w\in B(\mathcal{G}_{\rho}),

    Sρ,h,r[μ](w+c,i,k)=Sρ,h,r[μ](w,i,k)+c for all i𝒢ρ,k=0,,N1.S_{\rho,h,r}[\mu](w+c,i,k)=S_{\rho,h,r}[\mu](w,i,k)+c\text{ for all }i\in\mathcal{G}_{\rho},\ k=0,\ldots,N-1.

Assume also (ν\nu1): and (F2): .

  • (iv)

    (Consistency) Let ρn,hn,rnn0\rho_{n},h_{n},r_{n}\xrightarrow{n\to\infty}0 under CFL conditions ρn2hn,hnrnσ=o(1)\frac{\rho_{n}^{2}}{h_{n}},\frac{h_{n}}{r_{n}^{\sigma}}=o(1), grid points (tkn,xin)(t,x)(t_{k_{n}},x_{i_{n}})\to(t,x), and μn,μC([0,T];P(d))\mu_{n},\mu\in C([0,T];P(\mathbb{R}^{d})) such that μnμ\mu_{n}\to\mu. Then, for every ϕCc(d×[0,T))\phi\in C_{c}^{\infty}(\mathbb{R}^{d}\times[0,T)),

    limn1hn[ϕ(tkn+1,xin)\displaystyle\lim_{n\to\infty}\frac{1}{h_{n}}\big{[}\phi(t_{k_{n}+1},x_{i_{n}})- Sρn,hn,rn[μn](ϕ,kn+1,in,kn)]\displaystyle S_{\rho_{n},h_{n},r_{n}}[\mu_{n}](\phi_{\cdot,k_{n}+1},i_{n},k_{n})\big{]}
    =\displaystyle= tϕ(t,x)infαd[L(x,α)Dϕα]ϕ(x)F(x,μ(t)).\displaystyle-\partial_{t}\phi(t,x)-\inf_{\alpha\in\mathbb{R}^{d}}\big{[}L(x,\alpha)-D\phi\cdot\alpha\big{]}-\mathcal{L}\phi(x)-F(x,\mu(t)).
Proof.

(i) Since

h(α):=ehλr2dm=1dI[ϕ](xi+hb¯(α)+hσrm)+I[ϕ](xi+hb¯(α)hσrm)\displaystyle h(\alpha):=\frac{e^{-h\lambda_{r}}}{2d}\sum_{m=1}^{d}I[\phi](x_{i}+h\bar{b}(\alpha)+\sqrt{h}\sigma_{r}^{m})+I[\phi](x_{i}+h\bar{b}(\alpha)-\sqrt{h}\sigma_{r}^{m})

is Lipschitz in α\alpha (maximum linear growth at infinity), while L(x,α)L(x,\alpha) is coercive (more than linear growth at infinity) by (L1): , there exists a ball BRB_{R}, where RR depends on the Lipschitz constant of I[ϕ]I[\phi] and the growth of LL, such that the minimizing control α¯\bar{\alpha} of Sρ,h,r[μ](ϕ,i,k)S_{\rho,h,r}[\mu](\phi,i,k) belongs to BRB_{R}.

(ii) and (iii) Follows directly from the definition of the scheme.

(iv) For ease of notation, we write ρ,h,r,μ\rho,h,r,\mu instead of ρn,hn,rn,μn\rho_{n},h_{n},r_{n},\mu_{n}. A 44th order Taylor expansion of ϕ\phi gives

ϕ(x+hb¯(α)±\displaystyle\phi(x+h\bar{b}(\alpha)\pm hdσrm)=ϕ(x)+Dϕ(x)(hb¯(α)±hdσrm)\displaystyle\sqrt{hd}\sigma_{r}^{m})=\phi(x)+D\phi(x)\cdot(h\bar{b}(\alpha)\pm\sqrt{hd}\sigma_{r}^{m})
+hd2(σrm)TD2ϕ(x)σrm±dh32b(α)TD2ϕ(x)σrm+h22b¯(α)TD2ϕ(x)b¯(α)\displaystyle+\frac{hd}{2}(\sigma_{r}^{m})^{T}D^{2}\phi(x)\sigma_{r}^{m}\pm\sqrt{d}\,h^{\frac{3}{2}}b(\alpha)^{T}D^{2}\phi(x)\sigma_{r}^{m}+\frac{h^{2}}{2}\bar{b}(\alpha)^{T}D^{2}\phi(x)\bar{b}(\alpha)
+|β|=3Dβϕ(x)β!(hb¯(α)±hdσrm)β+|β|=4Dβϕ(ξ±)β!(hb¯(α)±hdσrm)β,\displaystyle+\sum_{|\beta|=3}\frac{D^{\beta}\phi(x)}{\beta!}(h\bar{b}(\alpha)\pm\sqrt{hd}\,\sigma_{r}^{m})^{\beta}+\sum_{|\beta|=4}\frac{D^{\beta}\phi(\xi_{\pm})}{\beta!}(h\bar{b}(\alpha)\pm\sqrt{hd}\,\sigma_{r}^{m})^{\beta},

for some ξ±d\xi_{\pm}\in\mathbb{R}^{d}. Using that b¯(α)=αr|z|1zν(dz)\bar{b}(\alpha)=-\alpha-\int_{r\leq|z|\leq 1}z\nu(dz), and by (ν\nu1): r|z|1zν(dz)=O(r1σ)\int_{r\leq|z|\leq 1}z\nu(dz)=O(r^{1-\sigma}), we get that

(28) ϕ(x+hb¯(α)+hdσrm)+ϕ(x+hb¯(α)hdσrm)2ϕ(x)\displaystyle\phi(x+h\bar{b}(\alpha)+\sqrt{hd}\sigma_{r}^{m})+\phi(x+h\bar{b}(\alpha)-\sqrt{hd}\sigma_{r}^{m})-2\phi(x)
=2hD\displaystyle=-2hD ϕ(x)α2hr<|z|<1Dϕ(x)zν(dz)+hd(σrm)TD2ϕ(x)σrm+𝒪(h2r22σ).\displaystyle\phi(x)\cdot\alpha-2h\int_{r<|z|<1}D\phi(x)\cdot z\nu(dz)+hd(\sigma_{r}^{m})^{T}\cdot D^{2}\phi(x)\cdot\sigma_{r}^{m}+\mathcal{O}\big{(}h^{2}r^{2-2\sigma}\big{)}.

We used that h22b¯(α)TD2ϕ(x)b¯(α)\frac{h^{2}}{2}\bar{b}(\alpha)^{T}D^{2}\phi(x)\bar{b}(\alpha) is of order 𝒪(h2r22σ)\mathcal{O}(h^{2}r^{2-2\sigma}), the 33rd order terms are of order 𝒪(h3r33σ+h2r1σ)\mathcal{O}(h^{3}r^{3-3\sigma}+h^{2}r^{1-\sigma}), and the 44th order terms are of order (hdσr)4=𝒪(h2r42σ)(\sqrt{hd}\sigma_{r})^{4}=\mathcal{O}(h^{2}r^{4-2\sigma}). Then the error of the Taylor expansion is O(h2r22σ)O(h^{2}r^{2-2\sigma}). Using Lemma 3.1,

ϕ(xi)Sρ,h,r[μ](ϕ,i,k)\displaystyle\phi(x_{i})-S_{\rho,h,r}[\mu](\phi,i,k)
=ϕ(xi)infα[hF(xi,μ(tk+1))+hL(xi,α)+ehλr2dm=1d(2ϕ(xi)2hDϕ(xi)α\displaystyle=\ \phi(x_{i})-\inf_{\alpha}\bigg{[}hF(x_{i},\mu(t_{k+1}))+hL(x_{i},\alpha)+\frac{e^{-h\lambda_{r}}}{2d}\sum_{m=1}^{d}\Big{(}2\phi(x_{i})-2hD\phi(x_{i})\cdot\alpha
+hd(σrm)TD2ϕ(xi)σrm2hr<|z|<1Dϕ(xi)zν(dz))\displaystyle\hskip 71.13188pt+hd(\sigma_{r}^{m})^{T}D^{2}\phi(x_{i})\sigma_{r}^{m}-2h\int_{r<|z|<1}D\phi(x_{i})\cdot z\nu(dz)\Big{)}
+1ehλrλr|z|>rϕ(xi+z)ν(dz)+𝒪(ρ2)+𝒪(h2r22σ)]\displaystyle\hskip 71.13188pt+\frac{1-e^{-h\lambda_{r}}}{\lambda_{r}}\int_{|z|>r}\phi(x_{i}+z)\nu(dz)+\mathcal{O}(\rho^{2})+\mathcal{O}(h^{2}r^{2-2\sigma})\bigg{]}
(29) =hF(xi,μ(tk+1))infα[hL(xi,α)hehλrDϕ(xi)α]+(1ehλr)ϕ(xi)\displaystyle=\ hF(x_{i},\mu(t_{k+1}))-\inf_{\alpha}\bigg{[}hL(x_{i},\alpha)-he^{-h\lambda_{r}}D\phi(x_{i})\cdot\alpha\bigg{]}+(1-e^{-h\lambda_{r}})\phi(x_{i})
hehλr(rϕ(xi)+𝒪(r3σ))+hehλrr<|z|<1Dϕ(xi)zν(dz)\displaystyle\quad-he^{-h\lambda_{r}}\Big{(}\mathcal{L}_{r}\phi(x_{i})+\mathcal{O}(r^{3-\sigma})\Big{)}+he^{-h\lambda_{r}}\int_{r<|z|<1}D\phi(x_{i})\cdot z\nu(dz)
1ehλrλr|z|>rϕ(xi+z)ν(dz)+𝒪(ρ2+h2r22σ).\displaystyle\quad-\frac{1-e^{-h\lambda_{r}}}{\lambda_{r}}\int_{|z|>r}\phi(x_{i}+z)\nu(dz)+\mathcal{O}(\rho^{2}+h^{2}r^{2-2\sigma}).

Using that |z|<r|z|2ν(dz)Kr2σ\int_{|z|<r}|z|^{2}\nu(dz)\leq Kr^{2-\sigma} (by (ν\nu1): ), for the small jump operator r\mathcal{L}_{r} (defined in (12)) we have

(30) |rϕ(xi)ehλrrϕ(xi)|hλrr2σD2ϕ0.\displaystyle|\mathcal{L}_{r}\phi(x_{i})-e^{-h\lambda_{r}}\mathcal{L}_{r}\phi(x_{i})|\leq h\lambda_{r}\,r^{2-\sigma}\|D^{2}\phi\|_{0}.

Again, as r<|z|<1|z|ν(dz)Kr1σ\int_{r<|z|<1}|z|\nu(dz)\leq Kr^{1-\sigma} and |z|>1ν(dz)K\int_{|z|>1}\nu(dz)\leq K, for the long jump operator r\mathcal{L}^{r} (defined in (12)) we have that

|rϕ(xi)\displaystyle\Big{|}\mathcal{L}^{r}\phi(x_{i}) +ehλrr<|z|<1Dϕ(xi)zν(dz)1ehλrhλr|z|>r(ϕ(xi+z)ϕ(xi))ν(dz)|\displaystyle+e^{-h\lambda_{r}}\int_{r<|z|<1}D\phi(x_{i})\cdot z\nu(dz)-\frac{1-e^{-h\lambda_{r}}}{h\lambda_{r}}\int_{|z|>r}(\phi(x_{i}+z)-\phi(x_{i}))\nu(dz)\Big{|}
K(1ehλr)r1σDϕ0+K(11ehλrhλr)(r1σDϕ0+ϕ0)\displaystyle\leq K(1-e^{-h\lambda_{r}})r^{1-\sigma}\|D\phi\|_{0}+K\Big{(}1-\frac{1-e^{-h\lambda_{r}}}{h\lambda_{r}}\Big{)}\Big{(}r^{1-\sigma}\|D\phi\|_{0}+\|\phi\|_{0}\Big{)}
(31) K(hλrr1σDϕ0+hλrϕ0).\displaystyle\leq K\Big{(}h\lambda_{r}r^{1-\sigma}\|D\phi\|_{0}+h\lambda_{r}\|\phi\|_{0}\Big{)}.

Recalling that ϕ(xi)=rϕ(xi)+rϕ(xi)\mathcal{L}\phi(x_{i})=\mathcal{L}_{r}\phi(x_{i})+\mathcal{L}^{r}\phi(x_{i}), combining (5) with (30) and (5), we find

ϕ(xi)Sρ,h,r[μ](ϕ,i,k)\displaystyle\phi(x_{i})-S_{\rho,h,r}[\mu](\phi,i,k) =hF(xi,μ(tk+1))hinfα[L(xi,α)Dϕ(xi)α]hϕ(xi)\displaystyle=hF(x_{i},\mu(t_{k+1}))-h\inf_{\alpha}\bigg{[}L(x_{i},\alpha)-D\phi(x_{i})\cdot\alpha\bigg{]}-h\mathcal{L}\phi(x_{i})
+𝒪(h2λr+hr3σ+h2λrr1σ+ρ2+h2r22σ).\displaystyle\hskip 28.45274pt+\mathcal{O}\big{(}h^{2}\lambda_{r}+hr^{3-\sigma}+h^{2}\lambda_{r}r^{1-\sigma}+\rho^{2}+h^{2}r^{2-2\sigma}\big{)}.

As |λr|Crσ|\lambda_{r}|\leq Cr^{-\sigma}, we have

ϕ(tk,xi)ϕ(tk+1,xi)h+1h(ϕ(tk+1,xi)Sρ,h[μ](ϕ,k+1,i,k))\displaystyle\frac{\phi(t_{k},x_{i})-\phi(t_{k+1},x_{i})}{h}+\frac{1}{h}\Big{(}\phi(t_{k+1},x_{i})-S_{\rho,h}[\mu](\phi_{\cdot,k+1},i,k)\Big{)}
=tϕ(tk,xi)ϕ(tk+1,xi)+F(xi,μ(tk+1))infα[L(xi,α)Dϕ(tk+1,xi)α]\displaystyle=-\partial_{t}\phi(t_{k},x_{i})-\mathcal{L}\phi(t_{k+1},x_{i})+F(x_{i},\mu(t_{k+1}))-\inf_{\alpha}\bigg{[}L(x_{i},\alpha)-D\phi(t_{k+1},x_{i})\cdot\alpha\bigg{]}
+𝒪(h+hrσ+r3σ+hr12σ+ρ2h+hr22σ).\displaystyle\hskip 113.81102pt+\mathcal{O}\big{(}h+hr^{-\sigma}+r^{3-\sigma}+hr^{1-2\sigma}+\frac{\rho^{2}}{h}+hr^{2-2\sigma}\big{)}.

Hence the result follows by taking the limit nn\to\infty with ρn2hn,hnrnσ=o(1)\frac{\rho_{n}^{2}}{h_{n}},\frac{h_{n}}{r_{n}^{\sigma}}=o(1). ∎

Theorem 5.2.

(Comparison) Assume μ1,μ2C([0,T],P(d))\mu_{1},\mu_{2}\in C([0,T],P(\mathbb{R}^{d})), (ν\nu0): , and (L1): . Let uρ,h[μ1]u_{\rho,h}[\mu_{1}] and uρ,h[μ2]{u}_{\rho,h}[\mu_{2}] be defined by the scheme (20) for μ=μ1,μ2\mu=\mu_{1},\mu_{2}, respectively. Then,

uρ,h[μ1]uρ,h[μ2]0TF(,μ1)F(,μ2)0+G(,μ1(T))G(,μ2(T))0.\displaystyle\|u_{\rho,h}[\mu_{1}]-u_{\rho,h}[\mu_{2}]\|_{0}\leq T\|F(\cdot,\mu_{1})-F(\cdot,\mu_{2})\|_{0}+\|G(\cdot,\mu_{1}(T))-G(\cdot,\mu_{2}(T))\|_{0}.
Proof.

Let cm±(α):=hb¯(α)±hdσrmc^{\pm}_{m}(\alpha):=h\bar{b}(\alpha)\pm\sqrt{hd}\sigma_{r}^{m}, and note that

(32) I[u,k+1[μ1]](x)I[u,k+1[μ2]](x)=\displaystyle I[u_{\cdot,k+1}[\mu_{1}]](x)-I[u_{\cdot,k+1}[\mu_{2}]](x)= pdβp(x)(up,k+1[μ1]up,k+1[μ2]).\displaystyle\sum_{p\in\mathbb{Z}^{d}}\beta_{p}(x)(u_{p,k+1}[\mu_{1}]-u_{p,k+1}[\mu_{2}]).

By (18) and the definition of inf\inf, for any ϵ>0\epsilon>0, there is αϵd\alpha_{\epsilon}\in\mathbb{R}^{d} such that

ui,k\displaystyle u_{i,k} [μ2]hF(xi,μ2(tk))+hL(xi,αϵ)+ehλr2dm=1d[I[u,k+1[μ2]](xi+cm+(αϵ))\displaystyle[\mu_{2}]\geq\,hF(x_{i},\mu_{2}(t_{k}))+hL(x_{i},\alpha_{\epsilon})+\frac{e^{-h\lambda_{r}}}{2d}\sum_{m=1}^{d}\Big{[}I[u_{\cdot,k+1}[\mu_{2}]](x_{i}+c^{+}_{m}(\alpha_{\epsilon}))
(33) +I[u,k+1[μ2]](xi+cm(αϵ))]+1ehλrλr|z|rI[u,k+1[μ2]](xi+z)ν(dz)ϵ.\displaystyle+I[u_{\cdot,k+1}[\mu_{2}]](x_{i}+c^{-}_{m}(\alpha_{\epsilon}))\Big{]}+\frac{1-e^{-h\lambda_{r}}}{\lambda_{r}}\int_{|z|\geq r}I[u_{\cdot,k+1}[\mu_{2}]](x_{i}+z)\nu(dz)-\epsilon.

We then find, using (18), (32), (33),

ui,k[μ1]\displaystyle u_{i,k}[\mu_{1}] ui,k[μ2]h(F(xi,μ1(tk))F(xi,μ2(tk))+h(L(xi,αϵ)L(xi,αϵ))\displaystyle-u_{i,k}[\mu_{2}]\leq h\big{(}F(x_{i},\mu_{1}(t_{k}))-F(x_{i},\mu_{2}(t_{k})\big{)}+h(L(x_{i},\alpha_{\epsilon})-L(x_{i},\alpha_{\epsilon}))
+pd[ehλr2dm=1d(βp(cm+(αϵ))+βp(cm(αϵ)))(up+i,k+1[μ1]up+i,k+1[μ2])\displaystyle\quad+\sum_{p\in\mathbb{Z}^{d}}\bigg{[}\frac{e^{-h\lambda_{r}}}{2d}\sum_{m=1}^{d}\Big{(}\beta_{p}(c^{+}_{m}(\alpha_{\epsilon}))+\beta_{p}(c^{-}_{m}(\alpha_{\epsilon}))\Big{)}\big{(}u_{p+i,k+1}[\mu_{1}]-u_{p+i,k+1}[\mu_{2}]\big{)}
+1ehλrλr|z|rβp(z)(up+i,k+1[μ1]up+i,k+1[μ2])ν(dz)]+ϵ\displaystyle\quad+\frac{1-e^{-h\lambda_{r}}}{\lambda_{r}}\int_{|z|\geq r}\beta_{p}(z)\big{(}u_{p+i,k+1}[\mu_{1}]-u_{p+i,k+1}[\mu_{2}]\big{)}\nu(dz)\bigg{]}+\epsilon
hF(,μ1)F(,μ2)0+csupi|ui,k+1u~i,k+1|+ϵ,\displaystyle\leq h\|F(\cdot,\mu_{1})-F(\cdot,\mu_{2})\|_{0}+c\sup_{i}|u_{i,k+1}-\tilde{u}_{i,k+1}|+\epsilon,

where since pβp1\sum_{p}\beta_{p}\equiv 1,

c=ehλr2dm=1dpd(βp(cm+(αϵ))+βp(ci(αϵ)))+1ehλrλr|z|rpdβp(z)ν(dz)=1.\displaystyle c=\frac{e^{-h\lambda_{r}}}{2d}\sum_{m=1}^{d}\sum_{p\in\mathbb{Z}^{d}}\Big{(}\beta_{p}(c^{+}_{m}(\alpha_{\epsilon}))+\beta_{p}(c^{-}_{i}(\alpha_{\epsilon}))\Big{)}+\frac{1-e^{-h\lambda_{r}}}{\lambda_{r}}\int_{|z|\geq r}\sum_{p\in\mathbb{Z}^{d}}\beta_{p}(z)\nu(dz)=1.

Since |ui,N[μ1]ui,N[μ2]|G(,μ1(tN))G(,μ2(tN))0|u_{i,N}[\mu_{1}]-u_{i,N}[\mu_{2}]|\leq\|G(\cdot,\mu_{1}(t_{N}))-G(\cdot,\mu_{2}(t_{N}))\|_{0}, a symmetry and iteration argument shows that

|ui,k[μ1]ui,k[μ2]|(Nk)hF(,μ1)F(,μ2)0+G(,μ1(tN))G(,μ2(tN))0.\displaystyle\big{|}u_{i,k}[\mu_{1}]-u_{i,k}[\mu_{2}]\big{|}\leq(N-k)h\|F(\cdot,\mu_{1})-F(\cdot,\mu_{2})\|_{0}+\|G(\cdot,\mu_{1}(t_{N}))-G(\cdot,\mu_{2}(t_{N}))\|_{0}.

The result then follows from interpolation and T=NhT=Nh. ∎

The SL scheme is very stable in the sense that we have uniform in h,ρ,μh,\rho,\mu boundedness, Lipschitz continuity, and semi-concavity of the solutions ui,ku_{i,k}.

Lemma 5.3.

Let μC([0,T],P(d))\mu\in C([0,T],P(\mathbb{R}^{d})) and ui,k[μ]u_{i,k}[\mu] be defined by the scheme (18).

  1. (a)

    (Lipschitz continuity) Assume (ν\nu0): , (L2): and (F2): . Then

    |ui,kuj,k||xixj|(LF+LL)(Ttk)+LG,i,jd,k{0,1,N}.\displaystyle\frac{|u_{i,k}-u_{j,k}|}{|x_{i}-x_{j}|}\leq(L_{F}+L_{L})(T-t_{k})+L_{G},\quad i,j\in\mathbb{Z}^{d},\ k\in\{0,1,\ldots N\}.
  2. (b)

    (Semi-concavity) Assume (ν\nu0): , (L3): and (F3): . Then

    ui+j,k2ui,k+uij,k|xj|2(cF+cL)(Ttk)+cG,i,jd,k{0,1,N}.\displaystyle\frac{u_{i+j,k}-2u_{i,k}+u_{i-j,k}}{|x_{j}|^{2}}\leq(c_{F}+c_{L})(T-t_{k})+c_{G},\quad i,j\in\mathbb{Z}^{d},\ k\in\{0,1,\ldots N\}.
  3. (c)

    (Uniformly bounded) Assume (ν\nu0): , (L0): (L2): , (F1): , and (F2): . Then

    |ui,k|(CF+CL(K))(Ttk)+CG,i,jd,k{0,1,N},\displaystyle|u_{i,k}|\leq(C_{F}+C_{L}(K))(T-t_{k})+C_{G},\quad i,j\in\mathbb{Z}^{d},\ k\in\{0,1,\ldots N\},

    where KK is defined in Theorem 5.1 (i).

Proof.

(a) Note that since βm(xj+x)=βmj(x)\beta_{m}(x_{j}+x)=\beta_{m-j}(x),

(34) I[u,k+1](xj+x)I[u,k+1](xi+x)=\displaystyle I[u_{\cdot,k+1}](x_{j}+x)-I[u_{\cdot,k+1}](x_{i}+x)= pdβp(x)(up+j,k+1up+i,k+1).\displaystyle\sum_{p\in\mathbb{Z}^{d}}\beta_{p}(x)(u_{p+j,k+1}-u_{p+i,k+1}).

Then, by (L2): , (F2): , and similar computations as in Theorem 5.2, we find that

uj,kui,kh(Lf+LL)|xixj|+supj|ui,k+1uj,k+1|+ϵ,\displaystyle u_{j,k}-u_{i,k}\leq h(L_{f}+L_{L})|x_{i}-x_{j}|+\sup_{j}|u_{i,k+1}-u_{j,k+1}|+\epsilon,

Since |ui,N+1uj,N+1|=|G(xi,m(tN+1))G(xj,m(tN+1))|LG|xixj||u_{i,N+1}-u_{j,N+1}|=|G(x_{i},m(t_{N+1}))-G(x_{j},m(t_{N+1}))|\leq L_{G}|x_{i}-x_{j}| by (F2): , the result follows by iteration.
(b) Similar to (34) we see

I[u,k+1](xi+j+x)2I[u,k+1](xi+x)+I[u,k+1](xij+x)\displaystyle I[u_{\cdot,k+1}](x_{i+j}+x)-2I[u_{\cdot,k+1}](x_{i}+x)+I[u_{\cdot,k+1}](x_{i-j}+x)
=\displaystyle= pdβp(xi+x)(up+j,k+12up,k+1+upj,k+1).\displaystyle\sum_{p\in\mathbb{Z}^{d}}\beta_{p}(x_{i}+x)(u_{p+j,k+1}-2u_{p,k+1}+u_{p-j,k+1}).

Then, by (L3): , (F3): , and similar computations as in Theorem 5.2, we find that

ui+j,k2ui,k+uij,k(cL+cF)h|xj|2+supi(ui+j,k+12ui,k+1+uij,k+1).\displaystyle u_{i+j,k}-2u_{i,k}+u_{i-j,k}\leq(c_{L}+c_{F})h|x_{j}|^{2}+\sup_{i}(u_{i+j,k+1}-2u_{i,k+1}+u_{i-j,k+1}).

Since ui+j,N2ui,N+uij,NcG|xj|2u_{i+j,N}-2u_{i,N}+u_{i-j,N}\leq c_{G}|x_{j}|^{2} by (F3): , the result follows by iteration.

(c) By part (a) and Theorem 5.1 (i), |α|K|\alpha|\leq K, and then a direct calculation shows that

sup|α|K(h(|F|+|L|)+supj|uj,k+1|)ui,ksup|α|K(h(|F|+|L|)+supj|uj,k+1|).\displaystyle-\sup_{|\alpha|\leq K}\Big{(}h(|F|+|L|)+\sup_{j}|u_{j,k+1}|\Big{)}\leq u_{i,k}\leq\sup_{|\alpha|\leq K}\Big{(}h(|F|+|L|)+\sup_{j}|u_{j,k+1}|\Big{)}.

The result follows from (L1): and (F1): . ∎

Theorem 5.4.

(Convergence of the HJB scheme) Assume (ν\nu0): , (ν\nu1): , (F1): , (F2): , (L2): , ρn,hn,rnn0\rho_{n},h_{n},r_{n}\xrightarrow{n\to\infty}0 under CFL conditions ρn2hn,hnrnσ=o(1)\frac{\rho_{n}^{2}}{h_{n}},\frac{h_{n}}{r_{n}^{\sigma}}=o(1), μnμ\mu_{n}\rightarrow\mu in C([0,T],P(d))C([0,T],P(\mathbb{R}^{d})), and uρn,hn[μn]u_{\rho_{n},h_{n}}[\mu_{n}] is the solution of the scheme (18) defined by (20). Then there is a continuous bounded function u[μ]u[\mu] such that uρn,hn[μn]u[μ]u_{\rho_{n},h_{n}}[\mu_{n}]\rightarrow u[\mu] locally uniformly in d×[0,T]\mathbb{R}^{d}\times[0,T], and u[μ]u[\mu] is the viscosity solution of the HJB equation in (1) for m=μm=\mu.

Proof.

The result follows from the Barles-Perthame-Souganidis relaxed limit method [12], using the monotonicity, consistency, and LL^{\infty}-stability properties of the scheme (cf. Theorem 5.1 (ii), (iii), and Lemma 5.3 (c)), and the strong comparison principle for the HJB equation in Proposition 2.5 (a). We refer to the proof of [23, Theorem 3.3] for a standard but more detailed proof in a similar case. ∎

We recall that the continuous extensions uρ,h[μ](t,x)u_{\rho,h}[\mu](t,x) and uρ,hϵ[μ](t,x)u_{\rho,h}^{\epsilon}[\mu](t,x) are defined in (20) and (22), respectively. The results of Lemma 5.3 transfers to uρ,hϵ[μ](t,x)u_{\rho,h}^{\epsilon}[\mu](t,x).

Lemma 5.5.

Let μC([0,T],P(d))\mu\in C([0,T],P(\mathbb{R}^{d})) and uρ,hϵ[μ]u_{\rho,h}^{\epsilon}[\mu] be given by (22).

  1. (a)

    (Lipschitz continuity) Assume (ν\nu0): , (L2): and (F2): . Then

    |uρ,hϵ[μ](t,x)uρ,hϵ[μ](t,y)|((LL+LF)T+LG)|xy|.\displaystyle\big{|}u_{\rho,h}^{\epsilon}[\mu](t,x)-u_{\rho,h}^{\epsilon}[\mu](t,y)\big{|}\leq((L_{L}+L_{F})T+L_{G})|x-y|.
  2. (b)

    (Approximate semiconcavity) Assume (ν\nu0): , (L2): ,(L3): , (F2): , and (F3): . Then there exist a constant c1>0c_{1}>0, independent of ρ,h,ϵ\rho,h,\epsilon and μ\mu, such that

    uρ,hϵ[μ](t,x+y)2uρ,hϵ[μ](t,x)+uρ,hϵ[μ](t,xy)c1(|y|2+ρ2+ρ2ϵ),and\displaystyle u_{\rho,h}^{\epsilon}[\mu](t,x+y)-2u_{\rho,h}^{\epsilon}[\mu](t,x)+u_{\rho,h}^{\epsilon}[\mu](t,x-y)\leq c_{1}(|y|^{2}+\rho^{2}+\frac{\rho^{2}}{\epsilon}),\ \mbox{and}
    Duρ,hϵ[μ](t,y)Duρ,hϵ[μ](t,x),yxc1(|xy|2+ρ2ϵ2).\displaystyle\langle Du_{\rho,h}^{\epsilon}[\mu](t,y)-Du_{\rho,h}^{\epsilon}[\mu](t,x),y-x\rangle\leq c_{1}\Big{(}|x-y|^{2}+\frac{\rho^{2}}{\epsilon^{2}}\Big{)}.
  3. (c)

    Assume d=1d=1, (ν\nu0): , (L3): , and (F3): . Then there exists a constant c2>0c_{2}>0, independent of ρ,h,ϵ\rho,h,\epsilon and μ\mu, such that for each i,jdi,j\in\mathbb{Z}^{d} and k𝒩hk\in\mathcal{N}_{h}

    Duρ,hϵ[μ](tk,xj)Duρ,hϵ[μ](tk,xi),xjxic2|xjxi|2.\displaystyle\langle Du_{\rho,h}^{\epsilon}[\mu](t_{k},x_{j})-Du_{\rho,h}^{\epsilon}[\mu](t_{k},x_{i}),x_{j}-x_{i}\rangle\leq c_{2}|x_{j}-x_{i}|^{2}.
Proof.

(a) Since ui,ku_{i,k} satisfies the discrete Lipschitz bound of Lemma 5.3 (a), uρ,h[μ]u_{\rho,h}[\mu] is Lipschitz with same Lipschitz constant as ui,ku_{i,k} by properties of linear interpolation, and uρ,hϵ[μ]u_{\rho,h}^{\epsilon}[\mu] is Lipschitz with same constant as uρ,h[μ]u_{\rho,h}[\mu] by properties of mollifiers (Lemma 3.2).

(b) For i,jdi,j\in\mathbb{Z}^{d} we have by Lemma 5.3 (b), ui+j+uij2uic|xj|2.u_{i+j}+u_{i-j}-2u_{i}\leq c|x_{j}|^{2}. Multiplying both sides by βi(x)\beta_{i}(x), and summing over d\mathbb{Z}^{d}, we get

uρ,h(x+xj)+uρ,h(xxj)2uρ,h(x)c|xj|2.\displaystyle u_{\rho,h}(x+x_{j})+u_{\rho,h}(x-x_{j})-2u_{\rho,h}(x)\leq c|x_{j}|^{2}.

Letting xxzx\to x-z, multiplying by a positive mollifier ρϵ(z)\rho_{\epsilon}(z) and integrating, we get

uρ,hϵ(x+xj)+uρ,hϵ(xxj)2uρ,hϵ(x)c|xj|2.\displaystyle u_{\rho,h}^{\epsilon}(x+x_{j})+u_{\rho,h}^{\epsilon}(x-x_{j})-2u_{\rho,h}^{\epsilon}(x)\leq c|x_{j}|^{2}.

We multiply both sides with βj(y)\beta_{j}(y), and sum over d\mathbb{Z}^{d},

I[uρ,hϵ](x+y)+I[uρ,hϵ](xy)2I[uρ,hϵ](x)cI[||2](y)c(|y|2+ρ2).\displaystyle I[u_{\rho,h}^{\epsilon}](x+y)+I[u_{\rho,h}^{\epsilon}](x-y)-2I[u_{\rho,h}^{\epsilon}](x)\leq cI[|\cdot|^{2}](y)\leq c(|y|^{2}+\rho^{2}).

By Lemma 3.2 and part (a), we have that |I[uρ,hϵ](ξ)uρ,hϵ(ξ)|KD2uρ,hϵ0ρ2Kρ2ϵ|I[u_{\rho,h}^{\epsilon}](\xi)-u_{\rho,h}^{\epsilon}(\xi)|\leq K\|D^{2}u_{\rho,h}^{\epsilon}\|_{0}\rho^{2}\leq K\frac{\rho^{2}}{\epsilon}, where the Lipschitz bound KK depends on the constants in (L2): and (F2): . Thus,

uρ,hϵ(x+y)+uρ,hϵ(xy)2uρ,hϵ(x)c(|y|2+ρ2+ρ2ϵ).\displaystyle u_{\rho,h}^{\epsilon}(x+y)+u_{\rho,h}^{\epsilon}(x-y)-2u_{\rho,h}^{\epsilon}(x)\leq c(|y|^{2}+\rho^{2}+\frac{\rho^{2}}{\epsilon}).

The second part of (b) then follows as in [3, Remark 6].

(c) The proof is given in [23, Lemma 3.6]. ∎

Under our assumptions, the continuous HJB equation has a (viscosity) solution u(t)W1,(d)u(t)\in W^{1,\infty}(\mathbb{R}^{d}), that is, the derivative exists almost everywhere [37, Theorem 4.3]. We have the following result for Duρ,hϵ[μ]Du_{\rho,h}^{\epsilon}[\mu].

Theorem 5.6.

Assume (ν\nu0): , (ν\nu1): , (L1): (L2): , (F1): (F2): , ρn,hn,rn,ϵnn0\rho_{n},h_{n},r_{n},\epsilon_{n}\xrightarrow{n\to\infty}0 under CFL conditions ρn2hn,hnrnσ=o(1)\frac{\rho_{n}^{2}}{h_{n}},\frac{h_{n}}{r_{n}^{\sigma}}=o(1), and μnμ\mu_{n}\rightarrow\mu in C([0,T],P(d))C([0,T],P(\mathbb{R}^{d})). Let uρn,hnϵn[μn]u_{\rho_{n},h_{n}}^{\epsilon_{n}}[\mu_{n}] be defined by (22) and u[μ]u[\mu] the viscosity solution of the HJB equation in (1) for m=μm=\mu. Then

  • (i)

    uρn,hnϵn[μn]u[μ]u_{\rho_{n},h_{n}}^{\epsilon_{n}}[\mu_{n}]\rightarrow u[\mu] locally uniformly,

  • (ii)

    Assume also (L3): , (F3): and ρnϵn=o(1)\frac{\rho_{n}}{\epsilon_{n}}=o(1). Then Duρn,hnϵn[μn](t,x)Du[μ](t,x)Du_{\rho_{n},h_{n}}^{\epsilon_{n}}[\mu_{n}](t,x)\rightarrow Du[\mu](t,x) whenever Du(t,x)Du(t,x) exists, that is, the convergence is almost everywhere.

  • (iii)

    Assume also (L3): , (F3): , ρnϵn=o(1)\frac{\rho_{n}}{\epsilon_{n}}=o(1), and (U): . Then Duρn,hnϵn[μn]Du[μ]Du_{\rho_{n},h_{n}}^{\epsilon_{n}}[\mu_{n}]\rightarrow Du[\mu] locally uniformly.

Proof.

(i) This follows from the convergence result Theorem 5.4 and Lemma 3.2.

(ii) and (iii). We refer to [23, Theorem 3.53.5] and [25, Proposition 5.15.1]. Estimates from Lemma 5.5 are needed. For completeness we give the proof in Appendix B. ∎

6. On the dual SL scheme for the FPK equation

In this section we establish more properties of the discrete FPK equation (24), including tightness, equicontinuity in time, L1L^{1}-stability of solutions with respect to μ\mu, and LpL^{p}-bounds in dimension d=1d=1. To prove tightness we will use a result from [31].

Proposition 6.1.

Assume (ν\nu0): and (M): . Then there exists a function 0ΨC2(d)0\leq\Psi\in C^{2}(\mathbb{R}^{d}) with DΨ0,\|D\Psi\|_{0}, D2Ψ0<\|D^{2}\Psi\|_{0}<\infty, and lim|x|Ψ(x)=\displaystyle\lim_{|x|\rightarrow\infty}\Psi(x)=\infty, such that

(35) supxd||z|>1(Ψ(x+z)Ψ(z))ν(dz)|<anddΨ(x)m0(dx)<.\displaystyle\sup_{x\in\mathbb{R}^{d}}\Big{|}\int_{|z|>1}\big{(}\Psi(x+z)-\Psi(z)\big{)}\nu(dz)\Big{|}<\infty\quad\mbox{and}\quad\int_{\mathbb{R}^{d}}\Psi(x)\,m_{0}(dx)<\infty.
Proof.

We use [31, Lemma 4.9] on the family of measures {ν1,m0}\{\nu_{1},m_{0}\}, where ν1\nu_{1} is defined in (11), to get a function Ψ(x)=V0(1+|x|2)\Psi(x)=V_{0}(\sqrt{1+|x|^{2}}) such that V0:[0,)[0,)V_{0}:[0,\infty)\rightarrow[0,\infty) is a non-decreasing sub-additive function, V00,V0′′0<\|V_{0}^{\prime}\|_{0},\|V_{0}^{\prime\prime}\|_{0}<\infty, limxV0(x)=\displaystyle\lim_{x\rightarrow\infty}V_{0}(x)=\infty, and

dΨ(x)μ(dx)<forμ{ν1,m0}.\displaystyle\int_{\mathbb{R}^{d}}\Psi(x)\,\mu(dx)<\infty\qquad\mbox{for}\qquad\mu\in\{\nu_{1},m_{0}\}.

We immediately get the result except for the first part of (35). But this estimate follows from sub-additivity and ν1\nu_{1}-integrability of V0V_{0}, see [31, Lemma 4.13 (ii)]. ∎

Remark 6.2.

(a) If dνdzC|z|d+σ1\frac{d\nu}{dz}\leq\frac{C}{|z|^{d+\sigma_{1}}} for |z|>1|z|>1 and d|x|σ2m0(dx)<\int_{\mathbb{R}^{d}}|x|^{\sigma_{2}}\,m_{0}(dx)<\infty for σ1,σ2>0\sigma_{1},\sigma_{2}>0, then Ψ(z)=log(1+|z|2)\Psi(z)=\log(\sqrt{1+|z|^{2}}) is a possible explicit choice for the function in Proposition 6.1.

(b) Since ΨC2(d)\Psi\in C^{2}(\mathbb{R}^{d}), the first part of (35) is equivalent to Ψ0<\|\mathcal{L}\Psi\|_{0}<\infty (see [31, Lemma 4.13 (ii)]).

Lemma 6.3.

Assume {μα}αAP(d)\{\mu_{\alpha}\}_{\alpha\in A}\subset P(\mathbb{R}^{d}) and there exists a function 0ψC(d)0\leq\psi\in C(\mathbb{R}^{d}) such that lim|x|ψ(x)=\lim_{|x|\rightarrow\infty}\psi(x)=\infty and supαdψ(x)μα(dx)C\sup_{\alpha}\int_{\mathbb{R}^{d}}\psi(x)\mu_{\alpha}(dx)\leq C. Then {μα}α\{\mu_{\alpha}\}_{\alpha} is tight.

This result is classical and can be proved in a similar way as the Chebychev inequality.

Theorem 6.4 (Tightness).

Assume (ν\nu0): , (ν\nu1): , (L1): (L2): , (F2): , (H1): , (M): , the CFL condition ρ2h,hr12σ=𝒪(1)\frac{\rho^{2}}{h},hr^{1-2\sigma}=\mathcal{O}(1), μC([0,T],P(d))\mu\in C([0,T],P(\mathbb{R}^{d})), and mρ,hϵ[μ]m^{\epsilon}_{\rho,h}[\mu] is defined by (26). Take Ψ\Psi as in Proposition 6.1. Then there exists C>0C>0, independent of ρ,h,ϵ\rho,h,\epsilon and μ\mu, such that

dΨ(x)𝑑mρ,hϵ[μ](t)Cfor any t[0,T].\displaystyle\int_{\mathbb{R}^{d}}\Psi(x)\,dm^{\epsilon}_{\rho,h}[\mu](t)\leq C\qquad\mbox{for any }\qquad t\in[0,T].
Proof.

Essentially we start by multiplying the scheme (24) by Ψ\Psi and integrating in space. By the definition of mρ,hϵ=mρ,hϵ[μ]m^{\epsilon}_{\rho,h}=m^{\epsilon}_{\rho,h}[\mu] in (26) and (24), we find that

dΨ(x)𝑑mρ,hϵ(tk+1)\displaystyle\int_{\mathbb{R}^{d}}\Psi(x)dm^{\epsilon}_{\rho,h}(t_{k+1}) =1ρdidmi,k+1EiΨ(x)𝑑x\displaystyle=\frac{1}{\rho^{d}}\sum_{i\in\mathbb{Z}^{d}}m_{i,k+1}\int_{E_{i}}\Psi(x)dx
=id1ρdEiΨ(x)𝑑xjmj,k𝐁ρ,h,r[Hp(,Duρ,hϵ)](i,j,k).\displaystyle=\sum_{i\in\mathbb{Z}^{d}}\frac{1}{\rho^{d}}\int_{E_{i}}\Psi(x)dx\sum_{j}m_{j,k}\,\mathbf{B}_{\rho,h,r}[H_{p}(\cdot,Du_{\rho,h}^{\epsilon})](i,j,k).

By the definition of 𝐁ρ,h,r\mathbf{B}_{\rho,h,r} in (25) and interchanging the order of summation and integration, we have

dΨ(x)𝑑mρ,hϵ(tk+1)=\displaystyle\int_{\mathbb{R}^{d}}\Psi(x)d\,m^{\epsilon}_{\rho,h}(t_{k+1})= jdmj,kρd[eλrh2dp=1didEiΨ(x)(βi(Φj,k,pϵ,+)+βi(Φj,k,pϵ,))dx\displaystyle\sum_{j\in\mathbb{Z}^{d}}\frac{m_{j,k}}{\rho^{d}}\bigg{[}\frac{e^{-\lambda_{r}h}}{2d}\sum_{p=1}^{d}\sum_{i\in\mathbb{Z}^{d}}\int_{E_{i}}\Psi(x)\big{(}\beta_{i}(\Phi^{\epsilon,+}_{j,k,p})+\beta_{i}(\Phi^{\epsilon,-}_{j,k,p})\big{)}dx
+1eλrhλr|z|>ridEiΨ(x)βi(xj+z)dxν(dz)].\displaystyle\hskip 48.36958pt+\frac{1-e^{-\lambda_{r}h}}{\lambda_{r}}\int_{|z|>r}\sum_{i\in\mathbb{Z}^{d}}\int_{E_{i}}\Psi(x)\beta_{i}(x_{j}+z)dx\,\nu(dz)\bigg{]}.

Since ΨC2(d)\Psi\in C^{2}(\mathbb{R}^{d}), by properties of midpoint approximation and linear/multilinear interpolation we have |1ρdEiΨ(x)𝑑xΨ(xi)|+|Ψ(x)idβi(x)Ψ(xi)|𝒪(ρ2)\big{|}\frac{1}{\rho^{d}}\int_{E_{i}}\Psi(x)dx-\Psi(x_{i})\big{|}+\big{|}\Psi(x)-\sum_{i\in\mathbb{Z}^{d}}\beta_{i}(x)\Psi(x_{i})\big{|}\leq\mathcal{O}(\rho^{2}). Therefore

(36) dΨ(x)𝑑mρ,hϵ(tk+1)\displaystyle\int_{\mathbb{R}^{d}}\Psi(x)d\,m^{\epsilon}_{\rho,h}(t_{k+1})\leq jdmj,k[eλrh2dp=1d(Ψ(Φj,k,pϵ,+)+Ψ(Φj,k,pϵ,))\displaystyle\sum_{j\in\mathbb{Z}^{d}}m_{j,k}\bigg{[}\frac{e^{-\lambda_{r}h}}{2d}\sum_{p=1}^{d}\big{(}\Psi\big{(}\Phi^{\epsilon,+}_{j,k,p}\big{)}+\Psi\big{(}\Phi^{\epsilon,-}_{j,k,p}\big{)}\big{)}
+1eλrhλr|z|>rΨ(xj+z)ν(dz)]+𝒪(ρ2).\displaystyle\hskip 28.45274pt+\frac{1-e^{-\lambda_{r}h}}{\lambda_{r}}\int_{|z|>r}\Psi(x_{j}+z)\,\nu(dz)\bigg{]}+\mathcal{O}(\rho^{2}).

We estimate the terms on the right hand side. Let Φj,k,pϵ,±=xj±ah,j±\Phi^{\epsilon,\pm}_{j,k,p}=x_{j}\pm a^{\pm}_{h,j} where

(37) ah,j±=h(DpH(xj,Duρ,hϵ(tk,xj))+Brσ)±hσrp.\displaystyle a^{\pm}_{h,j}=h\,\Big{(}D_{p}H\big{(}x_{j},Du^{\epsilon}_{\rho,h}(t_{k},x_{j})\big{)}+B_{r}^{\sigma}\Big{)}\pm\sqrt{h}\sigma_{r}^{p}.

By the fundamental theorem of Calculus,

(38) Ψ(xjah,j+)+Ψ(xjah,j)=2Ψ(xj)(ah,j++ah,j)DΨ(xj)+E1\displaystyle\Psi(x_{j}-a^{+}_{h,j})+\Psi(x_{j}-a^{-}_{h,j})=2\Psi(x_{j})-(a^{+}_{h,j}+a^{-}_{h,j})\cdot D\Psi(x_{j})+E_{1}

where ah,j++ah,j=2h(DpH(xj,Duρ,hϵ(tk,xj))+Brσ)a^{+}_{h,j}+a^{-}_{h,j}=2h\,\big{(}D_{p}H\big{(}x_{j},Du^{\epsilon}_{\rho,h}(t_{k},x_{j})\big{)}+B_{r}^{\sigma}\big{)} and

E1=01[ah,j+(DΨ(xjtah,j+)DΨ(xj))+ah,j(DΨ(xjtah,j)DΨ(xj))]𝑑t.\displaystyle E_{1}=-\int_{0}^{1}\Big{[}a^{+}_{h,j}\cdot\big{(}D\Psi(x_{j}-ta^{+}_{h,j})-D\Psi(x_{j})\big{)}+a^{-}_{h,j}\cdot\big{(}D\Psi(x_{j}-ta^{-}_{h,j})-D\Psi(x_{j})\big{)}\Big{]}dt.

By Lemma 5.5 (a) and (H1): , we find that DpH(,Duρ,hϵ)0CR\|D_{p}H(\cdot,Du_{\rho,h}^{\epsilon})\|_{0}\leq C_{R} with R=(LL+LF)T+LG+1R=(L_{L}+L_{F})T+L_{G}+1, and then that

|E1|\displaystyle|E_{1}| D2Ψ0(|ah,j+|2+|ah,j|2)4D2Ψ0(h2(CR2+|Brσ|2)+h|σrp|2).\displaystyle\leq\|D^{2}\Psi\|_{0}(|a^{+}_{h,j}|^{2}+|a^{-}_{h,j}|^{2})\leq 4\|D^{2}\Psi\|_{0}\big{(}h^{2}(C_{R}^{2}+|B_{r}^{\sigma}|^{2})+h|\sigma_{r}^{p}|^{2}\big{)}.

To estimate the nonlocal term, we write

|z|>rΨ(xj+z)ν(dz)=|z|>1Ψ(xj+z)ν(dz)\displaystyle\int_{|z|>r}\Psi(x_{j}+z)\,\nu(dz)=\int_{|z|>1}\Psi(x_{j}+z)\nu(dz)
+r<|z|<1{Ψ(xj)+zDΨ(xj)+01z[DΨ(xj+tz)DΨ(xj)]𝑑t}ν(dz)\displaystyle\quad+\int_{r<|z|<1}\Big{\{}\Psi(x_{j})+z\cdot D\Psi(x_{j})+\int_{0}^{1}z\cdot\Big{[}D\Psi(x_{j}+tz)-D\Psi(x_{j})\Big{]}dt\Big{\}}\,\nu(dz)
||z|>1(Ψ(xj+z)Ψ(xj))ν(dz)|+λrΨ(xj)+BrσDΨ(xj)\displaystyle\leq\Big{|}\int_{|z|>1}\big{(}\Psi(x_{j}+z)-\Psi(x_{j})\big{)}\nu(dz)\Big{|}+\lambda_{r}\Psi(x_{j})+B_{r}^{\sigma}\cdot D\Psi(x_{j})
+D2Ψ0r<|z|<1|z|2ν(dz)\displaystyle\hskip 184.9429pt+\|D^{2}\Psi\|_{0}\int_{r<|z|<1}|z|^{2}\nu(dz)
λrΨ(xj)+BrσDΨ(xj)+E2,\displaystyle\leq\,\lambda_{r}\Psi(x_{j})+B_{r}^{\sigma}\cdot D\Psi(x_{j})+E_{2},

where E2E_{2} is finite and independent of ρ,h,ϵ\rho,h,\epsilon by Proposition 6.1 and |z|<1|z|2ν(dz)<\int_{|z|<1}|z|^{2}\nu(dz)<\infty. Going back to (36) and using the above estimates then leads to

dΨ(x)𝑑mρ,hϵ(tk+1)\displaystyle\int_{\mathbb{R}^{d}}\Psi(x)d\,m^{\epsilon}_{\rho,h}(t_{k+1})
jdmj,k[eλrh2dp=1d(2Ψ(xj)2h[DpH(xj,Duρ,hϵ(tk,xj))+Brσ]DΨ(xj)+|E1|)\displaystyle\leq\sum_{j\in\mathbb{Z}^{d}}m_{j,k}\bigg{[}\frac{e^{-\lambda_{r}h}}{2d}\sum_{p=1}^{d}\Big{(}2\Psi(x_{j})-2h\,\big{[}D_{p}H\big{(}x_{j},Du^{\epsilon}_{\rho,h}(t_{k},x_{j})\big{)}+B_{r}^{\sigma}\big{]}\cdot D\Psi(x_{j})+|E_{1}|\Big{)}
+1eλrhλr(λrΨ(xj)+BrσDΨ(xj)+E2)]+Cρ2\displaystyle\hskip 56.9055pt+\frac{1-e^{-\lambda_{r}h}}{\lambda_{r}}\Big{(}\lambda_{r}\Psi(x_{j})+B_{r}^{\sigma}\cdot D\Psi(x_{j})+E_{2}\Big{)}\bigg{]}+C\rho^{2}
jdmj,kΨ(xj)+C(h2λr|Brσ|+h2|Brσ|2+h+ρ2),\displaystyle\leq\sum_{j\in\mathbb{Z}^{d}}m_{j,k}\,\Psi(x_{j})+C\Big{(}h^{2}\lambda_{r}|B_{r}^{\sigma}|+h^{2}|B_{r}^{\sigma}|^{2}+h+\rho^{2}\Big{)},

where we used |heλrh+1eλrhλr|32λrh2|-he^{-\lambda_{r}h}+\frac{1-e^{-\lambda_{r}h}}{\lambda_{r}}|\leq\frac{3}{2}\lambda_{r}h^{2} and 1eλrhλrh\frac{1-e^{-\lambda_{r}h}}{\lambda_{r}}\leq h to get the last inequality.

With Ak+1=dΨ(x)𝑑mρ,hϵ(tk+1)A_{k+1}=\int_{\mathbb{R}^{d}}\Psi(x)d\,m^{\epsilon}_{\rho,h}(t_{k+1}), the above estimate becomes Ak+1Ak+EA_{k+1}\leq A_{k}+E where E=C(λrh2|Brσ|+h2|Brσ|2+h+ρ2)E=C(\lambda_{r}h^{2}|B_{r}^{\sigma}|+h^{2}|B_{r}^{\sigma}|^{2}+h+\rho^{2}). By iteration, |Brσ|2λr|Brσ|Cr12σ|B^{\sigma}_{r}|^{2}\leq\lambda_{r}|B_{r}^{\sigma}|\leq Cr^{1-2\sigma} (by (ν\nu0): , (ν\nu1): ), and kNChk\leq N\leq\frac{C}{h}, we find that

(39) Ak+1\displaystyle A_{k+1} A0+(k+1)EA0+C(hr12σ+1+ρ2h).\displaystyle\leq\,A_{0}+(k+1)E\leq A_{0}+C\Big{(}hr^{1-2\sigma}+1+\frac{\rho^{2}}{h}\Big{)}.

By assumption ρ2h,hr12σ=𝒪(1)\frac{\rho^{2}}{h},hr^{1-2\sigma}=\mathcal{O}(1), and by Proposition 6.1, A0=dΨ(x)𝑑m0<A_{0}=\int_{\mathbb{R}^{d}}\Psi(x)d\,m_{0}<\infty. Therefore

dΨ(x)𝑑mρ,hϵ(tk)Cfork=0,1,,N,\displaystyle\int_{\mathbb{R}^{d}}\Psi(x)d\,m^{\epsilon}_{\rho,h}(t_{k})\leq C\qquad\mbox{for}\qquad k=0,1,\dots,N,

for some constant C>0C>0 independent of ρ,h,ϵ,μ\rho,h,\epsilon,\mu, and hence by (26) the result follows for t[0,T]t\in[0,T]. ∎

Theorem 6.5 (Equicontinuity in time).

Assume (ν\nu0): , (ν\nu1): , (L1): (L2): , (F2): , (H1): , (M): , μC([0,T],P(d))\mu\in C([0,T],P(\mathbb{R}^{d})), and mρ,hϵ[μ]m^{\epsilon}_{\rho,h}[\mu] is defined by (26). Let ρ2h,hrσ=𝒪(1)\frac{\rho^{2}}{h},\frac{h}{r^{\sigma}}=\mathcal{O}(1) if σ(0,1)\sigma\in(0,1), or ρ2h,hr12σ=𝒪(1)\frac{\rho^{2}}{h},hr^{1-2\sigma}=\mathcal{O}(1) if σ(1,2)\sigma\in(1,2). Then there exists a constant C0>0C_{0}>0, independent of ρ,h,ϵ\rho,h,\epsilon and μ\mu, such that for any t1,t2[0,T]t_{1},t_{2}\in[0,T],

d0(mρ,hϵ[μ](t1),mρ,hϵ[μ](t2))C0|t1t2|.\displaystyle d_{0}(m^{\epsilon}_{\rho,h}[\mu](t_{1}),m^{\epsilon}_{\rho,h}[\mu](t_{2}))\leq C_{0}\sqrt{|t_{1}-t_{2}|}.
Proof.

We start by the case σ>1\sigma>1. For δ>0\delta>0, let ϕδ:=ϕρδ\phi_{\delta}:=\phi*\rho_{\delta} for ρδ\rho_{\delta} defined just before Lemma 3.2. With mρ,hϵ=mρ,hϵ[μ]m^{\epsilon}_{\rho,h}=m^{\epsilon}_{\rho,h}[\mu] we first note that

d0(mρ,hϵ(t1),mρ,hϵ(t2))=supϕLip1,1dϕ(x)(mρ,hϵ(t1)mρ,hϵ(t2))𝑑x\displaystyle d_{0}(m^{\epsilon}_{\rho,h}(t_{1}),m^{\epsilon}_{\rho,h}(t_{2}))=\sup_{\phi\in\mbox{Lip}_{1,1}}\int_{\mathbb{R}^{d}}\phi(x)(m^{\epsilon}_{\rho,h}(t_{1})-m^{\epsilon}_{\rho,h}(t_{2}))dx
=supϕLip1,1{d(ϕϕδ)(mρ,hϵ(t1)mρ,hϵ(t2))𝑑x+dϕδ(mρ,hϵ(t1)mρ,hϵ(t2))𝑑x}\displaystyle=\sup_{\phi\in\mbox{Lip}_{1,1}}\Big{\{}\int_{\mathbb{R}^{d}}(\phi-\phi_{\delta})(m^{\epsilon}_{\rho,h}(t_{1})-m^{\epsilon}_{\rho,h}(t_{2}))dx+\int_{\mathbb{R}^{d}}\phi_{\delta}\,(m^{\epsilon}_{\rho,h}(t_{1})-m^{\epsilon}_{\rho,h}(t_{2}))dx\Big{\}}
(40)  2δDϕ0+supϕLip1,1dϕδ(mρ,hϵ(t1)mρ,hϵ(t2))𝑑x,\displaystyle\leq\,2\delta\|D\phi\|_{0}+\sup_{\phi\in\mbox{Lip}_{1,1}}\int_{\mathbb{R}^{d}}\phi_{\delta}\,(m^{\epsilon}_{\rho,h}(t_{1})-m^{\epsilon}_{\rho,h}(t_{2}))dx,

where Lemma 3.2 was used to estimate the ϕϕδ\phi-\phi_{\delta} term and mρ,hϵ𝑑x=1\int m^{\epsilon}_{\rho,h}dx=1. Since mρ,hϵm^{\epsilon}_{\rho,h} and dϕδ(x)mρ,hϵ(t,x)𝑑x\int_{\mathbb{R}^{d}}\phi_{\delta}(x)m^{\epsilon}_{\rho,h}(t,x)dx are affine on each interval [tk,tk+1][t_{k},t_{k+1}], dϕδ(x)mρ,hϵ(,x)𝑑xW1,[0,T]\int_{\mathbb{R}^{d}}\phi_{\delta}(x)\,m^{\epsilon}_{\rho,h}(\cdot,x)dx\in W^{1,\infty}[0,T] and

ddtdϕδ(x)mρ,hϵ(,x)𝑑x0supk|Ik|.\displaystyle\Big{\|}\frac{d}{dt}\int_{\mathbb{R}^{d}}\phi_{\delta}(x)\,m^{\epsilon}_{\rho,h}(\cdot,x)dx\Big{\|}_{0}\leq\sup_{k}|I_{k}|.

where Ik=dϕδ(x)mρ,hϵ(tk+1,x)mρ,hϵ(tk,x)h𝑑xI_{k}=\int_{\mathbb{R}^{d}}\phi_{\delta}(x)\,\frac{m^{\epsilon}_{\rho,h}(t_{k+1},x)-m^{\epsilon}_{\rho,h}(t_{k},x)}{h}dx. It follows that

(41) dϕδ(mρ,hϵ(t1,x)mρ,hϵ(t2,x))𝑑x|t1t2|supk|Ik|.\displaystyle\int_{\mathbb{R}^{d}}\phi_{\delta}\,(m^{\epsilon}_{\rho,h}(t_{1},x)-m^{\epsilon}_{\rho,h}(t_{2},x))dx\leq|t_{1}-t_{2}|\sup_{k}|I_{k}|.

Let us estimate IkI_{k}. By (26), (24), (25), the midpoint quadrature approximation error bound, and the linear/multi-linear interpolation error bound, we have

Ik=\displaystyle I_{k}= 1hi1ρdEiϕδ(x)𝑑x[mi,k+1mi,k]\displaystyle\frac{1}{h}\sum_{i}\frac{1}{\rho^{d}}\int_{E_{i}}\phi_{\delta}(x)\,dx[m_{i,k+1}-m_{i,k}]
=\displaystyle= 1hρdj,i(Eiϕδ(x)𝑑x)[mj,k𝐁ρ,h,r[Hp(,Duρ,hϵ)](i,j,k)mi,kδi,j]\displaystyle\frac{1}{h\rho^{d}}\sum_{j,i}\Big{(}\int_{E_{i}}\phi_{\delta}(x)dx\Big{)}\Big{[}m_{j,k}\,\mathbf{B}_{\rho,h,r}[H_{p}(\cdot,Du_{\rho,h}^{\epsilon})](i,j,k)-m_{i,k}\,\delta_{i,j}\Big{]}
=\displaystyle= 1hjmj,k[iϕδ(xi)𝐁ρ,h,r[Hp(,Duρ,hϵ)](i,j,k)ϕδ(xj)+CD2ϕδ0ρ2]\displaystyle\frac{1}{h}\sum_{j}m_{j,k}\Big{[}\sum_{i}\phi_{\delta}(x_{i})\mathbf{B}_{\rho,h,r}[H_{p}(\cdot,Du_{\rho,h}^{\epsilon})](i,j,k)-\phi_{\delta}(x_{j})+C\|D^{2}\phi_{\delta}\|_{0}\rho^{2}\Big{]}
=\displaystyle= 1hjmj,k[eλrh2d(p=1dϕδ(Φj,k,pϵ,+)+ϕδ(Φj,k,pϵ,)2ϕδ(xj))\displaystyle\frac{1}{h}\sum_{j}m_{j,k}\Big{[}\frac{e^{-\lambda_{r}h}}{2d}\Big{(}\sum^{d}_{p=1}\phi_{\delta}(\Phi^{\epsilon,+}_{j,k,p})+\phi_{\delta}(\Phi^{\epsilon,-}_{j,k,p})-2\phi_{\delta}(x_{j})\Big{)}
+1eλrhλr|z|>r(ϕδ(xj+z)ϕδ(xj))ν(dz)+CD2ϕδ0ρ2].\displaystyle\hskip 56.9055pt+\frac{1-e^{-\lambda_{r}h}}{\lambda_{r}}\int_{|z|>r}\big{(}\phi_{\delta}(x_{j}+z)-\phi_{\delta}(x_{j})\big{)}\nu(dz)+C\|D^{2}\phi_{\delta}\|_{0}\rho^{2}\Big{]}.

Since Φj,k,pϵ,±=xj+ah,j±\Phi^{\epsilon,\pm}_{j,k,p}=x_{j}+a^{\pm}_{h,j} by (37), a 2nd order Taylor’s expansion gives us

|Ik|\displaystyle\big{|}I_{k}\big{|}\leq 1hjmj,k[eλrh((hDpH(xj,Duρ,hϵ[μ](tk,xj))hBrσ)Dϕδ(xj)\displaystyle\frac{1}{h}\sum_{j}m_{j,k}\bigg{[}e^{-\lambda_{r}h}\Big{(}(-hD_{p}H\big{(}x_{j},Du^{\epsilon}_{\rho,h}[\mu](t_{k},x_{j})\big{)}-hB_{r}^{\sigma})\cdot D\phi_{\delta}(x_{j})
+D2ϕδ02dp=1d(|ah,j+|2+|ah,j|2)+1eλrhλr(BrσDϕδ(xj)\displaystyle\hskip 28.45274pt+\frac{\|D^{2}\phi_{\delta}\|_{0}}{2d}\sum^{d}_{p=1}\big{(}|a^{+}_{h,j}|^{2}+|a^{-}_{h,j}|^{2}\big{)}+\frac{1-e^{-\lambda_{r}h}}{\lambda_{r}}\Big{(}B^{\sigma}_{r}\cdot D\phi_{\delta}(x_{j})
+D2ϕδ0r<|z|<1|z|2ν(dz)+2ϕδ0|z|>1ν(dz)+CD2ϕδ0ρ2)]\displaystyle\hskip 28.45274pt+\|D^{2}\phi_{\delta}\|_{0}\int_{r<|z|<1}|z|^{2}\nu(dz)+2\|\phi_{\delta}\|_{0}\int_{|z|>1}\nu(dz)+C\|D^{2}\phi_{\delta}\|_{0}\rho^{2}\Big{)}\bigg{]}
\displaystyle\leq 1h[(hDpH(,Duρ,hϵ)0+h2λr|Brσ|)Dϕδ0+c3hϕδ0\displaystyle\,\frac{1}{h}\bigg{[}\Big{(}h\|D_{p}H(\cdot,Du_{\rho,h}^{\epsilon})\|_{0}+h^{2}\lambda_{r}|B^{\sigma}_{r}|\Big{)}\|D\phi_{\delta}\|_{0}+c_{3}h{\|\phi_{\delta}\|}_{0}
+c1(h2DpH(,Duρ,hϵ)2+h2|Brσ|2+h|σr|2+h+ρ2)D2ϕδ0]jmj,k.\displaystyle\hskip 14.22636pt+c_{1}\Big{(}h^{2}{\|D_{p}H(\cdot,Du_{\rho,h}^{\epsilon})\|}^{2}+h^{2}{|B^{\sigma}_{r}|}^{2}+h{|\sigma_{r}|}^{2}+h+\rho^{2}\Big{)}{\|D^{2}\phi_{\delta}\|}_{0}\bigg{]}\sum_{j}m_{j,k}.

The above inequality follows since (1eλrhλrhehλr)h2λr(\frac{1-e^{-\lambda_{r}h}}{\lambda_{r}}-he^{-h\lambda_{r}})\leq h^{2}\lambda_{r} (used for the BrσDϕδB_{r}^{\sigma}\cdot D\phi_{\delta}-terms), and r<|z|<1|z|2ν(dz)+|z|>1ν(dz)C\int_{r<|z|<1}{|z|}^{2}\nu(dz)+\int_{|z|>1}\nu(dz)\leq C independently of rr by (ν\nu0): and (ν\nu1): . By Lemma 5.5 (a) and (H1): , DpH(,Duρ,hϵ)0CR\|D_{p}H(\cdot,Du_{\rho,h}^{\epsilon})\|_{0}\leq C_{R} with R=(LL+LF)T+LG+1R=(L_{L}+L_{F})T+L_{G}+1. Since mj,k=1\sum m_{j,k}=1, ϕLip1,1\phi\in\textup{Lip}_{1,1}, D2ϕδ0Dϕ0δ\|D^{2}\phi_{\delta}\|_{0}\leq\frac{\|D\phi\|_{0}}{\delta}, and |Brσ|2λr|Brσ|Kr12σ|B^{\sigma}_{r}|^{2}\leq\lambda_{r}|B_{r}^{\sigma}|\leq Kr^{1-2\sigma} (by (ν\nu0): , (ν\nu1): ), we get that

|Ik|C(1+hr12σ)+C(1+h+hr12σ+ρ2h)1δ.\displaystyle|I_{k}|\leq C(1+hr^{1-2\sigma})+C\big{(}1+h+hr^{1-2\sigma}+\frac{\rho^{2}}{h}\big{)}\frac{1}{\delta}.

To conclude the proof in the case σ>1\sigma>1, we go back to (40) and (41). In view of the above estimate on IkI_{k} and the assumption that ρ2h,hr12σ=𝒪(1)\frac{\rho^{2}}{h},hr^{1-2\sigma}=\mathcal{O}(1), we find that

d0(mρ,hϵ(t1),mρ,hϵ(t2))\displaystyle d_{0}(m^{\epsilon}_{\rho,h}(t_{1}),m^{\epsilon}_{\rho,h}(t_{2})) 2δ+C|t1t2|(1+1δ).\displaystyle\leq 2\delta+C|t_{1}-t_{2}|\Big{(}1+\frac{1}{\delta}\Big{)}.

Finally taking δ=|t1t2|\delta=\sqrt{|t_{1}-t_{2}|} we get d0(mρ,hϵ(t1),mρ,hϵ(t2))C|t1t2|.d_{0}(m^{\epsilon}_{\rho,h}(t_{1}),m^{\epsilon}_{\rho,h}(t_{2}))\leq C\sqrt{|t_{1}-t_{2}|}.

When σ<1\sigma<1, we find that |Brσ|C|B_{r}^{\sigma}|\leq C and hence that

|Ik|C(1+hrσ)+C(1+h+hrσ+ρ2h)1δ.\displaystyle|I_{k}|\leq C(1+hr^{-\sigma})+C\big{(}1+h+hr^{-\sigma}+\frac{\rho^{2}}{h}\big{)}\frac{1}{\delta}.

By assumption hrσ+ρ2h=𝒪(1)hr^{-\sigma}+\frac{\rho^{2}}{h}=\mathcal{O}(1), so again we find that

d0(mρ,hϵ(t1),mρ,hϵ(t2))2δ+C|t1t2|(1+1δ),\displaystyle d_{0}(m^{\epsilon}_{\rho,h}(t_{1}),m^{\epsilon}_{\rho,h}(t_{2}))\leq 2\delta+C|t_{1}-t_{2}|\Big{(}1+\frac{1}{\delta}\Big{)},

and can conclude as before. ∎

We also need a L1L^{1}-stability result for mρ,hϵ[μ]m^{\epsilon}_{\rho,h}[\mu] with respect to variations in μ\mu.

Lemma 6.6 (L1L^{1}-stability).

Assume (ν\nu0): , (H1): , and mρ,hϵ[μ]m^{\epsilon}_{\rho,h}[\mu] is defined by (26). Then for μ1,μ2C([0,T],P(d))\mu_{1},\mu_{2}\in C([0,T],P(\mathbb{R}^{d})),

supt[0,T]mρ,hϵ[μ1](t,)mρ,hϵ[μ2](t,)L1(d)\displaystyle\sup_{t\in[0,T]}\|m^{\epsilon}_{\rho,h}[\mu_{1}](t,\cdot)-m^{\epsilon}_{\rho,h}[\mu_{2}](t,\cdot)\|_{L^{1}(\mathbb{R}^{d})}
cKTρehλrDpH(,Duρ,hϵ[μ1])DpH(,Duρ,hϵ[μ2])0.\displaystyle\leq\frac{cKT}{\rho}e^{-h\lambda_{r}}\big{\|}D_{p}H(\cdot,Du^{\epsilon}_{\rho,h}[\mu_{1}])-D_{p}H(\cdot,Du^{\epsilon}_{\rho,h}[\mu_{2}])\big{\|}_{0}.
Proof.

Let α=DpH(,Duρ,hϵ[μ1])\alpha=D_{p}H(\cdot,Du^{\epsilon}_{\rho,h}[\mu_{1}]), α~=DpH(,Duρ,hϵ[μ2])\tilde{\alpha}=D_{p}H(\cdot,Du^{\epsilon}_{\rho,h}[\mu_{2}]), mj,k=mj,k[μ1]m_{j,k}=m_{j,k}[\mu_{1}], and m~j,k=mj,k[μ2]\tilde{m}_{j,k}=m_{j,k}[\mu_{2}]. By (25) and Lemma 3.3, 𝐁ρ,h,r[α](i,j,k)0\mathbf{B}_{\rho,h,r}[\alpha](i,j,k)\geq 0 and mj,k0m_{j,k}\geq 0, so that

i|mi,k+1m~i,k+1|=i|j(mj,k𝐁ρ,h,r[α](i,j,k)m~j,k𝐁ρ,h,r[α~](i,j,k))|\displaystyle\sum_{i}\big{|}m_{i,k+1}-\tilde{m}_{i,k+1}\big{|}=\sum_{i}\big{|}\sum_{j}(m_{j,k}\,\mathbf{B}_{\rho,h,r}[\alpha](i,j,k)-\tilde{m}_{j,k}\,\mathbf{B}_{\rho,h,r}[\tilde{\alpha}](i,j,k))\big{|}
ij(mj,k|𝐁ρ,h,r[α](i,j,k)𝐁ρ,h,r[α~](i,j,k)|+|mj,km~j,k|𝐁ρ,h,r[α~](i,j,k)).\displaystyle\leq\sum_{i}\sum_{j}\Big{(}m_{j,k}\big{|}\mathbf{B}_{\rho,h,r}[\alpha](i,j,k)-\mathbf{B}_{\rho,h,r}[\tilde{\alpha}](i,j,k)\big{|}+\big{|}m_{j,k}-\tilde{m}_{j,k}\big{|}\mathbf{B}_{\rho,h,r}[\tilde{\alpha}](i,j,k)\Big{)}.

Since i𝐁ρ,h,r[α~](i,j,k)=1\sum_{i}\mathbf{B}_{\rho,h,r}[\tilde{\alpha}](i,j,k)=1 (follows from iβi=1\sum_{i}\beta_{i}=1 and (25)),

ij|mj,km~j,k|𝐁ρ,h,r[α~](i,j,k)=j|mj,km~j,k|.\displaystyle\sum_{i}\sum_{j}\big{|}m_{j,k}-\tilde{m}_{j,k}\big{|}\mathbf{B}_{\rho,h,r}[\tilde{\alpha}](i,j,k)=\sum_{j}\big{|}m_{j,k}-\tilde{m}_{j,k}\big{|}.

Moreover, since only a finite number KdK_{d} of βi\beta_{i}’s are non-zero at any given point, βi\beta_{i} is Lipschitz with constant cρ\frac{c}{\rho}, and jmj,k=1\sum_{j}m_{j,k}=1 by Lemma 3.3, by the definitions of 𝐁ρ,h,r\mathbf{B}_{\rho,h,r} (25) and Φj,k,p±\Phi_{j,k,p}^{\pm} (23),

ijmj,k|𝐁ρ,h,r[α](i,j,k)𝐁ρ,h,r[α~](i,j,k)|\displaystyle\sum_{i}\sum_{j}m_{j,k}\big{|}\mathbf{B}_{\rho,h,r}[\alpha](i,j,k)-\mathbf{B}_{\rho,h,r}[\tilde{\alpha}](i,j,k)\big{|}
jmj,kehλr2dp=1di|βi(Φj,k,p+[μ1])βi(Φj,k,p+[μ2])\displaystyle\leq\sum_{j}m_{j,k}\frac{e^{-h\lambda_{r}}}{2d}\sum_{p=1}^{d}\sum_{i}\,\big{|}\beta_{i}(\Phi_{j,k,p}^{+}[\mu_{1}])-\beta_{i}(\Phi_{j,k,p}^{+}[\mu_{2}])
+βi(Φj,k,p[μ1])βi(Φj,k,p[μ2])|Kdchehλrραα~0.\displaystyle\hskip 56.9055pt+\beta_{i}(\Phi_{j,k,p}^{-}[\mu_{1}])-\beta_{i}(\Phi_{j,k,p}^{-}[\mu_{2}])\big{|}\leq K_{d}\frac{che^{-h\lambda_{r}}}{\rho}\|\alpha-\tilde{\alpha}\|_{0}.

An iteration then shows that

i|mi,k+1m~i,k+1|i|mi,0m~i,0|+cKdTρehλrαα~0.\displaystyle\sum_{i}\big{|}m_{i,k+1}-\tilde{m}_{i,k+1}\big{|}\leq\sum_{i}\big{|}m_{i,0}-\tilde{m}_{i,0}\big{|}+\frac{cK_{d}T}{\rho}e^{-h\lambda_{r}}\|\alpha-\tilde{\alpha}\|_{0}.

Since mi,0=m~i,0=Eim0𝑑xm_{i,0}=\tilde{m}_{i,0}=\int_{E_{i}}m_{0}\,dx, the result follows by interpolation. ∎

We end this section by a uniform LpL^{p}-bound on mρ,hϵm^{\epsilon}_{\rho,h} in dimension d=1d=1.

Theorem 6.7 (LpL^{p} bounds).

Assume d=1d=1, (ν\nu0): , (ν\nu1): , (L1): , (L3): , (F3): , (H2): , (M’): , μC([0,T],P(d))\mu\in C([0,T],P(\mathbb{R}^{d})), and mρ,hϵ[μ]m^{\epsilon}_{\rho,h}[\mu] be defined by (26). Then mρ,hϵ[μ]Lp()m_{\rho,h}^{\epsilon}[\mu]\in L^{p}(\mathbb{R}) and there exist a constant K>0K>0 independent of ϵ,h,ρ\epsilon,h,\rho and μ\mu such that

mρ,hϵ[μ](,t)Lp()eKTm0Lp().\displaystyle\|m^{\epsilon}_{\rho,h}[\mu](\cdot,t)\|_{L^{p}(\mathbb{R})}\leq e^{KT}\|m_{0}\|_{L^{p}(\mathbb{R})}.

To prove the theorem we need few technical lemmas.

Lemma 6.8.

Assume d=1d=1, (ν\nu0): , (ν\nu1): , (L1): , (L3): , (F3): , and (H2): . There exists a constant c0>0c_{0}>0 independent of ρ,h,ϵ,μ\rho,h,\epsilon,\mu such that

(DpH(xj,Duρ,hϵ(tk,xj))DpH(xi,Duρ,hϵ(tk,xi)))(xjxi)c0|xjxi|2.\Big{(}D_{p}H\big{(}x_{j},Du^{\epsilon}_{\rho,h}(t_{k},x_{j})\big{)}-D_{p}H\big{(}x_{i},Du^{\epsilon}_{\rho,h}(t_{k},x_{i})\big{)}\Big{)}(x_{j}-x_{i})\leq c_{0}|x_{j}-x_{i}|^{2}.
Proof.

By (L1): and (H2): for R=((LF+LL)T+LG)+1R=((L_{F}+L_{L})T+L_{G})+1 we have

(DpH(xj,Duρ,hϵ(tk,xj))DpH(xi,Duρ,hϵ(tk,xi)))(xjxi)\displaystyle\Big{(}D_{p}H\big{(}x_{j},Du^{\epsilon}_{\rho,h}(t_{k},x_{j})\big{)}-D_{p}H\big{(}x_{i},Du^{\epsilon}_{\rho,h}(t_{k},x_{i})\big{)}\Big{)}(x_{j}-x_{i})
=(xjxi)01ddt(DpH(xj,tDuρ,hϵ(tk,xj)+(1t)Duρ,hϵ(tk,xi)))𝑑t\displaystyle=(x_{j}-x_{i})\int_{0}^{1}\frac{d}{dt}\Big{(}D_{p}H\big{(}x_{j},t\,Du^{\epsilon}_{\rho,h}(t_{k},x_{j})+(1-t)Du^{\epsilon}_{\rho,h}(t_{k},x_{i})\big{)}\Big{)}dt
+(xjxi)(DpH(xj,Duρ,hϵ(tk,xi))DpH(xi,Duρ,hϵ(tk,xi)))\displaystyle\quad\ +(x_{j}-x_{i})\Big{(}D_{p}H\big{(}x_{j},Du^{\epsilon}_{\rho,h}(t_{k},x_{i})\big{)}-D_{p}H\big{(}x_{i},Du^{\epsilon}_{\rho,h}(t_{k},x_{i})\big{)}\Big{)}
=(xjxi)01DppH(xj,tDuρ,hϵ(tk,xj)\displaystyle=(x_{j}-x_{i})\int_{0}^{1}D_{pp}H\Big{(}x_{j},t\,Du^{\epsilon}_{\rho,h}(t_{k},x_{j})
+(1t)Duρ,hϵ(tk,xi))(Duρ,hϵ(tk,xj)Duρ,hϵ(tk,xi))dt\displaystyle\hskip 113.81102pt+(1-t)Du^{\epsilon}_{\rho,h}(t_{k},x_{i})\Big{)}\big{(}Du^{\epsilon}_{\rho,h}(t_{k},x_{j})-Du^{\epsilon}_{\rho,h}(t_{k},x_{i})\big{)}dt
+(xjxi)(DpH(xj,Duρ,hϵ(tk,xi))DpH(xi,Duρ,hϵ(tk,xi)))\displaystyle\quad\ +(x_{j}-x_{i})\Big{(}D_{p}H\big{(}x_{j},Du^{\epsilon}_{\rho,h}(t_{k},x_{i})\big{)}-D_{p}H\big{(}x_{i},Du^{\epsilon}_{\rho,h}(t_{k},x_{i})\big{)}\Big{)}
CRc2|xjxi|2+CR|xjxi|2,\displaystyle\leq C_{R}\,c_{2}|x_{j}-x_{i}|^{2}+C_{R}|x_{j}-x_{i}|^{2},

where the last inequality follows from convexity of HH (since LL is convex by (L1): ), semiconcavity of uρ,hϵu^{\epsilon}_{\rho,h} in Lemma 5.5 (c), and regularity of HH in (H2): . ∎

Lemma 6.9.

Assume d=1d=1, (ν\nu0): , (ν\nu1): , (L1): , (L3): , (F3): , (H2): , μC([0,T],P(d))\mu\in C([0,T],P(\mathbb{R}^{d})), and let Φj,kϵ,±[μ]\Phi^{\epsilon,\pm}_{j,k}[\mu] be defined in (23). There exist a constant K0>0K_{0}>0 independent of ϵ,ρ,h,μ\epsilon,\rho,h,\mu, such that for all ii\in\mathbb{Z} and k=𝒩hk=\mathcal{N}_{h},

max{jβi(Φj,kϵ,+)[μ],jβi(Φj,kϵ,)[μ]}1+K0h.\displaystyle\max\Big{\{}\sum_{j\in\mathbb{Z}}\beta_{i}(\Phi^{\epsilon,+}_{j,k})[\mu],\sum_{j\in\mathbb{Z}}\beta_{i}(\Phi^{\epsilon,-}_{j,k})[\mu]\Big{\}}\leq 1+K_{0}h.

The proof of this result is similar to the proof of [23, Lemma 3.8] – a slightly expanded proof is given in Appendix C. A similar result holds for the integral-term:

Lemma 6.10.

Assume d=1d=1. Then we have

1λrj|z|>rβi(xj+z)ν(dz)=1.\displaystyle\frac{1}{\lambda_{r}}\sum_{j\in\mathbb{Z}}\int_{|z|>r}\beta_{i}(x_{j}+z)\nu(dz)=1.
Proof.

By (11) and properties of the basis functions βj\beta_{j} we have

1λrj|z|>rβi(xj+z)ν(dz)=1λr|z|>rjβij(z)ν(dz)=1λr|z|>rν(dz)=1.\displaystyle\frac{1}{\lambda_{r}}\sum_{j\in\mathbb{Z}}\int_{|z|>r}\beta_{i}(x_{j}+z)\nu(dz)=\frac{1}{\lambda_{r}}\int_{|z|>r}\sum_{j\in\mathbb{Z}}\beta_{i-j}(z)\nu(dz)=\frac{1}{\lambda_{r}}\int_{|z|>r}\nu(dz)=1.\qquad\quad\qed
Proof of Theorem 6.7.

By definition of mρ,hϵm^{\epsilon}_{\rho,h} in (26) and the scheme (24),

(mρ,hϵ(x,tk+1))p𝑑x=(1ρimi,k+1𝟙Ei(x))p𝑑x\displaystyle\int_{\mathbb{R}}(m^{\epsilon}_{\rho,h}(x,t_{k+1}))^{p}dx=\int_{\mathbb{R}}\Big{(}\frac{1}{\rho}\sum_{i}m_{i,k+1}\mathbbm{1}_{E_{i}}(x)\Big{)}^{p}dx
=1ρp1i(mi,k+1)p=1ρp1i(jmj,k𝐁ρ,h,r(i,j,k))p,\displaystyle=\frac{1}{\rho^{p-1}}\sum_{i\in\mathbb{Z}}(m_{i,k+1})^{p}=\frac{1}{\rho^{p-1}}\sum_{i}\Big{(}\sum_{j}m_{j,k}\,\mathbf{B}_{\rho,h,r}(i,j,k)\Big{)}^{p},

where 𝐁ρ,h,r=𝐁ρ,h,r[Hp(,Duρ,hϵ[μ])]\mathbf{B}_{\rho,h,r}=\mathbf{B}_{\rho,h,r}[H_{p}(\cdot,Du_{\rho,h}^{\epsilon}[\mu])] is defined in (25). By Jensen’s inequality we have

i(jmj,k𝐁ρ,h,r(i,j,k))pi(p𝐁ρ,h,r(i,p,k))p1(j(mj,k)p𝐁ρ,h,r(i,j,k)),\displaystyle\sum_{i\in\mathbb{Z}}\Big{(}\sum_{j}m_{j,k}\,\mathbf{B}_{\rho,h,r}(i,j,k)\Big{)}^{p}\leq\sum_{i\in\mathbb{Z}}\Big{(}\sum_{p\in\mathbb{Z}}\mathbf{B}_{\rho,h,r}(i,p,k)\Big{)}^{p-1}\Big{(}\sum_{j}\big{(}m_{j,k}\big{)}^{p}\,\mathbf{B}_{\rho,h,r}(i,j,k)\Big{)},

and by Lemma 6.9 and 6.10,

p𝐁ρ,h,r(i,p,k)1+K0h,\sum_{p\in\mathbb{Z}}\mathbf{B}_{\rho,h,r}(i,p,k)\leq 1+K_{0}h,

where K0K_{0} is independent of i,ρ,h,ϵi,\rho,h,\epsilon and μ\mu. Since i𝐁ρ,h,r(i,p,k)=1\sum_{i}\mathbf{B}_{\rho,h,r}(i,p,k)=1 (follows from iβi=1\sum_{i}\beta_{i}=1), we find that

i(mi,k+1)p\displaystyle\sum_{i\in\mathbb{Z}}(m_{i,k+1})^{p} (1+K0h)p1j(mj,k)pi𝐁ρ,h,r(i,j,k)\displaystyle\leq(1+K_{0}h)^{p-1}\sum_{j}\big{(}m_{j,k}\big{)}^{p}\sum_{i}\mathbf{B}_{\rho,h,r}(i,j,k)
ρp1mρ,hϵ(tk,)Lp()p(1+K0h)p1.\displaystyle\leq\rho^{p-1}\|m^{\epsilon}_{\rho,h}(t_{k},\cdot)\|^{p}_{L^{p}(\mathbb{R})}(1+K_{0}h)^{p-1}.

By iteration and mρ,hϵ(,t0)Lp=m0Lp\|m^{\epsilon}_{\rho,h}(\cdot,t_{0})\|_{L^{p}}=\|m_{0}\|_{L^{p}}, mρ,hϵ(tk+1,)LpeK0T(11p)m0Lp\|m^{\epsilon}_{\rho,h}(t_{k+1},\cdot)\|_{L^{p}}\leq e^{K_{0}T(1-\frac{1}{p})}\|m_{0}\|_{L^{p}}, and the result follows for p[1,)p\in[1,\infty).

The proof of p=p=\infty is simpler, and in view of Lemma 6.9 and 6.10, the proof follows as in [24] for 2nd order case. ∎

7. Proof of convergence – Theorem 4.1 and 4.3

The main structure of the proofs are similar, so we present the proofs together. We proceed by several steps.

Step 1. (Compactness of mρn,hnϵnm^{\epsilon_{n}}_{\rho_{n},h_{n}})  In view of Theorem 6.4 and 6.5, mρ,hϵm^{\epsilon}_{\rho,h} is precompact in C([0,T],P(d))C([0,T],P(\mathbb{R}^{d})) by the Prokhorov and Arzelà-Ascoli Theorem. Hence there exist a subsequence {mρn,hnϵn}\{m^{\epsilon_{n}}_{\rho_{n},h_{n}}\} and mm in C([0,T],P(d))C([0,T],P(\mathbb{R}^{d})) such that

mρn,hnϵnminC([0,T],P(d)).\displaystyle m^{\epsilon_{n}}_{\rho_{n},h_{n}}\rightarrow m\quad\mbox{in}\quad C([0,T],P(\mathbb{R}^{d})).

This proves Theorem 4.3 (a) (ii) and the first part of Theorem 4.1 (a) (ii).

If (M’): holds with p=p=\infty, then Theorem 6.7 and Helly’s weak * compactness theorem imply that {mρ,hϵ}\{m^{\epsilon}_{\rho,h}\} is weak * precompact in L([0,T]×)L^{\infty}([0,T]\times\mathbb{R}) and there is a subsequence {mρn,hnϵn}\{m^{\epsilon_{n}}_{\rho_{n},h_{n}}\} and function mm such that mρn,hnϵnmm^{\epsilon_{n}}_{\rho_{n},h_{n}}\overset{\ast}{\rightharpoonup}m in L([0,T]×)L^{\infty}([0,T]\times\mathbb{R}). If (M’): holds with p(1,)p\in(1,\infty), then {mρ,hϵ}\{m^{\epsilon}_{\rho,h}\} is equiintegrable in [0,T]×[0,T]\times\mathbb{R} by Theorem 6.4 and 6.7 and de la Vallée Poussin’s theorem. By Dunford-Pettis’ theorem, it is then weakly precompact in L1([0,T]×)L^{1}([0,T]\times\mathbb{R}) and there exists a subsequence {mρn,hnϵn}\{m^{\epsilon_{n}}_{\rho_{n},h_{n}}\} and function mm such that mρn,hnϵnmm^{\epsilon_{n}}_{\rho_{n},h_{n}}\rightharpoonup m in L1([0,T]×)L^{1}([0,T]\times\mathbb{R}). The second part of Theorem 4.1 (a) (ii) follows.

Step 2. (Compactness and limit points for uρn,hnu_{\rho_{n},h_{n}})  Part (i) and limit points uu as viscosity solutions in part (iii) of both Theorem 4.1 and 4.3 follow from step 1 and Theorem 5.6 (i).

Step 3. (Consistency for mρn,hnϵnm^{\epsilon_{n}}_{\rho_{n},h_{n}})  Let (u,m)(u,m) be a limit point of {(uρn,hnϵn,mρn,hnϵn)}n\{(u^{\epsilon_{n}}_{\rho_{n},h_{n}},m^{\epsilon_{n}}_{\rho_{n},h_{n}})\}_{n}. Then by step 2, uu is a viscosity solution of the HJB equation in (1). We now show that mm is a very weak solution of the FPK equation in (1) with uu as the input data, i.e. mm satisfies (3) for t[0,T]t\in[0,T] and ϕCc(d)\phi\in C_{c}^{\infty}(\mathbb{R}^{d}). In the rest of the proof we use ρ,h,r,ϵ\rho,h,r,\epsilon instead of ρn,hn,rn,ϵn\rho_{n},h_{n},r_{n},\epsilon_{n} to simplify. We also let m^^=mρn,hnϵn\widehat{\widehat{m}}=m^{\epsilon_{n}}_{\rho_{n},h_{n}}, w=uρn,hnϵn[m^^]w=u_{\rho_{n},h_{n}}^{\epsilon_{n}}[\widehat{\widehat{m}}], and take tn=[thn]hnt_{n}=\big{[}\frac{t}{h_{n}}\big{]}h_{n}. Then we note that

dϕ(x)𝑑m^^(tn)(x)=dϕ(0)𝑑m0(x)+k=0n1dϕ(x)d[m^^(tk+1)m^^(tk)],\displaystyle\int_{\mathbb{R}^{d}}\phi(x)d\widehat{\widehat{m}}(t_{n})(x)=\int_{\mathbb{R}^{d}}\phi(0)dm_{0}(x)+\sum_{k=0}^{n-1}\int_{\mathbb{R}^{d}}\phi(x)d[\widehat{\widehat{m}}(t_{k+1})-\widehat{\widehat{m}}(t_{k})],

so to prove (3), we must estimate the sum on the right.

By the midpoint approximation and (26), the scheme (24), and (25) combined with linear/multilinear interpolation, and finally midpoint approximation again, we find that

dϕ(x)𝑑m^^(tk+1)=1ρdidmi,k+1Eiϕ(x)𝑑x=imi,k+1ϕ(xi)+𝒪(ρ2)\displaystyle\int_{\mathbb{R}^{d}}\phi(x)d\widehat{\widehat{m}}(t_{k+1})=\frac{1}{\rho^{d}}\sum_{i\in\mathbb{Z}^{d}}m_{i,k+1}\int_{E_{i}}\phi(x)dx=\sum_{i}m_{i,k+1}\phi(x_{i})+\mathcal{O}(\rho^{2})
=iϕ(xi)jmj,k𝐁ρ,h,r[Hp(,Dw)](i,j,k)+𝒪(ρ2)\displaystyle=\sum_{i}\phi(x_{i})\sum_{j}m_{j,k}\,\mathbf{B}_{\rho,h,r}[H_{p}(\cdot,Dw)](i,j,k)+\mathcal{O}(\rho^{2})
=jmj,k(eλrh2dp=1d[ϕ(Φj,k,pϵ,+)+ϕ(Φj,k,pϵ,)]+1eλrhλr|z|>rϕ(xj+z)ν(dz))+𝒪(ρ2)\displaystyle=\sum_{j}m_{j,k}\Big{(}\frac{e^{-\lambda_{r}h}}{2d}\sum_{p=1}^{d}[\phi(\Phi_{j,k,p}^{\epsilon,+})+\phi(\Phi_{j,k,p}^{\epsilon,-})]+\frac{1-e^{-\lambda_{r}h}}{\lambda_{r}}\int_{|z|>r}\phi(x_{j}+z)\nu(dz)\Big{)}+\mathcal{O}(\rho^{2})
=jmj,kρdEj(eλrh2dp=1d[ϕ(Φk,pϵ,+)(x)+ϕ(Φk,pϵ,)(x)]\displaystyle=\sum_{j}\frac{m_{j,k}}{\rho^{d}}\int_{E_{j}}\Big{(}\frac{e^{-\lambda_{r}h}}{2d}\sum_{p=1}^{d}[\phi(\Phi_{k,p}^{\epsilon,+})(x)+\phi(\Phi_{k,p}^{\epsilon,-})(x)]
+1eλrhλr|z|>rϕ(x+z)ν(dz))dx+𝒪(ρ2)+EΦ+Eν,\displaystyle\hskip 76.82234pt+\frac{1-e^{-\lambda_{r}h}}{\lambda_{r}}\int_{|z|>r}\phi(x+z)\nu(dz)\Big{)}dx+\mathcal{O}(\rho^{2})+E_{\Phi}+E_{\nu},

where Φj,k,pϵ,±\Phi_{j,k,p}^{\epsilon,\pm} is defined in (23), Φk,pϵ,±(x)=xh(Hp(x,Dw(tk,x))+Brσ)±hdσrp\Phi^{\epsilon,\pm}_{k,p}(x)=x-h\,\big{(}H_{p}(x,Dw(t_{k},x))+B_{r}^{\sigma}\big{)}\pm\sqrt{hd}\sigma_{r}^{p}, and EΦ+EνE_{\Phi}+E_{\nu} is the error of the last midpoint approximation. Since ϕ\phi is smooth, uρ,hu_{\rho,h} uniformly Lipschitz (Lemma 5.5 (a)), D2w0CDuρ,h0ϵ\|D^{2}w\|_{0}\leq\frac{C\|Du_{\rho,h}\|_{0}}{\epsilon}, and by assumption (H2): ,

|ϕ(Φj,k,pϵ,±)1ρdEjϕ(Φk,pϵ,±)(x)𝑑x|\displaystyle\Big{|}\phi(\Phi^{\epsilon,\pm}_{j,k,p})-\frac{1}{\rho^{d}}\int_{E_{j}}\phi(\Phi^{\epsilon,\pm}_{k,p})(x)dx\Big{|}
Dϕ0ρdEj|xxj|𝑑x+hDϕ0ρdEj|DpH(xj,Dw(tk,xj))DpH(x,Dw(tk,x))|𝑑x\displaystyle\leq\frac{\|D\phi\|_{0}}{\rho^{d}}\int_{E_{j}}|x-x_{j}|dx+\frac{h\|D\phi\|_{0}}{\rho^{d}}\int_{E_{j}}\big{|}D_{p}H(x_{j},Dw(t_{k},x_{j}))-D_{p}H(x,Dw(t_{k},x))\big{|}dx
Kρ(1+h(Hpp0D2w0+Hpx0))Kρ(1+hϵDuρ,h0),\displaystyle\leq K\rho\big{(}1+h(\|H_{pp}\|_{0}\|D^{2}w\|_{0}+\|H_{px}\|_{0})\big{)}\leq K\rho\big{(}1+\frac{h}{\epsilon}\|Du_{\rho,h}\|_{0}\big{)},

and hence EΦ=𝒪(hρϵ)E_{\Phi}=\mathcal{O}(\frac{h\rho}{\epsilon}). Similarly, Eν=𝒪(hρ2λr)=𝒪(hρ2rσ)E_{\nu}=\mathcal{O}(h\rho^{2}\lambda_{r})=\mathcal{O}(\frac{h\rho^{2}}{r^{\sigma}}).

From the above estimates, we find that

dϕ(x)d(m^^(tk+1)m^^(tk))(x)=d(eλrh2dp=1d[ϕ(Φk,pϵ,+)(x)+ϕ(Φk,pϵ,)(x)2ϕ(x)]\displaystyle\int_{\mathbb{R}^{d}}\phi(x)d\big{(}\widehat{\widehat{m}}(t_{k+1})-\widehat{\widehat{m}}(t_{k})\big{)}(x)=\int_{\mathbb{R}^{d}}\Big{(}\frac{e^{-\lambda_{r}h}}{2d}\sum_{p=1}^{d}[\phi(\Phi_{k,p}^{\epsilon,+})(x)+\phi(\Phi_{k,p}^{\epsilon,-})(x)-2\phi(x)]
+1eλrhλr|z|>r(ϕ(x+z)ϕ(x))ν(dz))dm^^(tk)(x)+𝒪(ρ2+hρϵ+hρ2rσ).\displaystyle\qquad\quad+\frac{1-e^{-\lambda_{r}h}}{\lambda_{r}}\int_{|z|>r}\big{(}\phi(x+z)-\phi(x)\big{)}\nu(dz)\Big{)}d\widehat{\widehat{m}}(t_{k})(x)+\mathcal{O}\big{(}\rho^{2}+\frac{h\rho}{\epsilon}+\frac{h\rho^{2}}{r^{\sigma}}\big{)}.

By a similar argument as in (28) and using Lemma 3.1,

ϕ(Φk,pϵ,+)(x)+ϕ(Φk,pϵ,)(x)2ϕ(x)=\displaystyle\phi(\Phi_{k,p}^{\epsilon,+})(x)+\phi(\Phi_{k,p}^{\epsilon,-})(x)-2\phi(x)= 2h(Dϕ(x)DpH(x,Dw(tk,x))+BrσDϕ(x))\displaystyle-2h\Big{(}D\phi(x)\cdot D_{p}H(x,Dw(t_{k},x))+B_{r}^{\sigma}\cdot D\phi(x)\Big{)}
+2hr[ϕ](x)+𝒪(h2r22σ+hr3σ).\displaystyle\,+2h\mathcal{L}_{r}[\phi](x)+\mathcal{O}(h^{2}r^{2-2\sigma}+hr^{3-\sigma}).

Hence using (30) and (5) we have

dϕ(x)d(m^^(tk+1)m^^(tk))(x)\displaystyle\int_{\mathbb{R}^{d}}\phi(x)d(\widehat{\widehat{m}}(t_{k+1})-\widehat{\widehat{m}}(t_{k}))(x)
=hd[Dϕ(x)DpH(x,Dw(tk,x))+r[ϕ](x)+r[ϕ](x)]𝑑m^^(tk)(x)\displaystyle=h\int_{\mathbb{R}^{d}}\big{[}-D\phi(x)\cdot D_{p}H(x,Dw(t_{k},x))+\mathcal{L}_{r}[\phi](x)+\mathcal{L}^{r}[\phi](x)\big{]}d\widehat{\widehat{m}}(t_{k})(x)
+𝒪(h2rσ+h2r12σ+h2r22σ)+𝒪(ρ2+hρϵ+hρ2rσ+h2r22σ+hr3σ).\displaystyle\quad+\mathcal{O}(h^{2}r^{-\sigma}+h^{2}r^{1-2\sigma}+h^{2}r^{2-2\sigma})+\mathcal{O}(\rho^{2}+\frac{h\rho}{\epsilon}+\frac{h\rho^{2}}{r^{\sigma}}+h^{2}r^{2-2\sigma}+hr^{3-\sigma}).

Summing from k=0k=0 to k=n1k=n-1 and approximating sums by integrals, we obtain

(42) dϕ(x)𝑑m^^(tn)(x)dϕ(x)𝑑m^^(t0)=hk=0n1d[Dϕ(x)DpH(x,Dw(tk,x))+[ϕ](x)]𝑑m^^(tk)(x)+n𝒪(ρ2+hρϵ+hρ2rσ+h2rσ+hr3σ)=d0tn[Dϕ(x)DpH(x,Dw(s,x))+[ϕ](x)]𝑑m^^(s)(x)𝑑s+𝒪(ρ2h+ρϵ+ρ2+hrσ+r3σ)+E,\displaystyle\begin{split}&\int_{\mathbb{R}^{d}}\phi(x)d\widehat{\widehat{m}}(t_{n})(x)-\int_{\mathbb{R}^{d}}\phi(x)d\widehat{\widehat{m}}(t_{0})\\ &=h\sum_{k=0}^{n-1}\int_{\mathbb{R}^{d}}\big{[}-D\phi(x)\cdot D_{p}H(x,Dw(t_{k},x))+\mathcal{L}[\phi](x)\big{]}d\widehat{\widehat{m}}(t_{k})(x)\\ &\qquad\qquad+n\,\mathcal{O}(\rho^{2}+\frac{h\rho}{\epsilon}+\frac{h\rho^{2}}{r^{\sigma}}+h^{2}r^{-\sigma}+hr^{3-\sigma})\\ &=\int_{\mathbb{R}^{d}}\int_{0}^{t_{n}}\big{[}-D\phi(x)\cdot D_{p}H(x,Dw(s,x))+\mathcal{L}[\phi](x)\big{]}d\widehat{\widehat{m}}(s)(x)\,ds\\ &\qquad\qquad+\mathcal{O}\Big{(}\frac{\rho^{2}}{h}+\frac{\rho}{\epsilon}+\frac{\rho^{2}+h}{r^{\sigma}}+r^{3-\sigma}\Big{)}+E,\end{split}

where EE is Riemann sum approximation error. Let Ik(x):=Dϕ(x)DpH(x,Dw(tk,x))I_{k}(x):=-D\phi(x)\cdot D_{p}H(x,Dw(t_{k},x)) +[ϕ](x)+\mathcal{L}[\phi](x) and use time-continuity m^^\widehat{\widehat{m}} in the d0d_{0}-metric (Theorem 6.5), that w(,x)w(\cdot,x) is constant on [tk,tk+1)[t_{k},t_{k+1}), (H1): , (H2): and D2w0CDuρ,h0ϵ\|D^{2}w\|_{0}\leq\frac{C\|Du_{\rho,h}\|_{0}}{\epsilon}, to conclude that for s[tk,tk+1)s\in[t_{k},t_{k+1})

tktk+1dIk(x)d(m^^(tk)m^^(s))(x)𝑑sh(Ik0+DIk0)C0sups[tk,tk+1)stk\displaystyle\int_{t_{k}}^{t_{k+1}}\int_{\mathbb{R}^{d}}I_{k}(x)d\big{(}\widehat{\widehat{m}}(t_{k})-\widehat{\widehat{m}}(s)\big{)}(x)ds\leq h\big{(}\|I_{k}\|_{0}+\|DI_{k}\|_{0}\big{)}C_{0}\sup_{s\in[t_{k},t_{k+1})}\sqrt{s-t_{k}}
Kh(1+Dw0+D2w0)hKh(1+1ϵ)h.\displaystyle\leq Kh\Big{(}1+\|Dw\|_{0}+\|D^{2}w\|_{0}\Big{)}\sqrt{h}\leq Kh\Big{(}1+\frac{1}{\epsilon}\Big{)}\sqrt{h}.

Summing over kk, we have E=|k=0n1tktk+1dIk(x)d(m^^(tk)m^^(s))(x)𝑑s|=𝒪(hϵ)E=\big{|}\sum_{k=0}^{n-1}\int_{t_{k}}^{t_{k+1}}\int_{\mathbb{R}^{d}}I_{k}(x)d\big{(}\widehat{\widehat{m}}(t_{k})-\widehat{\widehat{m}}(s)\big{)}(x)ds\big{|}=\mathcal{O}(\frac{\sqrt{h}}{\epsilon}).

Since m^^\widehat{\widehat{m}} converges to mm in C([0,T],P(d))C([0,T],P(\mathbb{R}^{d})) and ϕCc(d)\phi\in C_{c}^{\infty}(\mathbb{R}^{d}) implies [ϕ]Cb(d)\mathcal{L}[\phi]\in C_{b}(\mathbb{R}^{d}), we have

(43) d0tn[ϕ](x)𝑑m^^(s)(x)nd0t[ϕ](x)𝑑m(s)(x).\displaystyle\int_{\mathbb{R}^{d}}\int_{0}^{t_{n}}\mathcal{L}[\phi](x)d\widehat{\widehat{m}}(s)(x)\xrightarrow{n\rightarrow\infty}\int_{\mathbb{R}^{d}}\int_{0}^{t}\mathcal{L}[\phi](x)d{m}(s)(x).

It now remains to show convergence of the DpHD_{p}H-term and pass to the limit in (42) to get that mm is a very weak solution satisfying (3).

Step 4 (Proof of Theorem 4.1 (a) (iii)). Now d=1d=1 and part (ii) of Theorem 4.1 (a) implies that m^^m\widehat{\widehat{m}}\overset{\ast}{\rightharpoonup}m in L([0,t]×)L^{\infty}([0,t]\times\mathbb{R}) if m0L()m_{0}\in L^{\infty}(\mathbb{R}), or m^^m\widehat{\widehat{m}}\rightharpoonup m in L1([0,t]×)L^{1}([0,t]\times\mathbb{R}) if m0Lp()m_{0}\in L^{p}(\mathbb{R}) for p(1,)p\in(1,\infty). We also have Dw(t,x)=Duρ,hϵ(t,x)Du(t,x)Dw(t,x)=Du_{\rho,h}^{\epsilon}(t,x)\rightarrow Du(t,x) almost everywhere in [0,T]×[0,T]\times\mathbb{R} by Theorem 5.6 (ii). Since DϕCc()D\phi\in C_{c}^{\infty}(\mathbb{R}) and DpH(,Dw)D_{p}H(\cdot,Dw) uniformly bounded, by the triangle inequality and the dominated convergence Theorem we find that

0tnDϕ(x)\displaystyle\int_{\mathbb{R}}\int_{0}^{t_{n}}D\phi(x)\cdot DpH(x,Dw(s,x))dm^^(s)(x)\displaystyle D_{p}H(x,Dw(s,x))\,d\widehat{\widehat{m}}(s)(x)
0tDϕ(x)DpH(x,Du(s,x))𝑑m(s)(x).\displaystyle\longrightarrow\int_{\mathbb{R}}\int_{0}^{t}D\phi(x)\cdot D_{p}H(x,Du(s,x))\,d{m}(s)(x).

Then by passing to the limit in (42) using the above limit, (43), and the CFL conditions ρ2h,hrσ,hϵ=o(1)\frac{\rho^{2}}{h},\frac{h}{r^{\sigma}},\frac{\sqrt{h}}{\epsilon}=\mathit{o}(1) (note that ρ2h\rho^{2}\leq h for large nn), we see that (3) holds and m{m} is a very weak solution of the FPK equation. This completes the proof of Theorem 4.1 (a) (iii).

Step 5(Proof of Theorem 4.3(iii)).  Now (U): holds and Dw=Duρ,hϵDuDw=Du_{\rho,h}^{\epsilon}\rightarrow Du locally uniformly by Theorem 5.6 (iii). Since DϕCc(d)D\phi\in C^{\infty}_{c}(\mathbb{R}^{d}) and d𝑑m^^(s)(x)=1\int_{\mathbb{R}^{d}}d\widehat{\widehat{m}}(s)(x)=1, by continuity and uniform boundedness of DpH(,Dw)D_{p}H(\cdot,Dw), it follows that

(44) |d0tnDϕ(x)DpH(x,Dw(s,x))dm^^(s)(x)d0tnDϕ(x)DpH(x,Du(s,x))dm^^(s)(x)|TDϕ0DppH0DwDuL(supp(ϕ))d𝑑m^^(s)(x)0.\displaystyle\begin{split}&\Big{|}\int_{\mathbb{R}^{d}}\int_{0}^{t_{n}}D\phi(x)\cdot D_{p}H(x,Dw(s,x))\,d\widehat{\widehat{m}}(s)(x)\\ &\hskip 85.35826pt-\int_{\mathbb{R}^{d}}\int_{0}^{t_{n}}D\phi(x)\cdot D_{p}H(x,Du(s,x))\,d{\widehat{\widehat{m}}}(s)(x)\Big{|}\\ &\leq T\|D\phi\|_{0}\|D_{pp}H\|_{0}\|Dw-Du\|_{L^{\infty}(\textup{supp}(\phi))}\int_{\mathbb{R}^{d}}d\widehat{\widehat{m}}(s)(x)\longrightarrow 0.\end{split}

Since m^^m\widehat{\widehat{m}}\rightarrow m in C([0,T],P(d))C([0,T],P(\mathbb{R}^{d})) and DϕDpH(,Du)(t)Cb(d)D\phi\cdot D_{p}H(\cdot,Du)(t)\in C_{b}(\mathbb{R}^{d}) by (U): , we get

d0tnDϕ(x)\displaystyle\int_{\mathbb{R}^{d}}\int_{0}^{t_{n}}D\phi(x)\cdot DpH(x,Du(s,x))dm^^(s)(x)\displaystyle D_{p}H(x,Du(s,x))\,d\widehat{\widehat{m}}(s)(x)
d0tDϕ(x)DpH(x,Du(s,x))𝑑m(s)(x).\displaystyle\longrightarrow\int_{\mathbb{R}^{d}}\int_{0}^{t}D\phi(x)\cdot D_{p}H(x,Du(s,x))\,d{m}(s)(x).

Then by passing to the limit in (42) using the above limit, (44), (43), and the CFL conditions ρ2h,hrσ,hϵ=o(1)\frac{\rho^{2}}{h},\frac{h}{r^{\sigma}},\frac{\sqrt{h}}{\epsilon}=\mathit{o}(1), we see that (3) holds and m{m} is a very weak solution of the FPK equation. This completes the proof of Theorem 4.3(iii).

8. Numerical examples

For numerical experiments we look at

(45) {utσ2u+12|ux|2=f(t,x)+Kϕδm(t,x) in (0,T)×[a,b],mtσ2mdiv(mux)=0 in (0,T)×[a,b],u(T,x)=G(x,m(T)),m(x,0)=m0(x) in [a,b],\displaystyle\begin{cases}-u_{t}-\sigma^{2}\mathcal{L}u+\frac{1}{2}|u_{x}|^{2}=f(t,x)+K\ \phi_{\delta}\ast m(t,x)\quad&\text{ in }(0,T)\times[a,b],\\ m_{t}-\sigma^{2}\mathcal{L}^{*}m-\text{div}(mu_{x})=0\quad&\text{ in }(0,T)\times[a,b],\\ u(T,x)=G(x,m(T)),\qquad m(x,0)=m_{0}(x)\quad&\text{ in }[a,b],\end{cases}

where a<ba<b are real numbers, \mathcal{L} is a diffusion operator, ϕδ=1δ2πex22δ2\phi_{\delta}=\frac{1}{\delta\sqrt{2\pi}}e^{-\frac{x^{2}}{2\delta^{2}}}, KK some real number, and ff is some bounded smooth function. We will specify these quantities in the examples below.

Artificial boundary conditions

Our schemes (18) and (24) for approximating (45) are posed in all of \mathbb{R}. To work in a bounded domain we impose (artificial) exterior conditions:

  1. (U1)

    uu00+TfL((0,T)×(a,b))u\equiv\|u_{0}\|_{0}+T\cdot\|f\|_{L^{\infty}((0,T)\times(a,b))} in ([a,b])×[0,T](\mathbb{R}\setminus[a,b])\times[0,T],

  2. (M1)

    m0m\equiv 0 in ([a,b])×[0,T](\mathbb{R}\setminus[a,b])\times[0,T], and m0m_{0} is compactly supported in [a,b][a,b].

Condition (U1) penalize being in [a,b]c[a,b]^{c} ensuring that optimal controls α\alpha in (18) are such that xihα±hσr[a,b]x_{i}-h\alpha\pm\sqrt{h}\sigma_{r}\in[a,b]. Moreover, the contributions to non-local operators of uu from [a,b]c[a,b]^{c} will be small away from the boundary. Condition (M1) ensures that the mass of mm is essentially contained in [a,b][a,b] up to some finite time (but some mass will leak out due to nonlocal effects), and there is no contribution from [a,b]c[a,b]^{c} when we compute non-local operators of mm. We will present numerical results from a region of interest that is far away from the boundary of [a,b][a,b], and where the influence of the (artificial) exterior data is expected to be negligible.

Evaluating the integrals

To implement the scheme, we need to evaluate the integral

|z|rI[f](xi+z)ν(dz)=jf[xi]ωji,ν,\displaystyle\int_{|z|\geq r}I[f](x_{i}+z)\nu(dz)=\sum_{j\in\mathbb{Z}}f[x_{i}]\omega_{j-i,\nu},

where

ωji,ν=|z|rβji(z)ν(dz),\displaystyle\omega_{j-i,\nu}=\int_{|z|\geq r}\beta_{j-i}(z)\nu(dz),

see (17). In addition, we need to compute the values of σr,br\sigma_{r},b_{r}, and λr\lambda_{r} (see (9), (8), and (11)). To compute the weights ωji,ν\omega_{j-i,\nu} we use two different methods. For the fractional Laplacians, we use the explicit weights of [45], while for CGMY diffusions we calculate the weights numerically using the inbuilt integral function in MATLAB. When tested on the fractional Laplacian, the MATLAB integrator produced an error of less than 101510^{-15}. Below the quantities σr,br,λr\sigma_{r},b_{r},\lambda_{r} are computed explicitly, except in the CGMY case where we use numerical integration.

Solving the coupled system

We use a fixed point iteration scheme: (i) Let μ=m0\mu=m_{0}, and solve for uρ,hu_{\rho,h} in (18)–(20). (ii) With approximate optimal control Duρ,hϵDu_{\rho,h}^{\epsilon} as in (21), we solve for mρ,hϵm_{\rho,h}^{\epsilon} in (24). (iii) Let μnew=(mρ,hϵ+μ)/2\mu_{\text{new}}=(m_{\rho,h}^{\epsilon}+\mu)/2, and repeat the process with μ=μnew\mu=\mu_{\text{new}}. We continue until we have converged to a fixed point to within machine accuracy.

Remark 8.1.

Instead μnew=mρ,hϵ\mu_{\text{new}}=m_{\rho,h}^{\epsilon}, we take μnew=(mρ,hϵ+μ)/2\mu_{\text{new}}=(m_{\rho,h}^{\epsilon}+\mu)/2. I.e. we use a fixed point iteration with some memory. This gives much faster convergence in our examples.

Example 1.

Problem (45) with [0,T]×[a,b]=[0,2]×[0,1][0,T]\times[a,b]=[0,2]\times[0,1], G=0G=0, f(t,x)=5(x0.5(1sin(2πt)))2f(t,x)=5(x-0.5(1-\sin(2\pi t)))^{2}, m0(x)=Ce(x0.5)20.12m_{0}(x)=Ce^{-\frac{(x-0.5)^{2}}{0.1^{2}}}, where CC is such that abm0=1\int_{a}^{b}m_{0}=1. Furthermore, in accordance with the CFL-conditions of Theorem 4.1, we let h=ρ=0.005h=\rho=0.005, r=h12sr=h^{\frac{1}{2s}}, ϵ=h0.0707\epsilon=\sqrt{h}\approx 0.0707, σ=0.09\sigma=0.09, δ=0.4\delta=0.4, K=1K=1.

For the diffusions, we consider =(Δ)s2\mathcal{L}=(-\Delta)^{\frac{s}{2}} for s=0.5,1.5,1.9s=0.5,1.5,1.9, =Δ\mathcal{L}=\Delta, and 0\mathcal{L}\equiv 0. In figure 1 we plot the different solutions at time t=0.5t=0.5 and t=1.5t=1.5.

Refer to caption
(a) t=0.5t=0.5
Refer to caption
(b) t=1.5t=1.5
Figure 1. The solutions mm in Example 1.

In figure 2 we plot the solution with s=1.5s=1.5 on the time interval [0,2][0,2].

Refer to caption
(a) m(t,x)m(t,x)
Refer to caption
(b) u(t,x)u(t,x)
Figure 2. Solution mm and uu in Example 1 with diffusion parameter s=1.5s=1.5
Example 2.

Problem (45) with the same cost functions as in Example 1, but different diffusions with parameter s=1.5s=1.5:

  1. (i)

    =σ2(Δ)s2,\mathcal{L}=\sigma^{2}(-\Delta)^{\frac{s}{2}},

  2. (ii)

    =σ2Cd,s[u(x+y)u(x)Du(x)y𝟙|y|<1] 1[0,+)dy|y|1+s\mathcal{L}=\sigma^{2}C_{d,s}\int_{\mathbb{R}}[u(x+y)-u(x)-Du(x)\cdot y\mathbbm{1}_{|y|<1}]\,\mathbbm{1}_{[0,+\infty)}\,\frac{dy}{|y|^{1+s}},

  3. (iii)

    =σ2Cd,s[u(x+y)u(x)Du(x)y𝟙|y|<1] 1[0.5,0.5]cdy|y|1+s\mathcal{L}=\sigma^{2}C_{d,s}\int_{\mathbb{R}}[u(x+y)-u(x)-Du(x)\cdot y\mathbbm{1}_{|y|<1}]\,\mathbbm{1}_{[-0.5,0.5]^{c}}\,\frac{dy}{|y|^{1+s}},

  4. (iv)

    =σ2Cd,s[u(x+y)u(x)Du(x)y𝟙|y|<1]e10yy+dy|y|1+s\mathcal{L}=\sigma^{2}C_{d,s}\int_{\mathbb{R}}[u(x+y)-u(x)-Du(x)\cdot y\mathbbm{1}_{|y|<1}]\,e^{-10y^{-}-y^{+}}\,\frac{dy}{|y|^{1+s}},

where Cd,sC_{d,s} is the normalizing constant for the fractional Laplacian (see [45]). Case (i) is the reference solution, a symmetric and uniformly elliptic operator. Case (ii) is non-symmetric and non-degenerate, case (iii) is symmetric and degenerate, and case (iv) is a CGMY-diffusion (see e.g. [34]). We have plotted mm at t=0.5t=0.5 and t=1.5t=1.5 in Figure 3.

Refer to caption
(a) t=0.5t=0.5
Refer to caption
(b) t=1.5t=1.5
Figure 3. The solutions mm in Example 2
Example 3.

(Long time behaviour). Under certain conditions (see e.g. [22, 21]), the solution of time dependent MFG systems will quickly converge to the solution of the corresponding stationary ergodic MFG system, as the time horizon TT increases. We check numerically that this is also the case for nonlocal diffusions. In (45), we take =(Δ)s2\mathcal{L}=(-\Delta)^{\frac{s}{2}}, with s=1.5s=1.5, [0,T]×[a,b]=[0,10]×[1,2][0,T]\times[a,b]=[0,10]\times[-1,2], G(x)=(x2)2G(x)=(x-2)^{2}, f(t,x)=x2f(t,x)=x^{2}, and m0(x)=𝟙[1,2](x)m_{0}(x)=\mathbbm{1}_{[1,2]}(x). We expect (from the cost functions ff and GG) that the solution mm will approach the line x=0x=0 quite fast, and then travel along this line, until it goes towards the point x=2x=2 in the very end. Our numerical simulations shows that this is the case also for nonlocal diffusions. Here we have considered the cases K=0K=0 (no coupling in the uu equation) and K=0.4K=0.4 (some coupling). The parameters used in the simulations are h=ρ=0.01h=\rho=0.01, ϵ=h\epsilon=\sqrt{h}, r=h1/2sr=h^{1/2s}, and the results are shown in Figure 4.

Refer to caption
(a) ff
Refer to caption
(b) f+0.4ϕδf+0.4\phi_{\delta}
Figure 4. The solutions mm in Example 3

The players want to avoid each other in the case of K=0.4K=0.4, so the solution is more spread out in space direction than in the case of K=0K=0.

Example 4.

We compute the convergence rate when ff, GG, m0m_{0} are as in Example 1, s=1.5s=1.5, ν=0.2\nu=0.2, δ=0.4\delta=0.4, and the domain [0,T]×[a,b]=[0,0.5]×[0,1][0,T]\times[a,b]=[0,0.5]\times[0,1]. We take ρ=h\rho=h, r=h12sr=h^{\frac{1}{2s}}, and for simplicity ϵ=0.25\epsilon=0.25.

We calculate solutions for different values of hh, and compare with a reference solution computed at h=210h=2^{-10}. We calculate LL^{\infty} and L1L^{1} relative errors restricted to the xx-interval [13,23][\frac{1}{3},\frac{2}{3}] (to avoid boundary effects), and t=0t=0 for uu and t=Tt=T for mm:

ERRu:=uρ,h(0,)uref(0,)L(13,23)uref(0,)L(13,23),ERRm:=mρ,hϵ(T,)mref(T,)L1(13,23)mref(T,)L1(13,23).\displaystyle\textup{ERR}_{u}:=\frac{\|u_{\rho,h}(0,\cdot)-u_{\text{ref}}(0,\cdot)\|_{L^{\infty}(\frac{1}{3},\frac{2}{3})}}{\|u_{\text{ref}}(0,\cdot)\|_{L^{\infty}(\frac{1}{3},\frac{2}{3})}},\quad\textup{ERR}_{m}:=\frac{\|m_{\rho,h}^{\epsilon}(T,\cdot)-m_{\text{ref}}(T,\cdot)\|_{L^{1}(\frac{1}{3},\frac{2}{3})}}{\|m_{\text{ref}}(T,\cdot)\|_{L^{1}(\frac{1}{3},\frac{2}{3})}}.

The results are given in the table below.

h 222^{-2} 232^{-3} 242^{-4} 252^{-5} 262^{-6} 272^{-7} 282^{-8} 292^{-9}
ERRu 0.3155 0.1951 0.0920 0.0446 0.0218 0.0097 0.0035 0.0013
ERRm 0.8055 0.4583 0.2886 0.1869 0.1023 0.0596 0.0300 0.0186

We see that when we halve hh, the error is halved, i.e we observe an error of order O(h)O(h).

Appendix A Proof of Proposition 3.4

The proof is an adaptation of the Schauder fixed point argument used to prove existence for MFGs. We will use a direct consequence of Theorem 6.4 and 6.5:

Corollary A.1.

Assume (ν\nu0): ,(ν\nu1): , (L1): (L2): , (H1): , (F2): , (M): , Ψ\Psi is given by Proposition 6.1, and mρ,hϵ[μ]m^{\epsilon}_{\rho,h}[\mu] is defined by (26). Then there is Cρ,h,ϵ>0C_{\rho,h,\epsilon}>0, such that for any μC([0,T],P(d))\mu\in C([0,T],P(\mathbb{R}^{d})) and t,s[0,T]t,s\in[0,T],

dΨ(x)𝑑mρ,hϵ[μ](t)+d0(mρ,hϵ[μ](t),mρ,hϵ[μ](s))|ts|Cρ,h,ϵ.\displaystyle\int_{\mathbb{R}^{d}}\Psi(x)\,dm^{\epsilon}_{\rho,h}[\mu](t)+\frac{d_{0}(m^{\epsilon}_{\rho,h}[\mu](t),m^{\epsilon}_{\rho,h}[\mu](s))}{\sqrt{|t-s|}}\leq C_{\rho,h,\epsilon}.

The point is that ρ,h,ϵ\rho,h,\epsilon are fixed in this result. Let

𝒞:={\displaystyle\mathcal{C}:=\Big{\{} μC(0,T;P(d)):μ(0)=m0,\displaystyle\mu\in C(0,T;P(\mathbb{R}^{d})):\mu(0)=m_{0},
supt,s[0,T][dψ(x)dμ(t,x)+d0(μ(t),μ(s))|ts|]Cρ,h,ϵ},\displaystyle\quad\sup_{t,s\in[0,T]}\Big{[}\int_{\mathbb{R}^{d}}\psi(x)d\mu(t,x)+\frac{d_{0}(\mu(t),\mu(s))}{\sqrt{|t-s|}}\Big{]}\leq C_{\rho,h,\epsilon}\Big{\}},

where Cρ,h,ϵC_{\rho,h,\epsilon} is defined in Corollary A.1. For μ𝒞\mu\in\mathcal{C}, let uρ,h[μ]u_{\rho,h}[\mu] be solution of (18) and uρ,hϵ[μ]u_{\rho,h}^{\epsilon}[\mu] defined by (22). Then mρ,hϵ=S(μ)m_{\rho,h}^{\epsilon}=S(\mu) is defined to the corresponding solution of (24). Note that a fixed point of SS will give a solution (u,m)(u,m) of the scheme (27). We now conclude the proof by applying Schauder’s fixed point theorem since:

1.  (𝒞\mathcal{C} is a convex, closed, compact set). It is a convex and closed by standard arguments and compact by the Prokhorov and Arzelà-Ascoli theorems.

2.  (SS is a self-map on 𝒞\mathcal{C}). The map SS maps 𝒞\mathcal{C} into itself by Corollary A.1 (tightness and equicontinuity), and Lemma 3.3 (positivity and mass preservation).

3.  (SS is continuous). Let μnμ\mu_{n}\to\mu in 𝒞\mathcal{C}. By Theorem 5.2 (comparison) and (F2): ,

uρ,h[μn]uρ,h[μ]0\displaystyle\|u_{\rho,h}[\mu_{n}]-u_{\rho,h}[\mu]\|_{0}
Tsupt,x|F(x,μn(t))F(x,μ(t))|+supx|G(x,μn(T))G(x,μ(T))|\displaystyle\leq T\sup_{t,x}|F(x,\mu_{n}(t))-F(x,\mu(t))|+\sup_{x}|G(x,\mu_{n}(T))-G(x,\mu(T))|
TLFsuptd0(μn(t),μ(t))+LGd0(μn(T),μ(T))0.\displaystyle\leq TL_{F}\,\sup_{t}d_{0}(\mu_{n}(t),\mu(t))+L_{G}\,d_{0}(\mu_{n}(T),\mu(T))\to 0.

Then supi|ui,k[μn]uij,k[μn]ρui,k[μ]uij,k[μ]ρ|0\sup_{i}\big{|}\frac{u_{i,k}[\mu_{n}]-u_{i-j,k}[\mu_{n}]}{\rho}-\frac{u_{i,k}[\mu]-u_{i-j,k}[\mu]}{\rho}\big{|}\to 0 uniformly for |ij|=1|i-j|=1, Duρ,hϵ[μn]Duρ,hϵ[μ]00\|Du_{\rho,h}^{\epsilon}[\mu_{n}]-Du_{\rho,h}^{\epsilon}[\mu]\|_{0}\to 0, and finally by Lemma 6.6,

supt[0,T]mρ,hϵ[μn](t,)mρ,hϵ[μ](t,)L1(d)cKTρehλrDuρ,hϵ[μn]Duρ,hϵ[μ]00.\displaystyle\sup_{t\in[0,T]}\|m_{\rho,h}^{\epsilon}[\mu_{n}](t,\cdot)-m_{\rho,h}^{\epsilon}[\mu](t,\cdot)\|_{L^{1}(\mathbb{R}^{d})}\leq\frac{cKT}{\rho}e^{-h\lambda_{r}}\|Du_{\rho,h}^{\epsilon}[\mu_{n}]-Du_{\rho,h}^{\epsilon}[\mu]\|_{0}\to 0.

Hence 𝒮\mathcal{S} is continuous.

Appendix B Proof of Lemma 5.6 (ii) and (iii)

Fix (t,x)[0,T]×d(t,x)\in[0,T]\times\mathbb{R}^{d} and consider a sequence (tk,xk)(t,x)(t_{k},x_{k})\to(t,x). For any ydy\in\mathbb{R}^{d}, a Taylor expansion shows that

(46) uρn,hnϵn[μn](tk,xk+y)uρn,hnϵn[μn](tk,xk)Duρn,hnϵn[μn](tk,xk)y=01(Duρn,hnϵn[μn](tk,xk+sy)Duρn,hnϵn[μn](tk,xk))y𝑑s:=01I(s)y𝑑s.\displaystyle\begin{split}&u_{\rho_{n},h_{n}}^{\epsilon_{n}}[\mu_{n}](t_{k},x_{k}+y)-u_{\rho_{n},h_{n}}^{\epsilon_{n}}[\mu_{n}](t_{k},x_{k})-Du_{\rho_{n},h_{n}}^{\epsilon_{n}}[\mu_{n}](t_{k},x_{k})\cdot y\\ &=\int_{0}^{1}\big{(}Du_{\rho_{n},h_{n}}^{\epsilon_{n}}[\mu_{n}](t_{k},x_{k}+sy)-Du_{\rho_{n},h_{n}}^{\epsilon_{n}}[\mu_{n}](t_{k},x_{k})\big{)}\cdot y\,ds:=\int_{0}^{1}I(s)\cdot y\,ds.\end{split}

Using first Lemma 5.5 (a) and then part two of Lemma 5.5 (b), we find that

0ρnϵn|y|I(s)y𝑑s\displaystyle\int_{0}^{\frac{\rho_{n}}{\epsilon_{n}|y|}}I(s)\cdot y\,ds 2Duρn,hnϵn[μn]0ρnϵn2((LL+LF)T+LG)ρnϵn\displaystyle\leq 2\|Du_{\rho_{n},h_{n}}^{\epsilon_{n}}[\mu_{n}]\|_{0}\frac{\rho_{n}}{\epsilon_{n}}\leq 2((L_{L}+L_{F})T+L_{G})\frac{\rho_{n}}{\epsilon_{n}}
ρnϵn|y|1I(s)y𝑑s\displaystyle\int^{1}_{\frac{\rho_{n}}{\epsilon_{n}|y|}}I(s)\cdot y\,ds c1ρnϵn|y|11s(|sy|2+ρn2ϵn2)𝑑s=c1|y|2ρnϵn|y|1(s+1sρn2|y|2ϵn2)𝑑s\displaystyle\leq c_{1}\int^{1}_{\frac{\rho_{n}}{\epsilon_{n}|y|}}\frac{1}{s}\Big{(}|sy|^{2}+\frac{\rho_{n}^{2}}{\epsilon_{n}^{2}}\Big{)}ds=c_{1}|y|^{2}\int^{1}_{\frac{\rho_{n}}{\epsilon_{n}|y|}}\Big{(}s\,+\frac{1}{s}\frac{\rho_{n}^{2}}{|y|^{2}\epsilon_{n}^{2}}\Big{)}\,ds
c1|y|2ρnϵn|y|1(s+1ss2)𝑑sc1|y|2.\displaystyle\leq c_{1}|y|^{2}\int^{1}_{\frac{\rho_{n}}{\epsilon_{n}|y|}}\Big{(}s\,+\frac{1}{s}s^{2}\Big{)}\,ds\leq c_{1}|y|^{2}.

By Lemma 5.5 (a), the sequence Duρn,hnϵn[μn](tk,xk)Du_{\rho_{n},h_{n}}^{\epsilon_{n}}[\mu_{n}](t_{k},x_{k}) is precompact. Now take any convergent subsequence as n,kn,k\to\infty and ρnϵn=o(1)\frac{\rho_{n}}{\epsilon_{n}}=o(1). If pp is the limit, then by passing to the limit in (46) along this subsequence we have

u[μ](x+y)u[μ](x)pyc1|y|2for everyyd,\displaystyle u[\mu](x+y)-u[\mu](x)-p\cdot y\leq c_{1}|y|^{2}\quad\text{for every}\quad y\in\mathbb{R}^{d},

and pD+u[μ](t,x)p\in D^{+}u[\mu](t,x), the superdifferential of u[μ](t,x)u[\mu](t,x). At points (x,t)(x,t) where u[μ]u[\mu] is differentiable, D+u[μ](t,x)={Du[μ](t,x)}D^{+}u[\mu](t,x)=\{Du[\mu](t,x)\} and p=Du[μ](t,x)p=Du[\mu](t,x), and then since the subsequence was arbitrary in the above argument and all limit points pp coincide,

(47) lim sup(tk,xk)(t,x),nDuρn,hnϵn[μn](tk,xk)=lim inf(tk,xk)(t,x),nDuρn,hnϵn[μn](tk,xk)=Du(t,x).\displaystyle\begin{split}&\limsup_{(t_{k},x_{k})\to(t,x),n\to\infty}Du_{\rho_{n},h_{n}}^{\epsilon_{n}}[\mu_{n}](t_{k},x_{k})\\ &\qquad=\liminf_{(t_{k},x_{k})\to(t,x),n\to\infty}Du_{\rho_{n},h_{n}}^{\epsilon_{n}}[\mu_{n}](t_{k},x_{k})\\ &\qquad=Du(t,x).\end{split}

We conclude that Duρn,hnϵn[μn]Du[μ]Du_{\rho_{n},h_{n}}^{\epsilon_{n}}[\mu_{n}]\rightarrow Du[\mu] at (t,x)(t,x). Part (ii) now follows since u[μ]u[\mu] is Lipschitz in space by Proposition 2.5 (c) and then xx-differentiable for a.e. xx and every tt.

To prove part (iii), we note that uu is C1C^{1} by (U): , so now (47) holds for every (t,x)(t,x). Then in view of the uniform Lipschitz estimate from Lemma 5.5 (a), local uniform convergence follows from [11, Chapter V, Lemma 1.9]. The proof is complete.

Appendix C Proof of Lemma 6.9

We first show strong separation between any two characteristics Φϵ,±\Phi^{\epsilon,\pm}: By Lemma 6.8,

|Φj,kϵ,±Φi,kϵ,±|2=|xjxi±hσrhσrh(DpH(xj,Duρ,hϵ(tk,xj))+Brσ\displaystyle\big{|}\Phi^{\epsilon,\pm}_{j,k}-\Phi^{\epsilon,\pm}_{i,k}\big{|}^{2}=\,\Big{|}x_{j}-x_{i}\pm\sqrt{h}\sigma_{r}\mp\sqrt{h}\sigma_{r}-h\Big{(}D_{p}H(x_{j},Du^{\epsilon}_{\rho,h}(t_{k},x_{j}))+B^{\sigma}_{r}
DpH(xi,Duρ,hϵ(tk,xi))Brσ)|2\displaystyle\hskip 142.26378pt-D_{p}H(x_{i},Du^{\epsilon}_{\rho,h}(t_{k},x_{i}))-B^{\sigma}_{r}\Big{)}\Big{|}^{2}
|xjxi|22h(DpH(xj,Duρ,hϵ(tk,xj))DpH(xi,Duρ,hϵ(tk,xi)))(xjxi)\displaystyle\geq\,|x_{j}-x_{i}|^{2}-2h\Big{(}D_{p}H\big{(}x_{j},Du^{\epsilon}_{\rho,h}(t_{k},x_{j})\big{)}-D_{p}H\big{(}x_{i},Du^{\epsilon}_{\rho,h}(t_{k},x_{i})\big{)}\Big{)}(x_{j}-x_{i})
(1c0h)|xjxi|2.\displaystyle\geq\,(1-c_{0}h)|x_{j}-x_{i}|^{2}.

Hence, we have

(48) min{|Φj,kϵ,+Φi,kϵ,+|,|Φj,kϵ,Φi,kϵ,|}1c0h|ji|ρ>ρ1c0h.\displaystyle\min\Big{\{}\big{|}\Phi^{\epsilon,+}_{j,k}-\Phi^{\epsilon,+}_{i,k}\big{|},\big{|}\Phi^{\epsilon,-}_{j,k}-\Phi^{\epsilon,-}_{i,k}\big{|}\Big{\}}\geq\sqrt{1-c_{0}h}|j-i|\rho>\rho\sqrt{1-c_{0}h}.

The result now holds following the proof of [23, Lemma 3.8]. We give the proof for completeness.

Since the diameter of the support of a (hat) basis functions βi\beta_{i} is 2ρ2\rho, by (48) there can be at most 3 characteristics inside the supp(βi)\textup{supp}(\beta_{i}) for small enough hh. The result is trivial if there is only one in characteristic supp(βi)\textup{supp}(\beta_{i}). When supp(βi)\textup{supp}(\beta_{i}) contains 2 characteristics, say Φj1,kϵ,+\Phi^{\epsilon,+}_{j_{1},k} and Φj2,kϵ,+\Phi^{\epsilon,+}_{j_{2},k}, we see by (48) (check the different orderings of xkx_{k}, Φj1,kϵ,+\Phi^{\epsilon,+}_{j_{1},k}, Φj2,kϵ,+\Phi^{\epsilon,+}_{j_{2},k}) that

βi(Φj1,kϵ,+)+βi(Φj2,kϵ,+)=\displaystyle\beta_{i}(\Phi^{\epsilon,+}_{j_{1},k})+\beta_{i}(\Phi^{\epsilon,+}_{j_{2},k})=  1|xiΦj1,kϵ,+|ρ+1|xiΦj2,kϵ,+|ρ\displaystyle\,1-\frac{\Big{|}x_{i}-\Phi^{\epsilon,+}_{j_{1},k}\Big{|}}{\rho}+1-\frac{\Big{|}x_{i}-\Phi^{\epsilon,+}_{j_{2},k}\Big{|}}{\rho}
\displaystyle\leq  2|Φj1,kϵ,+Φj2,kϵ,+|ρ21c0h 1+K0h.\displaystyle\,2-\frac{\Big{|}\Phi^{\epsilon,+}_{j_{1},k}-\Phi^{\epsilon,+}_{j_{2},k}\Big{|}}{\rho}\leq 2-\sqrt{1-c_{0}h}\leq\,1+K_{0}h.

Finally, assume support(βi)support(\beta_{i}) contains 3 characteristics Φj1,kϵ,+,Φj2,kϵ,+\Phi^{\epsilon,+}_{j_{1},k},\Phi^{\epsilon,+}_{j_{2},k} and Φj3,kϵ,+\Phi^{\epsilon,+}_{j_{3},k}. By (48) that all three characteristics can not be on one side (left or right) of xix_{i}. Without loss of generality we assume Φj1,kϵ,+<xi<Φj2,kϵ,+<Φj3,kϵ,+\Phi^{\epsilon,+}_{j_{1},k}<x_{i}<\Phi^{\epsilon,+}_{j_{2},k}<\Phi^{\epsilon,+}_{j_{3},k}, and find

βi(Φj1,kϵ,+)+βi(Φj2,kϵ,+)+βi(Φj3,kϵ,+)=1xiΦj1,kϵ,+ρ+1Φj2,kϵ,+xiρ+1Φj3,kϵ,+xiρ\displaystyle\beta_{i}(\Phi^{\epsilon,+}_{j_{1},k})+\beta_{i}(\Phi^{\epsilon,+}_{j_{2},k})+\beta_{i}(\Phi^{\epsilon,+}_{j_{3},k})=1-\frac{x_{i}-\Phi^{\epsilon,+}_{j_{1},k}}{\rho}+1-\frac{\Phi^{\epsilon,+}_{j_{2},k}-x_{i}}{\rho}+1-\frac{\Phi^{\epsilon,+}_{j_{3},k}-x_{i}}{\rho}
 3Φj2,kϵ,+Φj1,kϵ,+ρΦj3,kϵ,+Φj2,kϵ,+ρ\displaystyle\leq\,3-\frac{\Phi^{\epsilon,+}_{j_{2},k}-\Phi^{\epsilon,+}_{j_{1},k}}{\rho}-\frac{\Phi^{\epsilon,+}_{j_{3},k}-\Phi^{\epsilon,+}_{j_{2},k}}{\rho}
 321c0h1+2(11c0h)1+K0h.\displaystyle\leq\,3-2\sqrt{1-c_{0}h}\leq 1+2(1-\sqrt{1-c_{0}h})\leq 1+K_{0}h.

Combining all three cases we get

jβi(Φj,kϵ,+)1+K0hfor anyi.\sum_{j\in\mathbb{Z}}\beta_{i}(\Phi^{\epsilon,+}_{j,k})\leq 1+K_{0}h\quad\mbox{for any}\,i\in\mathbb{Z}.

The estimate of jβi(Φj,kϵ,)\sum_{j\in\mathbb{Z}}\beta_{i}(\Phi^{\epsilon,-}_{j,k}) is similar. This completes the proof.

Acknowledgements

The authors are supported by the Toppforsk (research excellence) project Waves and Nonlinear Phenomena (WaNP), grant no. 250070 from the Research Council of Norway. IC is partially supported by the Croatian Science Foundation under the project 4197. The authors would like to thank Elisabetta Carlini for sharing the code of the numerical methods introduced in [23].

References

  • [1] Y. Achdou, F. Camilli, and I. Capuzzo-Dolcetta. Mean Field Games: Numerical methods for the planning problem. SIAM J. Control Optim, 50(1):77–109, 2012.
  • [2] Y. Achdou, F. Camilli, and I. Capuzzo-Dolcetta. Mean Field Games: Convergence of a finite difference method. SIAM J. Numer. Anal., 51(5):2585–2612, 2013.
  • [3] Y. Achdou, F. Camilli, and L. Corrias. On numerical approximation of the Hamilton-Jacobi-transport system arising in high frequency approximations. Discrete Contin. Dyn. Syst. Ser. B, 19(3):629–650, 2014.
  • [4] Y. Achdou and I. Capuzzo-Dolcetta. Mean Field Games: Numerical methods. SIAM J. Numer. Anal., 48(3):1136–1162, 2010.
  • [5] Y. Achdou, P. Cardaliaguet, F. Delarue, A. Porretta, F. Santambrogio. Mean field games. Lecture Notes in Mathematics, CIME vol. 2281, Springer, 2020.
  • [6] Y. Achdou and M. Laurière. Mean Field Games and applications: Numerical aspects. arXiv preprint arXiv:2003.04444, 2020.
  • [7] Y. Achdou and V. Perez. Iterative strategies for solving linearized discrete Mean Field Games systems. Netw. Heterog. Media, 7(2):197, 2012.
  • [8] Y. Achdou and A. Porretta. Convergence of a finite difference scheme to weak solutions of the system of partial differential equations arising in mean field games. SIAM J. Numer. Anal., 54(1):161–186, 2016.
  • [9] D. Applebaum. Lévy processes and stochastic calculus. Cambridge university press, 2009.
  • [10] S. Asmussen and J. Rosiński. Approximations of small jumps of Lévy processes with a view towards simulation. J. Appl. Probab., 38(2):482–493, 2001.
  • [11] M. Bardi and I. Capuzzo-Dolcetta. Optimal Control and Viscosity Solutions of Hamilton-Jacobi-Bellman Equations. Birkhäuser, 1997.
  • [12] G. Barles and P. E. Souganidis. Convergence of approximation schemes for fully nonlinear second order equations. Asymptotic Anal., 4(3):271–283, 1991.
  • [13] A. Bensoussan, J. Frehse, and P. Yam Mean field games and mean field type control theory SpringerBriefs in Mathematics, Springer, 2013.
  • [14] F. Camilli and M. Falcone. An approximation scheme for the optimal control of diffusion processes. RAIRO Modél. Math. Anal. Numér., 29(1):97–122, 1995.
  • [15] F. Camilli and E. R. Jakobsen. A finite element like scheme for integro-partial differential Hamilton–Jacobi–Bellman equations. SIAM J. Numer. Anal., 47(4):2407–2431, 2009.
  • [16] I. Capuzzo Dolcetta. On a discrete approximation of the Hamilton-Jacobi equation of dynamic programming. Appl. Math. Optim. 10(4): 367–377, 1983.
  • [17] P. Cardaliaguet. Weak solutions for first order Mean Field Games with local coupling. In Analysis and geometry in control theory and its applications, volume 11 of Springer INdAM Ser., pages 111–158. Springer, Cham, 2015.
  • [18] P. Cardaliaguet, F. Delarue, J.-M. Lasry, and P.-L. Lions. The Master Equation and the Convergence Problem in Mean Field Games. Annals of Mathematics Studies, vol. 201, Princeton University Press, 2019.
  • [19] P. Cardaliaguet and P. J. Graber. Mean Field Games systems of first order. ESAIM Control Optim. Calc. Var., 21(3):690–722, 2015.
  • [20] P. Cardaliaguet, P. J. Graber, A. Porretta, and D. Tonon. Second order Mean Field Games with degenerate diffusion and local coupling. NoDEA Nonlinear Differential Equations Appl., 22(5):1287–1317, 2015.
  • [21] 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.
  • [22] P. Cardaliaguet, J.-M. Lasry, P.-L. Lions, and A. Porretta. Long time average of mean field games. Netw. Heterog. Media, 7(2):279, 2012.
  • [23] E. Carlini and F. J. Silva. A fully discrete semi-Lagrangian scheme for a first order Mean Field Game problem. SIAM J. Numer. Anal., 52(1):45–67, 2014.
  • [24] E. Carlini and F. J. Silva. A semi-Lagrangian scheme for a degenerate second order Mean Field Game system. Discrete Contin. Dyn. Syst., 35(9):4269, 2015.
  • [25] E. Carlini and F. J. Silva. On the discretization of some nonlinear Fokker-Planck-Kolmogorov equations and applications. SIAM J. Numer. Anal., 56(4):2148–2177, 2018.
  • [26] R. Carmona and F. Delarue. Probabilistic analysis of Mean-Field Games. SIAM J. Control Optim., 51(4):2705–2734, 2013.
  • [27] R. Carmona and F. Delarue. Probabilistic theory of mean field games with applications. I-II, Probability Theory and Stochastic Modelling 84, Springer 2018.
  • [28] R. Carmona and M. Laurière. Convergence analysis of machine learning algorithms for the numerical solution of Mean Field Control and Games: II–the finite horizon case. arXiv preprint arXiv:1908.01613, 2019.
  • [29] R. Carmona, M. Laurière, and Z. Tan. Linear-quadratic Mean-Field reinforcement learning: Convergence of policy gradient methods. arXiv preprint arXiv:1910.04295, 2019.
  • [30] A. Cesaroni, M. Cirant, S. Dipierro, M. Novaga, and E. Valdinoci. On stationary fractional Mean Field Games. J. Math. Pures Appl., 122(9):1-22, 2017.
  • [31] I. Chowdhury, E. R. Jakobsen, and M. Krupski. On fully nonlinear parabolic mean field games with examples of nonlocal and local diffusions. arXiv preprint arXiv:2104.06985, 2021.
  • [32] M. Cirant. On the solvability of some ergodic control problems in d\mathbb{R}^{d}. SIAM J. Control Optim., 52(6):4001–4026, 2014.
  • [33] M. Cirant and A. Goffi. On the existence and uniqueness of solutions to time-dependent fractional MFG. SIAM J. Math. Anal., 51(2):913–954, 2019.
  • [34] R. Cont and P. Tankov. Financial modelling with jump processes. CRC press, 2003.
  • [35] K. Debrabant and E. R. Jakobsen. Semi-Lagrangian schemes for linear and fully non-linear diffusion equations. Math. Comp., 82(283):1433–1462, 2013.
  • [36] S. Elghanjaoui and K. H. Karlsen. A markov chain approximation scheme for a singular investment-consumption problem with Lévy driven stock prices. Online available url: https://citeseerx.ist.psu.edu/viewdoc/download?doi=10.1.1.15.2182&\&rep=rep1&\&type=pdf, 2002.
  • [37] O. Ersland and E. R. Jakobsen. On fractional and nonlocal parabolic Mean Field Games in the whole space. arXiv preprint arXiv:2003.12302, 2020.
  • [38] M. Falcone and R. Ferretti. Semi-Lagrangian approximation schemes for linear and Hamilton-Jacobi equations. Society for Industrial and Applied Mathematics (SIAM), 2014.
  • [39] D. A. Gomes, E. A. Pimentel, and V. Voskanyan. Regularity theory for mean-field game systems. SpringerBriefs in Mathematics, Springer, 2016
  • [40] O. Guéant. Mean Field Games equations with quadratic Hamiltonian: a specific approach. Math. Models Methods Appl. Sci., 22(09), 2012.
  • [41] O. Guéant. New numerical methods for Mean Field Games with quadratic costs. Netw. Heterog. Media, 7(2):315-336, 2012.
  • [42] O. Guéant. Mean Field Games with a quadratic Hamiltonian: a constructive scheme. Advances in dynamic games, pages 229–241. Springer, 2013.
  • [43] O. Guéant, J.-M. Lasry, and P.-L. Lions. Mean field games and applications, in Paris–Princeton Lectures on Mathematical Finance 2010. Lecture Notes in Mathematics 2003, 205–266, Springer, 2011.
  • [44] 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.
  • [45] Y. Huang and A. Oberman. Numerical methods for the fractional Laplacian: A finite difference-quadrature approach. SIAM J. Numer. Anal., 52(6):3056–3084, 2014.
  • [46] E. R. Jakobsen and K. H. Karlsen. Continuous dependence estimates for viscosity solutions of integro-PDEs. J. Differential Equations, 212(2):278–318, 2005.
  • [47] E. R. Jakobsen, K. H. Karlsen and C. La Chioma. Error estimates for approximate solutions to Bellman equations associated with controlled jump-diffusions. Numer. Math., 110(2):221–255, 2008.
  • [48] V. N. Kolokoltsov, M. S. Troeva, and W. Yang. Mean Field Games based on stable-like processes. Autom. Remote Control, 77(11):2044–2064, 2016.
  • [49] J.-M. Lasry and P.-L. Lions. Mean Field Games. Jpn. J. Math., 2(1):229–260, 2007.
  • [50] R. Metzler and J. Klafter. The random walk’s guide to anomalous diffusion: a fractional dynamics approach. Phys. Rep., 339(1):1–77, 2000.
  • [51] A. Porretta. Weak solutions to Fokker-Planck equations and Mean Field Games. Arch. Ration. Mech. Anal., 216(1):1–62, 2015.
  • [52] L. Ruthotto, S. J. Osher, W. Li, L. Nurbekyan, and S. W. Fung. A machine learning framework for solving high-dimensional Mean Field Game and Mean Field Control problems. Proc. Natl. Acad. Sci. U.S.A., 117(17):9183–9193, 2020.
  • [53] C. Villani. Optimal transport: Old and new. Springer Science & Business Media, 2008.
  • [54] W. A. Woyczyński. Lévy processes in the physical sciences. Lévy processes, pages 241–266. Springer, 2001.