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

Graphon Mean Field Games and the GMFG Equations

Peter E. Caines Department of Electrical and Computer Engineering, McGill University, Montreal, QC, Canada peterc@cim.mcgill.ca  and  Minyi Huang School of Mathematics and Statistics, Carleton University, Ottawa, ON, Canada mhuang@math.carleton.ca
(Date: Aug 24, 2020; revised Apr 12, 2021, Jun 15, 2021, Dec 28, 2021.)
Abstract.

The emergence of the graphon theory of large networks and their infinite limits has enabled the formulation of a theory of the centralized control of dynamical systems distributed on asymptotically infinite networks [16, 19]. Furthermore, the study of the decentralized control of such systems was initiated in [6, 7], where Graphon Mean Field Games (GMFG) and the GMFG equations were formulated for the analysis of non-cooperative dynamic games on unbounded networks. In that work, existence and uniqueness results were introduced for the GMFG equations, together with an ϵ\epsilon-Nash theory for GMFG systems which relates infinite population equilibria on infinite networks to finite population equilibria on finite networks. Those results are rigorously established in this paper.

Key words and phrases:
Mean field games, networks, graphons
2020 Mathematics Subject Classification:
49N80, 91A16, 91A43, 93E20
This work was supported by NSERC and AFOSR (P. E. Caines) and NSERC (M. Huang).

1. Introduction

One response to the problems arising in the analysis of systems of great complexity is to pass to an appropriately formulated infinite limit. This approach has a distinguished history since it is the conceptual principle underlying the celebrated Boltzmann Equation of statistical mechanics and that of the fundamental Navier-Stokes equation of fluid mechanics (see e.g. [38, 22, 14, 15]). Similarly the Fokker-Planck-Kolmogorov (FPK) equation for the macroscopic flow of probabilities [12, 27] is used to describe a vast range of phenomena which at a micro or mezzo level are modelled via the random interactions of discrete entities.

The work in this paper is formulated within two recent theories which were developed with an analogous motive to that above, namely Mean Field Game (MFG) theory for the analysis of equilibria in very large populations of non-cooperative agents (see [25, 23, 30, 31, 9, 10, 8]), and the graphon theory of the infinite limits of graphs and networks (see [33, 2, 3, 4, 32]).

A mathematically rigorous study of MFG systems with state values in finite graphs is provided in [21], and MFG systems where the agent subsystems are defined at the nodes (vertices) of finite random Erdös-Rényi graphs are treated in [11]. The system behaviour in [21] is subject to a fixed underlying network. The random graphs in [11] have unbounded growth but do not create spatial distinction of the agents due to symmetry properties of the interactions. However, graphon theory gives a rigorous formulation of the notion of limits for infinite sequences of networks of increasing size, and the first application of graphon theory in dynamics appears to be in the work of Medvedev [34, 35], and Kaliuzhnyi-Verbovetskyi and Medvedev [26]. The law of large numbers for graphon mean field systems is proven in [1] as a generalization of results for standard interacting particle systems. Furthermore, the work in [39] derives the McKean-Vlasov limit for a network of agents described by delay stochastic differential equations that are coupled by randomly generated connections.

The first applications of graphon theory in systems and control theory are those in [17, 18, 16, 19, 20] which treat the centralized and distributed control of arbitrarily large networks of linear dynamical control systems for which a direct solution would be intractable. Approximate control is achieved by solving control problems on the infinite limit graphon and then applying control laws derived from those solutions on the finite network of interest. The analogy with the strategies for finding feedback laws resulting in ϵ\epsilon-Nash equilibria in the MFG framework is obvious. In this connection we note that work on static game theoretic equilibria for infinite populations on graphons was reported in [37].

A natural framework for the formulation of game theoretic problems involving large populations of agents distributed over large networks is given by Mean Field Game theory defined on graphons. The resulting basic idea and the associated fundamental equations for what we term Graphon Mean Field Game (GMFG) systems and the GMFG equations are the subject of the current paper and its predecessors [6, 7]. The GMFG equations are of significant generality since they permit the study, in the limit, of both dense and sparse, infinite networks of non-cooperative dynamical agents. Moreover the classical MFG equations are retrieved as a special case. We observe that an early analysis of linear quadratic (LQ) models in mean field games on networks with non-uniform edge weightings can be found in [24]. However, in that work there was no application of graphon theory, and in the uniform system parameter case there is one agent per node and a single mean field, whereas in the present work there is a subpopulation with its own mean field at each node.

The basic ϵ\epsilon-Nash equilibrium result in MFG theory and its corresponding form in GMFG theory are vital for the application of MFG derived control laws. This is the case since the solution of the MFG and GMFG equations is necessarily simpler than the effectively intractable task of finding the solution to the game problems for the large finite population systems. Indeed, this was one of the original motives for the creation of MFG theory and it is a basic feature of graphon systems control theory [17].

The paper is organized as follows. Section 2 provides preliminary materials on graphons. Section 3 introduces the GMFG equation system and proves the existence and uniqueness of a solution. For the decentralized strategies determined by the GMFG equations, an ϵ\epsilon-Nash equilibrium theorem is proven in Section 4. The GMFG equations are illustrated by an LQ example in Section 5.

Table 1. Notation
GkG_{k} the kk-th graph in a sequence of graphs
gkg^{k} weights of GkG_{k} as a step function
MkM_{k} the number of nodes in GkG_{k}
𝒞i{\mathcal{C}}_{i} the cluster of agents residing at node ii of GkG_{k}
𝒞(i){\mathcal{C}}(i) the cluster that agent ii belongs to
IiI_{i}^{*}, I(i)I^{*}(i) the midpoint of an interval of length 1/Mk1/M_{k}
gg the graphon function
μα(t)\mu_{\alpha}(t) the local mean field generated by agents at vertex α[0,1]\alpha\in[0,1]
μG(t)\mu_{G}(t) an ensemble of local mean fields (μα(t))0α1(\mu_{\alpha}(t))_{0\leq\alpha\leq 1}
[0,T]{\mathcal{M}}_{[0,T]} a class of μG()\mu_{G}(\cdot) satisfying a Hölder continuity condition
CTC_{T} the space of continuous functions on [0,T][0,T]
T{\mathcal{F}}_{T} σ\sigma-algebra induced by cylindrical sets in CTC_{T}
(CT,T,mα)(C_{T},{\mathcal{F}}_{T},m_{\alpha}) probability measure space for the path space at vertex α\alpha
𝐌T{\bf M}_{T} the set of probability measures on (CT,T)(C_{T},{\mathcal{F}}_{T})
DTD_{T} Wasserstein metric on 𝐌T{\bf M}_{T}
𝐌TG{\bf M}_{T}^{G} the product space α[0,1]𝐌T\prod_{\alpha\in[0,1]}{\bf M}_{T}
𝐌TG0{\bf M}_{T}^{G0}, 𝐌TG1{\bf M}_{T}^{G1} subsets of 𝐌TG{\bf M}_{T}^{G}
mGm_{G} an ensemble of measures (mα)0α1𝐌TG(m_{\alpha})_{0\leq\alpha\leq 1}\in{\bf M}_{T}^{G}
Projα(mG){\rm Proj}_{\alpha}(m_{G}) the component mαm_{\alpha} at vertex α\alpha
Margt(mα){\rm Marg}_{t}(m_{\alpha}) the time tt-marginal of mαm_{\alpha}
xαx_{\alpha} the state of a generic agent at vertex α[0,1]\alpha\in[0,1]
wαw_{\alpha} a generic standard Brownian motion at vertex α\alpha
φ(t,xα|μG();gα)\varphi(t,x_{\alpha}|\mu_{G}(\cdot);g_{\alpha}) the best response at vertex α\alpha with μG()\mu_{G}(\cdot) given by the GMFG system;
abbreviated as φ(t,xα,gα)\varphi(t,x_{\alpha},g_{\alpha}) or φα\varphi_{\alpha}
ϕ(t,xα|μG();gα)\phi(t,x_{\alpha}|\mu_{G}(\cdot);g_{\alpha}) the best response at vertex α\alpha with respect to an arbitrary μG()\mu_{G}(\cdot);
abbreviated as ϕα(t,xα|μG())\phi_{\alpha}(t,x_{\alpha}|\mu_{G}(\cdot)) or ϕα\phi_{\alpha}

2. The Concept of a Graphon

The basic idea of the theory of graphons is that the edge structure of each finite cardinality network is represented by a step function density on the unit square in 2\mathbb{R}^{2} on which the so-called cut norm and cut metrics are defined. The set of finite graphs endowed with the cut metric then gives rise to a metric space, and the completion of this space is the space of graphons. Let 𝐆𝟎𝐬𝐩\mathbf{G_{0}^{sp}} denote the linear space of bounded symmetric Lebesgue measurable functions W:[0,1]2W:[0,1]^{2}\rightarrow\mathbb{R}, which are called kernels. The space 𝐆𝐬𝐩\mathbf{G^{sp}} of graphons is a subset of 𝐆𝟎𝐬𝐩\mathbf{G_{0}^{sp}} and consists of kernels W:[0,1]2[0,1]W:[0,1]^{2}\rightarrow[0,1] which can be interpreted as weighted graphs on the vertex set [0,1][0,1]. We note that functions W𝐆𝐬𝐩W\in\mathbf{G^{sp}} taking values in finite sets satisfy this definition and so, in particular, graphons are defined on finite graphs.

The cut norm of a kernel W𝐆𝟎𝐬𝐩W\in\mathbf{G_{0}^{sp}} then has the expression:

W=supM,T[0,1]|M×TW(x,y)𝑑x𝑑y|\|W\|_{\Box}=\sup_{M,T\subset[0,1]}\Big{|}\int_{M\times T}W(x,y)dxdy\Big{|}

with the supremum taking over all measurable subsets MM and TT of [0,1][0,1]. Denote the set of measure preserving bijections [0,1][0,1][0,1]\rightarrow[0,1] by S[0,1]S_{[0,1]}. The cut metric between two graphons VV and WW is then given by δ(W,V)=infϕS[0,1]WϕV\delta_{\Box}(W,V)=\inf_{\phi\in S_{[0,1]}}\|W^{\phi}-V\|_{\Box}, where Wϕ(x,y)W(ϕ(x),ϕ(y))W^{\phi}(x,y)\coloneqq W(\phi(x),\phi(y)) and any pair of graphons at zero distance are identified with each other. The space (𝐆𝐬𝐩,δ)(\mathbf{G^{sp}},\delta_{\Box}) is compact in the topology given by the cut metric [32]. Furthermore, sets in (𝐆𝐬𝐩,δ)(\mathbf{G^{sp}},\delta_{\Box}) which are compact with respect to the L2L^{2} metric are compact with respect to the cut metric. Since 𝐆𝐬𝐩\mathbf{G^{sp}} is compact in the cut metric all sequences of graphons have subsequential limits.

In this paper, we start with the modeling of the game of a finite population based on a finite graph. Specifically, the population resides on a weighted finite graph GkG_{k} with a set of nodes (or vertices) 𝒱k={1,,Mk}{\mathcal{V}}_{k}=\{1,\ldots,M_{k}\} and weights gijk[0,1]g^{k}_{ij}\in[0,1] for (i,j)𝒱k×𝒱k(i,j)\in{\mathcal{V}}_{k}\times{\mathcal{V}}_{k}, where a value giikg_{ii}^{k} is assigned in the case i=ji=j. We call gik(gi1k,,giMkk)g_{i}^{k}\coloneqq(g_{i1}^{k},\ldots,g_{iM_{k}}^{k}) a section of gkg^{k} at ii. Each node ll is occupied by a set of agents which is called a cluster of the population and hence the number of clusters is MkM_{k}. We list the clusters as 𝒞1,,𝒞Mk\mathcal{C}_{1},\ldots,\mathcal{C}_{M_{k}}. Without loss of generality, we assume the llth cluster occupies node ll. Let 𝒞(i)\mathcal{C}(i) denote the cluster that agent ii belongs to. So i𝒞(i)i\in\mathcal{C}(i). Our further analysis in the paper is based on the convergence of gkg^{k} to a graphon limit gg. We may naturally identify (gijk)1i,jMk(g^{k}_{ij})_{1\leq i,j\leq M_{k}} with a graphon gk(α,β)g^{k}(\alpha,\beta) as a step function defined on [0,1]×[0,1][0,1]\times[0,1] (see [32]). However, convergence in the cut norm or the cut metric is inadequate for the analysis in this paper as it does not capture sufficiently strong sectional information of the difference gkgg^{k}-g. We will adopt a different convergence notion strengthening the sectional requirement as in assumption (H11) below. To indicate its arguments, we may write g(α,β)g(\alpha,\beta) or alternatively gα,βg_{\alpha,\beta}. We define the section of gg at α\alpha by gα:βgα,βg_{\alpha}:\beta\mapsto g_{\alpha,\beta}, β[0,1]\beta\in[0,1].

Since clusters 𝒞i1\mathcal{C}_{i_{1}} and 𝒞i2\mathcal{C}_{i_{2}} reside on nodes i1i_{1} and i2i_{2} of GkG_{k}, respectively, we define g𝒞i1𝒞i2k=gi1i2kg^{k}_{\mathcal{C}_{i_{1}}\mathcal{C}_{i_{2}}}=g^{k}_{i_{1}i_{2}}. Similarly, we define the section g𝒞ik=gikg^{k}_{\mathcal{C}_{i}}=g^{k}_{i}.

We partition [0,1][0,1] into MkM_{k} subintervals of equal length. Here Ilk=[(l1)/Mk,l/Mk]I_{l}^{k}=[(l-1)/M_{k},l/M_{k}] for 1lMk1\leq l\leq M_{k}. When it is clear from the context, we omit the superscript kk and write IlI_{l}. To relate the clusters of agents to the vertex set [0,1][0,1], we let the cluster 𝒞l\mathcal{C}_{l} correspond to IlI_{l}.

Throughout this paper, C,C0,C1,C,C_{0},C_{1},\ldots denote generic constants, which do not depend on the graph index kk and population size NN and may vary from place to place.

3. Graphon MFG Systems and the GMFG Equations

3.1. The Standard MFG Model and Its Graphon Generalization

In the diffusion based models of large population games the state evolution of a collection of NN agents 𝒜i,1iN<,{\mathcal{A}}_{i},1\leq i\leq N<\infty, is specified by a set of NN controlled stochastic differential equations (SDEs). A simplified form of the general case is given by the following set of controlled SDEs which for each agent 𝒜i{\mathcal{A}}_{i} includes state coupling with all other agents:

(3.1) dxi(t)=1Nj=1Nf(xi(t),ui(t),xj(t))dt+σdwi(t),dx_{i}(t)=\frac{1}{N}\sum_{j=1}^{N}f(x_{i}(t),u_{i}(t),x_{j}(t))dt+\sigma dw_{i}(t),\\

where xinx_{i}\in\mathbb{R}^{n} is the state, uinuu_{i}\in\mathbb{R}^{n_{u}} the control input, and winww_{i}\in{\mathbb{R}}^{n_{w}} a standard Brownian motion, and where {wi,1iN}\{w_{i},1\leq i\leq N\} are independent processes. For simplicity, all collections of system initial conditions are taken to be independent and have finite second moment. The cost of agent 𝒜i{\mathcal{A}}_{i} is given by

(3.2) JiN(ui,ui)=E0T1Nj=1Nl(xi(t),ui(t),xj(t))dt,J_{i}^{N}(u_{i},u_{-i})=E\int_{0}^{T}\frac{1}{N}\sum_{j=1}^{N}l(x_{i}(t),u_{i}(t),x_{j}(t))dt,

where l()l(\cdot) is the pairwise running cost, and uiu_{-i} denotes the controls of all other agents.

The dynamics of a generic agent 𝒜i{\mathcal{A}}_{i} in the infinite population limit of this system is then described by the controlled McKean-Vlasov (MV) equation

(3.3) dxi=f[xi,ui,μt]dt+σdwi,0tT,dx_{i}=f[x_{i},u_{i},\mu_{t}]dt+\sigma dw_{i},\quad 0\leq t\leq T,

where μt\mu_{t} is the distribution of xi(t)x_{i}(t), f[x,u,μt]nf(x,u,y)μt(dy)f[x,u,\mu_{t}]\coloneqq\int_{\mathbb{R}^{n}}f(x,u,y)\mu_{t}(dy) and where the initial distribution μ0x\mu^{x}_{0} of xi(0)x_{i}(0) is specified. Setting l[x,u,μt]=nl(x,u,y)μt(dy)l[x,u,\mu_{t}]=\int_{\mathbb{R}^{n}}l(x,u,y)\mu_{t}(dy), the corresponding infinite population cost for 𝒜i{\mathcal{A}}_{i} takes the form

(3.4) Ji(ui;μ())E0Tl[xi(t),ui(t),μt]𝑑t.J_{i}(u_{i};\mu(\cdot))\coloneqq E\int_{0}^{T}l[x_{i}(t),u_{i}(t),\mu_{t}]dt.

For notational simplicity, we present the graphon MFG framework with scalar individual states and controls, i.e., n=nu=nw=1n=n_{u}=n_{w}=1. Its extension to the vector case is evident.

Now we consider a finite population distributed over the finite graph GkG_{k}. Let \mathboldxGk=l=1Mk{xi|i𝒞l}\mathbold{x}_{G_{k}}=\bigoplus_{l=1}^{M_{k}}\{x_{i}|i\in\mathcal{C}_{l}\} denote the states of all agents in the total set of clusters of the population. This gives a total of N=l=1Mk|𝒞l|N=\sum_{l=1}^{M_{k}}|\mathcal{C}_{l}| individual states. The key feature of the graphon MFG construction beyond the standard MFG scheme is that at any agent in a network the averaged dynamics (3.1) and cost function (3.2) decompose into averages of subpopulations distributed at that agent’s neighboring nodes plus an average term for the local cluster. In the limit, the summed subpopulation averages are given by an integral over the local mean fields of the neighbouring agents.

For 𝒜i{\mathcal{A}}_{i} in the cluster 𝒞(i)\mathcal{C}(i), two coupling terms in the dynamics take the form

(3.5) f0(xi,ui,𝒞(i))=1|𝒞(i)|j𝒞(i)f0(xi,ui,xj),\displaystyle f_{0}(x_{i},u_{i},{\mathcal{C}}(i))=\frac{1}{|{\mathcal{C}}(i)|}\sum_{j\in{\mathcal{C}}(i)}f_{0}(x_{i},u_{i},x_{j}),
(3.6) fGk(xi,ui,g𝒞(i)k)=1Mkl=1Mkg𝒞(i)𝒞lk1|𝒞l|j𝒞lf(xi,ui,xj).\displaystyle f_{G_{k}}(x_{i},u_{i},g^{k}_{\mathcal{C}(i)})=\frac{1}{M_{k}}\sum_{l=1}^{M_{k}}g^{k}_{\mathcal{C}(i)\mathcal{C}_{l}}\frac{1}{|\mathcal{C}_{l}|}\sum_{j\in\mathcal{C}_{l}}f(x_{i},u_{i},x_{j}).

They model intra- and inter-cluster couplings, respectively. The specification of fGkf_{G_{k}} relies on the sectional information g𝒞(i)kg^{k}_{\mathcal{C}(i)\bullet}. Concerning the coupling structure in (3.6) we observe that with respect to 𝒜i{\mathcal{A}}_{i}, all individuals residing in cluster 𝒞l\mathcal{C}_{l} are symmetric and their state average generates the overall impact of that cluster on 𝒜i{\mathcal{A}}_{i} mediated by the graphon weighting g𝒞(i)kg^{k}_{\mathcal{C}(i)\bullet}. The two coupling terms are combined additively resulting in the local dynamics

f~Gk(xi,ui,g𝒞(i)k)=f0(xi,ui,𝒞(i))+fGk(xi,ui,g𝒞(i)k).\tilde{f}_{G_{k}}(x_{i},u_{i},g^{k}_{\mathcal{C}(i)})=f_{0}(x_{i},u_{i},{\mathcal{C}}(i))+f_{G_{k}}(x_{i},u_{i},g^{k}_{\mathcal{C}(i)}).

Note that 𝒜i{\mathcal{A}}_{i} interacts with the overall population through a function of the complete system state \mathboldxGk{\mathbold x}_{G_{k}} and the cluster sizes. These details shall be suppressed in this paper and we only indicate the graph GkG_{k} and the section g𝒞(i)kg^{k}_{\mathcal{C}(i)}. The state process of 𝒜i{\mathcal{A}}_{i} is then given by the stochastic differential equation

dxi(t)=f~Gk(xi,ui,g𝒞(i)k)dt+σdwi,1iN,\displaystyle dx_{i}(t)=\tilde{f}_{G_{k}}(x_{i},u_{i},g^{k}_{\mathcal{C}(i)})dt+\sigma dw_{i},\quad 1\leq i\leq N,

where σ>0\sigma>0 and the initial states {xi(0),1iN}\{x_{i}(0),1\leq i\leq N\} are i.i.d. with distribution μ0x𝒫1()\mu^{x}_{0}\in{\mathcal{P}}_{1}(\mathbb{R}), the set of probability measures on \mathbb{R} with finite mean.

The limit of the two dynamic coupling terms of an agent at a node α\alpha (called an α\alpha-agent), as the number of nodes of the graph GkG_{k} and the subpopulation at each node tend to infinity, is described by the following expressions:

(3.7) f0[xα,uα,μα]f0(xα,uα,z)μα(dz),\displaystyle{f}_{0}[x_{\alpha},u_{\alpha},{\mu}_{\alpha}]\coloneqq\int_{{\mathbb{R}}}f_{0}(x_{\alpha},u_{\alpha},z)\mu_{\alpha}(dz),
(3.8) f[xα,uα,μG;gα]01f(xα,uα,z)g(α,β)μβ(dz)𝑑β,\displaystyle{f}[x_{\alpha},u_{\alpha},{\mu}_{G};g_{\alpha}]\coloneqq\int_{0}^{1}\int_{{\mathbb{R}}}f(x_{\alpha},u_{\alpha},z)g(\alpha,\beta)\mu_{\beta}(dz)d\beta,

which give the complete local graphon dynamics via

(3.9) f~[xα,uα,μG;gα]f0[xα,uα,μα]+f[xα,uα,μG;gα].\widetilde{f}[x_{\alpha},u_{\alpha},{\mu}_{G};g_{\alpha}]\coloneqq f_{0}[x_{\alpha},u_{\alpha},\mu_{\alpha}]+{f}[x_{\alpha},u_{\alpha},{\mu}_{G};g_{\alpha}].

We call μβ\mu_{\beta} the local mean field at node β\beta, which is interpreted as the limit of the empirical distributions of agents at node β\beta. And μG={μβ,0β1}\mu_{G}=\{\mu_{\beta},0\leq\beta\leq 1\} is the ensemble of local mean fields. Due to the integration with respect to β\beta, the dependence of f~\widetilde{f} on the graphon limit gg is through the section gαg_{\alpha}. Since μG\mu_{G} contains μα\mu_{\alpha}, we do not list μα\mu_{\alpha} as an argument of f~\widetilde{f}.

Parallel to the standard MFG case, in the graphon case the stochastic differential equation

(3.10) [MV-SDE](α)dxα(t)=f~[xα(t),uα(t),μG(t);gα]dt+σdwα(t),\displaystyle{\text{[MV-SDE]}}(\alpha)\quad dx_{\alpha}(t)=\widetilde{f}[x_{\alpha}(t),u_{\alpha}(t),{\mu}_{G}(t);g_{\alpha}]dt+\sigma dw_{\alpha}(t),
0tT,α[0,1],\displaystyle\quad 0\leq t\leq T,\quad\alpha\in[0,1],

generalizes the standard controlled MV equation (3.3). We note that in a parallel development of graphon based stochastic dynamical populations [1] the system disturbance intensity σ\sigma is also a function of graphon weighted state functions at other clusters. For simplicity, we consider a constant σ\sigma and our analysis may be generalized to the case of a state and mean field dependent diffusion term. Similarly, for simplicity our dynamics and cost do not include a separate parametrization by α\alpha.

Analogously, in the GMFG case, we define the cost coupling terms for 𝒜i{\mathcal{A}}_{i} to be

l0(xi,ui,𝒞(i))=1|𝒞(i)|j𝒞(i)l0(xi,ui,xj),\displaystyle l_{0}(x_{i},u_{i},\mathcal{C}(i))=\frac{1}{|\mathcal{C}(i)|}\sum_{j\in\mathcal{C}(i)}l_{0}(x_{i},u_{i},x_{j}),
lGk(xi,ui,g𝒞(i)k)=1Mkl=1Mkg𝒞(i)𝒞lk1|𝒞l|j𝒞ll(xi,ui,xj).\displaystyle l_{G_{k}}(x_{i},u_{i},g^{k}_{\mathcal{C}(i)})=\frac{1}{M_{k}}\sum_{l=1}^{M_{k}}g^{k}_{\mathcal{C}(i)\mathcal{C}_{l}}\frac{1}{|\mathcal{C}_{l}|}\sum_{j\in\mathcal{C}_{l}}l(x_{i},u_{i},x_{j}).

Define l~Gk(xi,ui,g𝒞(i)k)=l0(xi,ui,𝒞(i))+lGk(xi,ui,g𝒞(i)k).\tilde{l}_{G_{k}}(x_{i},u_{i},g^{k}_{\mathcal{C}(i)})=l_{0}(x_{i},u_{i},\mathcal{C}(i))+l_{G_{k}}(x_{i},u_{i},g^{k}_{\mathcal{C}(i)}). The cost of 𝒜i{\mathcal{A}}_{i} in a finite population on a finite graph GkG_{k} is given in the form

(3.11) Ji=E0Tl~Gk(xi,ui,g𝒞(i)k)𝑑t.\displaystyle J_{i}=E\int_{0}^{T}\tilde{l}_{G_{k}}(x_{i},u_{i},g^{k}_{\mathcal{C}(i)})dt.

Denote

l0[xα,uα,μα]=l0(xα,uα,z)μα(dz),\displaystyle l_{0}[x_{\alpha},u_{\alpha},\mu_{\alpha}]=\int_{\mathbb{R}}l_{0}(x_{\alpha},u_{\alpha},z)\mu_{\alpha}(dz),
l[xα,uα,μG;gα]=01l(xα,uα,z)g(α,β)μβ(dz)𝑑β,\displaystyle l[x_{\alpha},u_{\alpha},\mu_{G};g_{\alpha}]=\int_{0}^{1}\int_{\mathbb{R}}l(x_{\alpha},u_{\alpha},z)g(\alpha,\beta)\mu_{\beta}(dz)d\beta,
l~[xα,uα,μG;gα]=l0[xα,uα,μα]+l[xα,uα,μG;gα].\displaystyle\widetilde{l}[x_{\alpha},u_{\alpha},\mu_{G};g_{\alpha}]=l_{0}[x_{\alpha},u_{\alpha},\mu_{\alpha}]+l[x_{\alpha},u_{\alpha},\mu_{G};g_{\alpha}].

Then in the infinite population graphon case, the α\alpha-agent has the cost function given by

(3.12) Jα(uα;μG())=E0Tl~[xα(t),uα(t),μG(t);gα]𝑑t.\displaystyle J_{\alpha}(u_{\alpha};\mu_{G}(\cdot))=E\int_{0}^{T}\widetilde{l}[x_{\alpha}(t),u_{\alpha}(t),\mu_{G}(t);g_{\alpha}]dt.

3.2. The Graphon MFG Model and Its Equations

In this section the standard MFG equations (see e.g. [5, 8]) will be generalized so that they subsume the standard (implicitly uniform totally connected) dense network case and cover the fully general graphon limit network case. Specifically, agent 𝒜i{\mathcal{A}}_{i} in a population of NN agents will be located at the llth node in an MkM_{k} node network (identified with its graphon) and in the infinite population graphon limit that node will be taken to map to α[0,1]\alpha\in[0,1]. It is important to note here that although the limit network is assumed dense it is not assumed to be uniformly totally connected; indeed, the connection structure of the infinite network is represented precisely by its graphon g(α,β)g(\alpha,\beta), 0α,β1.0\leq\alpha,\beta\leq 1.

The generalized Graphon MFG scheme below on [0,T][0,T] is given for each α\alpha by (i) the Hamilton-Jacobi-Bellman (HJB) equation generating the value function VαV^{\alpha} when all other agents’ control laws and the ensemble μG\mu_{G} of local mean fields are given, (ii) the FPK equation generating the local mean field μα\mu_{\alpha} given μG\mu_{G}, and (iii) the specification of the best response (BR) feedback law.

Suppressing the time index on the measures for simplicity of notation, we have the Graphon Mean Field Game (GMFG) equations:

[HJB](α)Vα(t,x)t=infuU{f~[x,u,μG;gα]Vα(t,x)x\displaystyle{\text{[HJB]}}(\alpha)\quad-\frac{{\partial}V^{\alpha}(t,x)}{{\partial}t}=\inf_{u\in U}\bigg{\{}\widetilde{f}[x,u,{\mu}_{G};g_{\alpha}]\frac{{\partial}V^{\alpha}(t,x)}{{\partial}x}
(3.13) +l~[x,u,μG;gα]}+σ222Vα(t,x)x2,\displaystyle\hskip 113.81102pt+\widetilde{l}[x,u,{\mu}_{G};g_{\alpha}]\bigg{\}}+\frac{\sigma^{2}}{2}\frac{{\partial}^{2}V^{\alpha}(t,x)}{{\partial}x^{2}},
Vα(T,x)=0,(t,x)[0,T]×,α[0,1],\displaystyle V^{\alpha}(T,x)=0,\quad(t,x)\in[0,T]\times\mathbb{R},\quad\alpha\in[0,1],
[FPK](α)pα(t,x)t=\displaystyle{\text{[FPK]}}(\alpha)\quad\frac{{\partial}p_{\alpha}(t,x)}{{\partial}t}= {f~[x,u0,μG;gα]pα(t,x)}x\displaystyle-\frac{{\partial}\{\widetilde{f}[x,u^{0},\mu_{G};g_{\alpha}]p_{\alpha}(t,x)\}}{{\partial}x}
(3.14) +σ222pα(t,x)x2,\displaystyle+\frac{\sigma^{2}}{2}\frac{{\partial}^{2}p_{\alpha}(t,x)}{{\partial}x^{2}},
[BR](α)u0\displaystyle{\text{[BR]}}(\alpha)\quad u^{0} φ(t,x|μG;gα).\displaystyle\coloneqq\varphi(t,x|{\mu}_{G};g_{\alpha}).

Here pα(t,x)p_{\alpha}(t,x) with initial condition pα(0)p_{\alpha}(0) is used to denote the density of the measure μα(t)\mu_{\alpha}(t) whenever a density is assumed to exist. The FPK equation may be replaced by the following closed-loop MV-SDE:

(3.15) [MV](α)dxα(t)=f~[xα(t),φ(t,xα(t)|μG;gα),μG(t);gα]dt+σdwα(t),\displaystyle\text{[MV]}(\alpha)\quad dx_{\alpha}(t)=\widetilde{f}[x_{\alpha}(t),\varphi(t,x_{\alpha}(t)|{\mu}_{G};g_{\alpha}),{\mu}_{G}(t);g_{\alpha}]dt+\sigma dw_{\alpha}(t),

where xα(0)x_{\alpha}(0) has distribution μ0x\mu_{0}^{x}. Our subsequent analysis will directly treat the pair (Vα(t,x),μα(t))(V^{\alpha}(t,x),\mu_{\alpha}(t)), where μα(t)\mu_{\alpha}(t) is specified as the law of xα(t)x_{\alpha}(t) in (3.15).

When a solution exists for the GMFG equations, the resulting BR feedback controls depend upon the ensemble μG\mu_{G} of local mean fields and the individual agent’s state. This is a natural generalization of the standard case. The standard MFG case is simply obtained by setting g(α,β)0,0α,β1g(\alpha,\beta)\equiv 0,0\leq\alpha,\beta\leq 1, which totally disconnects the network and results in f~[x,u,μG;gα]=f0[x,u,μ]\widetilde{f}[x,u,{\mu}_{G};g_{\alpha}]={f}_{0}[x,u,\mu] and l~[x,u,μG;gα]=l0[x,u,μ]\widetilde{l}[x,u,{\mu}_{G};g_{\alpha}]=l_{0}[x,u,\mu] [5, 8].

A collection of measures on some measurable space which are indexed by the vertex set [0,1][0,1] is called a measure ensemble. Thus, for each fixed tt, μG(t)\mu_{G}(t) is a measure ensemble.

On 𝒫1(){\mathcal{P}}_{1}({\mathbb{R}}) we endow the Wasserstein metric W1W_{1}: for any μ,ν𝒫1()\mu,\nu\in{\mathcal{P}}_{1}({\mathbb{R}}), W1(μ,ν)=infγ^|xy|γ^(dx,dy)W_{1}(\mu,\nu)=\inf_{\widehat{\gamma}}\int|x-y|\widehat{\gamma}(dx,dy), where γ^\widehat{\gamma} is a probability measure on 2\mathbb{R}^{2} with marginals μ,ν\mu,\nu.

Let C([0,1],𝒫1())C([0,1],{\mathcal{P}}_{1}({\mathbb{R}})) be the set of measure ensembles νG=(νβ)β[0,1]\nu_{G}=(\nu_{\beta})_{\beta\in[0,1]} satisfying νβ𝒫1()\nu_{\beta}\in{\mathcal{P}}_{1}({\mathbb{R}}), and limββW1(νβ,νβ)=0\lim_{\beta^{\prime}\to\beta}W_{1}(\nu_{\beta^{\prime}},\nu_{\beta})=0 for any β[0,1]\beta\in[0,1].

In order to analyze the solvability of the GMFG equations, we need to restrict μG()\mu_{G}(\cdot) to a certain class. We say {μG(t),0tT}\{\mu_{G}(t),0\leq t\leq T\} is from the admissible set [0,T]{\mathcal{M}}_{[0,T]} if:

(C1) For each fixed tt, μG(t)\mu_{G}(t) is in C([0,1],𝒫1())C([0,1],{\mathcal{P}}_{1}({\mathbb{R}})).

(C2) There exists η(0,1]\eta\in(0,1] such that for any bounded and Lipschitz continuous function ϕ\phi on \mathbb{R},

supβ[0,1]|ϕ(y)μβ(t1,dy)ϕ(y)μβ(t2,dy)|Ch|t1t2|η,\sup_{\beta\in[0,1]}\Big{|}\int_{\mathbb{R}}\phi(y)\mu_{\beta}(t_{1},dy)-\int_{\mathbb{R}}\phi(y)\mu_{\beta}(t_{2},dy)\Big{|}\leq C_{h}|t_{1}-t_{2}|^{\eta},

where ChC_{h} may be selected to depend only on the Lipschitz constant  Lip(ϕ)\mbox{ Lip}(\phi) for ϕ\phi.

Condition (C1) ensures that integration with respect to dβd\beta in (3.8) is well defined. Condition (C2) ensures that the drift term in the HJB equation (3.2) has a certain time continuity, which facilitates the subsequent existence analysis of the best response.

3.3. Existence Analysis

We introduce the following assumptions:

(H1) UU is a compact set.

(H2) f0(x,u,y)f_{0}(x,u,y), f(x,u,y)f(x,u,y), l0(x,u,y)l_{0}(x,u,y) and l(x,u,y)l(x,u,y) are continuous and bounded functions on ×U×\mathbb{R}\times U\times\mathbb{R} and are Lipschitz continuous in (x,y)(x,y), uniformly with respect to uu.

(H3) f0(x,u,y)f_{0}(x,u,y) and f(x,u,y)f(x,u,y) are Lipschitz continuous in uu, uniformly with respect to (x,y)(x,y).

(H4) For any qq\in\mathbb{R}, α[0,1]\alpha\in[0,1] and probability measure ensemble νGC([0,1],𝒫1()){\nu}_{G}\in C([0,1],{\mathcal{P}}_{1}({\mathbb{R}})), the set

(3.16) SανG(x,q)\displaystyle S_{\alpha}^{\nu_{G}}(x,q) =argminuU{q(f~[x,u,νG;gα])+l~[x,u,νG;gα]}\displaystyle=\arg\min_{u\in U}\{q(\widetilde{f}[x,u,{\nu}_{G};g_{\alpha}])+\widetilde{l}[x,u,{\nu}_{G};g_{\alpha}]\}

is a singleton, and for any given compact interval =[q¯,q¯]{\mathcal{I}}=[\underline{q},\bar{q}], the resulting uu as a function of (x,q)×(x,q)\in\mathbb{R}\times{\mathcal{I}} is Lipschitz continuous in (x,q)(x,q), uniformly with respect to νG{\nu}_{G} and gαg_{\alpha}, 0α10\leq\alpha\leq 1.

The next two assumptions will be used to ensure that the best responses have continuous dependence on α\alpha. In particular, (H5) is a continuity assumption on the graphon function g(α,β)g(\alpha,\beta). Under (H5), f~\widetilde{f} and l~\widetilde{l} have continuity in α\alpha.

(H5) For any bounded and measurable function h(β)h(\beta), the function 01g(α,β)h(β)𝑑β\int_{0}^{1}g(\alpha,\beta)h(\beta)d\beta is continuous in α[0,1]\alpha\in[0,1].

(H6) For given νGC([0,1],𝒫1())\nu_{G}\in C([0,1],{\mathcal{P}}_{1}({\mathbb{R}})), SανG(x,q)S_{\alpha}^{\nu_{G}}(x,q) is continuous in (α,x,q)(\alpha,x,q).

Although the GMFG equation system only involves {μG(t),0tT}\{\mu_{G}(t),0\leq t\leq T\}, which may be viewed as a collection of marginals at different vertices, it is necessary to develop the existence analysis in the underlying probability spaces (see related discussions in [25, p.240]).

We begin by introducing some analytic preliminaries. For the space CT=C([0,T],)C_{T}=C([0,T],\mathbb{R}), we specify a σ\sigma-algebra T{\mathcal{F}}_{T} induced by all cylindrical sets of the form {x()CT:x(ti)Bi,1ij for some j}\{x(\cdot)\in C_{T}:x(t_{i})\in B_{i},1\leq i\leq j\mbox{ for some }j\}, where BiB_{i} is a Borel set. Let 𝐌T{\bf M}_{T} denote the space of all probability measures on (CT,T)(C_{T},{\mathcal{F}}_{T}). The canonical process XX is defined by Xt(ω)=ωtX_{t}(\omega)=\omega_{t} for ωCT\omega\in C_{T}. On CTC_{T}, we introduce the metric ρ(x,y)=supt|x(t)y(t)|1\rho(x,y)=\sup_{t}|x(t)-y(t)|\wedge 1. Then (CT,ρ)(C_{T},\rho) is a complete metric space. Based on ρ\rho, we introduce the Wasserstein metric on 𝐌T{\bf M}_{T}. For m1,m2𝐌Tm_{1},m_{2}\in{\bf M}_{T}, denote

DT(m1,m2)=infm^CT×CT(supsT|Xs(ω1)Xs(ω2)|1)𝑑m^(ω1,ω2),\displaystyle D_{T}(m_{1},m_{2})=\inf_{\widehat{m}}\int_{C_{T}\times C_{T}}\Big{(}\sup_{s\leq T}|X_{s}(\omega_{1})-X_{s}(\omega_{2})|\wedge 1\Big{)}d\widehat{m}(\omega_{1},\omega_{2}),

where m^\widehat{m} is called a coupling as a probability measure on (CT,T)×(CT,T)(C_{T},{\mathcal{F}}_{T})\times(C_{T},{\mathcal{F}}_{T}) with the pair of marginals m1m_{1} and m2m_{2}, respectively. Then (𝐌T,DT)({\bf M}_{T},D_{T}) is a complete metric space [41].

We introduce the product of probability measure spaces α[0,1](CT,T,mα)\prod_{\alpha\in[0,1]}(C_{T},{\mathcal{F}}_{T},m_{\alpha}), where each individual space is interpreted as the path space of the agent at vertex α\alpha with a corresponding probability measure mαm_{\alpha}. Denote the product of spaces of probability measures 𝐌TG=α[0,1]𝐌T.{\bf M}_{T}^{G}=\prod_{\alpha\in[0,1]}{\bf M}_{T}. An element in 𝐌TG{\bf M}_{T}^{G} is a measure ensemble. Given mG𝐌TGm_{G}\in{\bf M}_{T}^{G}, the projection operator Projα{\rm Proj}_{\alpha} picks out its component mαm_{\alpha} associated with α[0,1]\alpha\in[0,1]. Let 𝐌TG0{\bf M}_{T}^{G0} consist of all (mα)α[0,1]𝐌TG(m_{\alpha})_{\alpha\in[0,1]}\in{\bf M}_{T}^{G} such that for any α[0,1]\alpha\in[0,1], DT(mα,mα)0D_{T}(m_{\alpha^{\prime}},m_{\alpha})\to 0 as αα\alpha^{\prime}\to\alpha.

For two measure ensembles mG(mα)α[0,1]m_{G}\coloneqq(m_{\alpha})_{\alpha\in[0,1]} and m¯G(m¯α)α[0,1]\bar{m}_{G}\coloneqq(\bar{m}_{\alpha})_{\alpha\in[0,1]} in 𝐌TG{\bf M}_{T}^{G}, define d(mG,m¯G)=supα[0,1]DT(mα,m¯α).d(m_{G},\bar{m}_{G})=\sup_{\alpha\in[0,1]}D_{T}(m_{\alpha},\bar{m}_{\alpha}).

Lemma 3.1.

(𝐌TG,d)({\bf M}_{T}^{G},d) is a complete metric space.

Proof.

If {mGk,k1}\{m_{G}^{k},k\geq 1\} is a Cauchy sequence in 𝐌TG{\bf M}_{T}^{G}, then for each given α\alpha, the sequence {Projα(mGk),k1}\{{\rm Proj}_{\alpha}(m_{G}^{k}),k\geq 1\} (of probability measures) is a Cauchy sequence in the complete metric space 𝐌T{\bf M}_{T} and so it contains a limit. This in turn determines a limit in 𝐌TG{\bf M}_{T}^{G}. ∎

Given the probability measure mα𝐌Tm_{\alpha}\in{\bf M}_{T}, we determine the tt-marginal μα(t)\mu_{\alpha}(t) by μα(t,B)=mα({x()CT:x(t)B})\mu_{\alpha}(t,B)=m_{\alpha}(\{x(\cdot)\in C_{T}:x(t)\in B\}) for any Borel set BB\subset\mathbb{R}, and denote the mapping from 𝐌T{\bf M}_{T} to 𝒫(){\mathcal{P}}(\mathbb{R}) (the set of probability measures on {\mathbb{R}}):

(3.17) μα(t)=Margt(mα).\displaystyle\mu_{\alpha}(t)={\rm Marg}_{t}(m_{\alpha}).

Consider the measure ensemble mG=(mα)α[0,1]𝐌TGm_{G}=(m_{\alpha})_{\alpha\in[0,1]}\in{\bf M}_{T}^{G} with μα(t)\mu_{\alpha}(t) given by (3.17). Define the time tt marginals by the following mapping

(3.18) Margt(mG)=(μα(t))α[0,1],\displaystyle{\rm Marg}_{t}(m_{G})=(\mu_{\alpha}(t))_{\alpha\in[0,1]},

where the right hand side is simply written as μG(t)\mu_{G}(t). For a given tt, μG(t)\mu_{G}(t) may be interpreted as a measure valued function defined on the vertex set [0,1][0,1]. Further denote the mapping Marg(mG)=(μG(t))t[0,T]=μG(){\rm Marg}(m_{G})=(\mu_{G}(t))_{t\in[0,T]}=\mu_{G}(\cdot).

Take a fixed

(3.19) μG()[0,T]\displaystyle\mu_{G}(\cdot)\in{\mathcal{M}}_{[0,T]}

with its associated Hölder parameter η\eta in (C2), and denote

f~α(t,x,u)=f~[x,u,μG(t);gα],l~α(t,x,u)=l~[x,u,μG(t);gα].\displaystyle\widetilde{f}_{\alpha}^{*}(t,x,u)=\widetilde{f}[x,u,\mu_{G}(t);g_{\alpha}],\quad\widetilde{l}_{\alpha}^{*}(t,x,u)=\widetilde{l}[x,u,\mu_{G}(t);g_{\alpha}].
Lemma 3.2.

Assume (H1)–(H2). For hα=f~α(t,x,u)h_{\alpha}=\widetilde{f}_{\alpha}^{*}(t,x,u) or l~α(t,x,u)\widetilde{l}_{\alpha}^{*}(t,x,u), there exist constants CC and CμGC_{\mu_{G}}, where the latter depends on μG()\mu_{G}(\cdot), such that

supt,u,α|hα(t,x,u)hα(t,y,u)|C|xy|,\displaystyle\sup_{t,u,\alpha}|h_{\alpha}(t,x,u)-h_{\alpha}(t,y,u)|\leq C|x-y|,
supx,u,α|hα(t,x,u)hα(s,x,u)|CμG|ts|η,\displaystyle\sup_{x,u,\alpha}|h_{\alpha}(t,x,u)-h_{\alpha}(s,x,u)|\leq C_{\mu_{G}}|t-s|^{\eta},

where the supremum is taken over t[0,T]t\in[0,T], xx\in\mathbb{R}, uUu\in U and α[0,1]\alpha\in[0,1].

Proof.

The Lipschitz continuity of f~α\widetilde{f}^{*}_{\alpha} with respect to xx follows from (H2) and (3.7)–(3.8). For t1,t2[0,T]t_{1},t_{2}\in[0,T], we estimate |f~[x,u,μG(t1);gα]f~[x,u,μG(t2);gα]||\widetilde{f}[x,u,\mu_{G}(t_{1});g_{\alpha}]-\widetilde{f}[x,u,\mu_{G}(t_{2});g_{\alpha}]| by using the Lipschitz condition of f0f_{0}, ff and condition (C2) for [0,T]{\mathcal{M}}_{[0,T]}. This establishes the Hölder continuity of f~α\widetilde{f}_{\alpha}^{*} in tt. The other cases can be similarly checked. ∎

In order to analyze the best response of the α\alpha-agent, we introduce the HJB equation

(3.20) Vtα(t,x)=infuU{f~α(t,x,u)Vxα(t,x)+l~α(t,x,u)}+σ22Vxxα(t,x),\displaystyle-V^{\alpha}_{t}(t,x)=\inf_{u\in U}\{\widetilde{f}_{\alpha}^{*}(t,x,u)V_{x}^{\alpha}(t,x)+\widetilde{l}_{\alpha}^{*}(t,x,u)\}+\frac{\sigma^{2}}{2}V_{xx}^{\alpha}(t,x),

where Vα(T,0)=0V^{\alpha}(T,0)=0. It differs from (3.2) by allowing an arbitrary μG()[0,T]\mu_{G}(\cdot)\in{\mathcal{M}}_{[0,T]}.

For studying (3.20), we introduce some standard definitions. Denote QT=(0,T)×Q_{T}=(0,T)\times\mathbb{R}, and Q¯T=[0,T]×\overline{Q}_{T}=[0,T]\times\mathbb{R}. Let C1,2(Q¯T)C^{1,2}(\overline{Q}_{T}) (resp., C1,2(QT)C^{1,2}(Q_{T})) denote the set of functions with continuous derivatives vt,vx,vxxv_{t},v_{x},v_{xx} on Q¯T\overline{Q}_{T} (resp., QTQ_{T}). Let Cb1,2(Q¯T)C_{b}^{1,2}(\overline{Q}_{T}) be the set of bounded functions in C1,2(Q¯T)C^{1,2}(\overline{Q}_{T}), and let the open (or closed) set QbQ_{b} be a bounded subset of QTQ_{T}. Wλ1,2(Qb)W^{1,2}_{\lambda}(Q_{b}), 1λ<1\leq\lambda<\infty, shall denote the Sobolev space consisting of functions vv such that each vv and its generalized derivatives vtv_{t}, vxv_{x}, vxxv_{xx} are in Lλ(Qb)L^{\lambda}(Q_{b}); further we have the norm

(3.21) vλ,Qb(2)=vλ,Qb+vtλ,Qb+vxλ,Qb+vxxλ,Qb,\displaystyle\|v\|_{\lambda,Q_{b}}^{(2)}=\|v\|_{\lambda,Q_{b}}+\|v_{t}\|_{\lambda,Q_{b}}+\|v_{x}\|_{\lambda,Q_{b}}+\|v_{xx}\|_{\lambda,Q_{b}},

where vλ,Qb=(Qb|v(t,x)|λ𝑑t𝑑x)1/λ\|v\|_{\lambda,Q_{b}}=(\int_{Q_{b}}|v(t,x)|^{\lambda}dtdx)^{1/\lambda}. Set |v|Qb=sup(t,x)Qb|v(t,x)||v|_{Q_{b}}=\sup_{(t,x)\in Q_{b}}|v(t,x)|. Now for Qb=(T1,T2)×Q_{b}=(T_{1},T_{2})\times{\mathcal{I}}, where {\mathcal{I}} is a bounded open subset of \mathbb{R}, and β(0,1)\beta\in(0,1), define the Hölder norms

|v|Qbβ=|v|Qb+supt(T1,T2),x,y|v(t,x)v(t,y)||xy|β\displaystyle|v|^{\beta}_{Q_{b}}=|v|_{Q_{b}}+\sup_{t\in(T_{1},T_{2}),x,y\in{\mathcal{I}}}|v(t,x)-v(t,y)|\cdot|x-y|^{-\beta}
+sups,t(T1,T2),x|v(s,x)v(t,x)||st|β/2,\displaystyle\qquad\qquad+\sup_{s,t\in(T_{1},T_{2}),x\in{\mathcal{I}}}|v(s,x)-v(t,x)|\cdot|s-t|^{-\beta/2},
|v|Qb1+β=|v|Qbβ+|vx|Qbβ,\displaystyle|v|^{1+\beta}_{Q_{b}}=|v|^{\beta}_{Q_{b}}+|v_{x}|^{\beta}_{Q_{b}},
|v|Qb2+β=|v|Qb1+β+|vt|Qbβ+|vxx|Qbβ.\displaystyle|v|^{2+\beta}_{Q_{b}}=|v|^{1+\beta}_{Q_{b}}+|v_{t}|^{\beta}_{Q_{b}}+|v_{xx}|^{\beta}_{Q_{b}}.
Lemma 3.3.

Under (H1)–(H4), the following holds:

(i) Equation (3.20) has a unique solution VαV^{\alpha} in Cb1,2(Q¯T)C_{b}^{1,2}(\overline{Q}_{T}) and moreover supQ¯T|Vxxα|C\sup_{\overline{Q}_{T}}|V^{\alpha}_{xx}|\leq C.

(ii) The best response

(3.22) uα=ϕα(t,x|μG()),α[0,1]\displaystyle u_{\alpha}=\phi_{\alpha}(t,x|\mu_{G}(\cdot)),\quad\alpha\in[0,1]

as the optimal control law solved from (3.20) is bounded and Borel measurable on [0,T]×[0,T]\times\mathbb{R}, and Lipschitz continuous in xx, uniformly with respect to α\alpha for the given μG()\mu_{G}(\cdot).

Proof.

(i) Denote

\mathboldHα(t,x,q)=minuU{qf~α(t,x,u)+l~α(t,x,u)}.\mathbold{H}_{\alpha}(t,x,q)=\min_{u\in U}\{q\widetilde{f}_{\alpha}^{*}(t,x,u)+\widetilde{l}_{\alpha}^{*}(t,x,u)\}.

Then (3.20) may be rewritten as

(3.23) Vtα(t,x)=\mathboldHα(t,x,Vxα)+σ22Vxxα,Vα(T,x)=0.\displaystyle-V^{\alpha}_{t}(t,x)=\mathbold{H}_{\alpha}(t,x,V^{\alpha}_{x})+\frac{\sigma^{2}}{2}V^{\alpha}_{xx},\qquad V^{\alpha}(T,x)=0.

As in the proof of [25, Theorem 5], we use Hölder and Lipschitz continuity (with respect to tt and xx, respectively) of f~α\widetilde{f}_{\alpha}^{*} and l~α\widetilde{l}_{\alpha}^{*} in Lemma 3.2, and follow the method in the proof of Theorem VI.6.2 of [13, p. 210] to show that (3.20) has a unique solution VαCb1,2(Q¯T)V^{\alpha}\in C_{b}^{1,2}(\overline{Q}_{T}), where uniqueness follows from a verification theorem using the closed-loop state process.

Next we show that VxxαV_{xx}^{\alpha} is bounded on Q¯T\overline{Q}_{T}. Take any x0x_{0}\in\mathbb{R}. Denote Br(x0)=(x0r,x0+r)B_{r}(x_{0})=(x_{0}-r,x_{0}+r) for r>0r>0, and QTx0,r=(0,T)×Br(x0)Q_{T}^{x_{0},r}=(0,T)\times B_{r}(x_{0}). We use two steps involving local estimates. Each step gets refined information about VαV^{\alpha} in a region based on available bound information in a larger region. It suffices to obtain a bound of VxxαV_{xx}^{\alpha} on QTx0,1Q_{T}^{x_{0},1} as long as this bound does not change with x0x_{0}.

Step 1. First, there exists a constant C1C_{1} such that

(3.24) supt,x,α|Vα|C1,supt,x,α|Vxα|C1.\displaystyle\sup_{t,x,\alpha}|V^{\alpha}|\leq C_{1},\quad\sup_{t,x,\alpha}|V_{x}^{\alpha}|\leq C_{1}.

The first inequality is obtained using (H1)–(H2) and the fact that VαV^{\alpha} is the value function of the associated optimal control problem. The second inequality is proven by the difference estimate of |Vα(t,x)Vα(t,y)||V^{\alpha}(t,x)-V^{\alpha}(t,y)| as in [13, p. 209].

By (H1), (H2) and (3.24), we have

supαsup(t,x)Q¯T|\mathboldHα(t,x,Vxα(t,x))|C2.\sup_{\alpha}\sup_{(t,x)\in\overline{Q}_{T}}|\mathbold{H}_{\alpha}(t,x,V_{x}^{\alpha}(t,x))|\leq C_{2}.

We use a typical method for analyzing semilinear parabolic equations. Once VαV^{\alpha} is known to be a solution of (3.23), we view VαV^{\alpha} as the solution of a linear equation with the free term \mathboldHα(t,x,Vxα)\mathbold{H}_{\alpha}(t,x,V_{x}^{\alpha}). For further estimates, we need λ>n+2\lambda>n+2 when using the norm (3.21). Fix λ=n+3=4\lambda=n+3=4. This yields the bound

Vαλ,QTx0,2(2)C3,\|V^{\alpha}\|^{(2)}_{\lambda,Q_{T}^{x_{0},2}}\leq C_{3},

where C3C_{3} depends on (C2,T,σ)(C_{2},T,\sigma) and the bound of (f,f0,l,l0)(f,f_{0},l,l_{0}) but not on x0x_{0}, α\alpha; see [13, p. 207] and also [29, p. 342] for local estimates of the Sobolev norm of solutions defined on unbounded domain using a cut-off function. Take β=1n+2λ=14\beta=1-\frac{n+2}{\lambda}=\frac{1}{4}. Subsequently, since λ>n+2\lambda>n+2, we have the Hölder estimate

(3.25) |Vα|QTx0,21+βC4Vαλ,QTx0,2(2)C3C4,\displaystyle|V^{\alpha}|_{Q^{x_{0},2}_{T}}^{1+\beta}\leq C_{4}\|V^{\alpha}\|^{(2)}_{{}_{\lambda,Q_{T}^{x_{0},2}}}\leq C_{3}C_{4},

where C4C_{4} is determined by λ=4\lambda=4 without depending on x0,αx_{0},\alpha; see [13, p. 207], [29, p. 343].

Step 2. On [0,T]××[C1,C1][0,T]\times\mathbb{R}\times[-C_{1},C_{1}], we can show \mathboldHα(t,x,q){\mathbold H}_{\alpha}(t,x,q) is Hölder continuous in tt and Lipschitz continuous in (x,q)(x,q). Denote β1=min{η,β}\beta_{1}=\min\{\eta,\beta\}. Next we view \mathboldHα(t,x,Vxα(t,x))\mathbold{H}_{\alpha}(t,x,V_{x}^{\alpha}(t,x)) as a function of (t,x)(t,x). Then by use of (3.25) we further obtain a bound on the Hölder norm:

(3.26) supαsupx0|\mathboldHα(,,Vxα)|QTx0,2β1C5.\displaystyle\sup_{\alpha}\sup_{x_{0}}|\mathbold{H}_{\alpha}(\cdot,\cdot,V_{x}^{\alpha})|^{\beta_{1}}_{Q^{x_{0},2}_{T}}\leq C_{5}.

Subsequently, by the method in [13, p. 207-208] with its cut-off function technique and [29, p. 351-352], we use (3.26) and local Hölder estimates of (3.23) to obtain

(3.27) |Vα|QTx0,12+β1C6,\displaystyle|V^{\alpha}|^{2+\beta_{1}}_{Q^{x_{0},1}_{T}}\leq C_{6},

where C6C_{6} depends on C5C_{5} but not on x0,αx_{0},\alpha. Since x0x_{0} is arbitrary, it follows that

(3.28) supαsupQ¯T|Vxxα|C6.\displaystyle\sup_{\alpha}\sup_{\overline{Q}_{T}}|V^{\alpha}_{xx}|\leq C_{6}.

(ii) By (H4), the optimal control law (3.22) as a function of (t,x)(t,x) is well defined and is bounded on [0,T]×[0,T]\times{\mathbb{R}} by compactness of UU. It is Borel measurable on Q¯T\overline{Q}_{T}; see [13, p.168]. Since SανG(x,q)S_{\alpha}^{\nu_{G}}(x,q) is Lipschitz continuous in (x,q)×[C1,C1](x,q)\in\mathbb{R}\times[-C_{1},C_{1}] and Vxα(t,x)V_{x}^{\alpha}(t,x) is Lipschitz continuous in xx\in\mathbb{R} by (3.28), uniformly with respect to α\alpha in each case, ϕα\phi_{\alpha} is uniformly Lipschitz continuous in xx. ∎

Denote

Ψα(t,x)=(Vα(t,x),Vtα(t,x),Vxα(t,x),Vxxα(t,x)),(t,x)Q¯T.\Psi^{\alpha}(t,x)=(V^{\alpha}(t,x),V_{t}^{\alpha}(t,x),V^{\alpha}_{x}(t,x),V^{\alpha}_{xx}(t,x)),\quad(t,x)\in\overline{Q}_{T}.

We prove the following continuity lemma for the solution of (3.20). For Q¯T\overline{Q}_{T}, define the compact subsets Bj={(t,x)|0tT,|x|j}B_{j}=\{(t,x)|0\leq t\leq T,|x|\leq j\}, jj\in\mathbb{N}.

Lemma 3.4.

Assume (H1)–(H5) hold and let μG()\mu_{G}(\cdot) in (3.19) be fixed. Then the following holds:

(i) For all compact set BjB_{j}, limαα|ΨαΨα|Bj=0\lim_{\alpha^{\prime}\to\alpha}|\Psi^{\alpha^{\prime}}-\Psi^{\alpha}|_{B_{j}}=0.

(ii) limααVxα(t,x)=Vxα(t,x)\lim_{\alpha^{\prime}\to\alpha}V_{x}^{\alpha^{\prime}}(t,x)=V_{x}^{\alpha}(t,x) for all (t,x)[0,T]×(t,x)\in[0,T]\times\mathbb{R}.

Proof.

It suffices to show (i) as (ii) follows immediately from (i).

Step 1. By (3.27) and the fact that the constant C6C_{6} can be selected without depending on α\alpha, there exists a constant CC such that supα|Vα|Bj2+β1C,\sup_{\alpha}|V^{\alpha}|^{2+\beta_{1}}_{B_{j}}\leq C, which implies that {Ψα,α[0,1]}\{\Psi^{\alpha},\alpha\in[0,1]\} is uniformly bounded and equicontinuous on BjB_{j}. For any sequence {αk,k1}\{\alpha_{k},k\geq 1\} converging to α\alpha, by Ascoli-Arzela’s lemma, for j=1j=1, there exists a subsequence denoted by {α¯k,k1}\{\bar{\alpha}_{k},k\geq 1\} such that Ψα¯k\Psi^{\bar{\alpha}_{k}} converges uniformly on B1B_{1}. By a diagonal argument, we may further extract a subsequence of {α¯k,k1}\{\bar{\alpha}_{k},k\geq 1\}, denoted by {α^k,k1}\{\hat{\alpha}_{k},k\geq 1\}, such that Ψα^k\Psi^{\hat{\alpha}_{k}} converges uniformly on each set BjB_{j}, j1j\geq 1. Hence there exists a function VV^{*} with continuous derivatives Vt,Vx,VxxV^{*}_{t},V^{*}_{x},V^{*}_{xx} on Q¯T\overline{Q}_{T} such that

(3.29) limkΨα^k(t,x)=Ψ(t,x),(t,x)Q¯T,\displaystyle\lim_{k\to\infty}\Psi^{\hat{\alpha}_{k}}(t,x)=\Psi^{*}(t,x),\qquad\forall(t,x)\in\overline{Q}_{T},

where Ψ=(V,Vt,Vx,Vxx)\Psi^{*}=(V^{*},V^{*}_{t},V^{*}_{x},V^{*}_{xx}). Since

Vtα^k(t,x)=\mathboldHαk(t,x,Vxα^k)+σ22Vxxα^k,Vαk(T,x)=0,-V_{t}^{\hat{\alpha}_{k}}(t,x)={\mathbold H}_{\alpha_{k}}(t,x,V_{x}^{\hat{\alpha}_{k}})+\frac{\sigma^{2}}{2}V_{xx}^{\hat{\alpha}_{k}},\quad V^{\alpha_{k}}(T,x)=0,

it follows from (3.29) that

Vt(t,x)=\mathboldHα(t,x,Vx)+σ22Vxx,V(T,x)=0.\displaystyle-V_{t}^{*}(t,x)={\mathbold H}_{\alpha}(t,x,V_{x}^{*})+\frac{\sigma^{2}}{2}V_{xx}^{*},\qquad V^{*}(T,x)=0.

We have used the fact that \mathboldHα(t,x,q){\mathbold H}_{\alpha}(t,x,q) is continuous in α\alpha due to (H5) and condition (C1) of [0,T]{\mathcal{M}}_{[0,T]}. It is clear that V=VαV^{*}=V^{\alpha} by uniqueness of the solution of (3.23). So Ψ=Ψα\Psi^{*}=\Psi^{\alpha}. Now it follows that

(3.30) limk|Ψα^kΨα|Bj=0,j.\displaystyle\lim_{k\to\infty}|\Psi^{\hat{\alpha}_{k}}-\Psi^{\alpha}|_{B_{j}}=0,\quad\forall j.

Step 2. Suppose (i) does not hold so that for some j^\hat{j} we have |ΨαΨα|Bj^|\Psi^{\alpha^{\prime}}-\Psi^{\alpha}|_{B_{\hat{j}}} does not converge to 0 as αα\alpha^{\prime}\to\alpha, which implies that there exist some ϵ0>0\epsilon_{0}>0 and a sequence {αk0}\{\alpha_{k}^{0}\} converging to α\alpha such that for each kk,

(3.31) |Ψαk0Ψα|Bj^ϵ0.\displaystyle|\Psi^{\alpha^{0}_{k}}-\Psi^{\alpha}|_{B_{\hat{j}}}\geq\epsilon_{0}.

Step 3. Recall that {αk}\{\alpha_{k}\} in Step 1 is arbitrary as long as it converges to α\alpha. Now we just take {αk}\{\alpha_{k}\} in Step 1 as {αk0}\{\alpha_{k}^{0}\}. By Step 1, there exists a subsequence of {αk0}\{\alpha^{0}_{k}\}, denoted by {α^k0}\{\hat{\alpha}^{0}_{k}\}, such that limk|Ψα^k0Ψα|Bj^=0\lim_{k\to\infty}|\Psi^{\hat{\alpha}_{k}^{0}}-\Psi^{\alpha}|_{B_{\hat{j}}}=0, which contradicts (3.31). Hence (i) holds. ∎

Lemma 3.5.

Assume (H1)–(H6). For given μG()[0,T]\mu_{G}(\cdot)\in{\mathcal{M}}_{[0,T]}, the best response ϕα(t,x|μG())\phi_{\alpha}(t,x|\mu_{G}(\cdot)) in (3.22) continuously depends on α\alpha. Specifically, for any α[0,1]\alpha\in[0,1],

(3.32) limααϕα(t,x|μG())=ϕα(t,x|μG()),t,x.\displaystyle\lim_{\alpha^{\prime}\to\alpha}\phi_{\alpha^{\prime}}(t,x|\mu_{G}(\cdot))=\phi_{\alpha}(t,x|\mu_{G}(\cdot)),\quad\forall t,x.
Proof.

The best response can be written as

ϕα(t,x|μG())=SαμG(t)(x,Vxα(t,x)),\displaystyle\phi_{\alpha}(t,x|\mu_{G}(\cdot))=S_{\alpha}^{\mu_{G}(t)}(x,V_{x}^{\alpha}(t,x)),
ϕα(t,x|μG())=SαμG(t)(x,Vxα(t,x)).\displaystyle\phi_{\alpha^{\prime}}(t,x|\mu_{G}(\cdot))=S_{\alpha^{\prime}}^{\mu_{G}(t)}(x,V_{x}^{\alpha^{\prime}}(t,x)).

It follows that

|SαμG(t)(x,Vxα(t,x))SαμG(t)(x,Vxα(t,x))|\displaystyle|S_{\alpha}^{\mu_{G}(t)}(x,V_{x}^{\alpha}(t,x))-S_{\alpha^{\prime}}^{\mu_{G}(t)}(x,V_{x}^{\alpha^{\prime}}(t,x))|
\displaystyle\leq |SαμG(t)(x,Vxα(t,x))SαμG(t)(x,Vxα(t,x))|\displaystyle|S_{\alpha}^{\mu_{G}(t)}(x,V_{x}^{\alpha}(t,x))-S_{\alpha}^{\mu_{G}(t)}(x,V_{x}^{\alpha^{\prime}}(t,x))|
+|SαμG(t)(x,Vxα(t,x))SαμG(t)(x,Vxα(t,x))|.\displaystyle+|S_{\alpha}^{\mu_{G}(t)}(x,V_{x}^{\alpha^{\prime}}(t,x))-S_{\alpha^{\prime}}^{\mu_{G}(t)}(x,V_{x}^{\alpha^{\prime}}(t,x))|.

Given μG()\mu_{G}(\cdot) we have the prior upper bound supα,t,x|Vxα(t,x)|C\sup_{\alpha,t,x}|V_{x}^{\alpha}(t,x)|\leq C. It suffices to show that (3.32) holds for any given C0>0C_{0}>0 and t[0,T]t\in[0,T], |x|C0|x|\leq C_{0}. By (H6), for the given μG(t)\mu_{G}(t), SαμG(t)(x,q)S_{\alpha}^{\mu_{G}(t)}(x,q) is uniformly continuous in α[0,1]\alpha\in[0,1], |x|C0|x|\leq C_{0}, q[C,C]q\in[-C,C]. For any ϵ>0\epsilon>0, there exists δ>0\delta>0 such that |αα|<δ|\alpha-\alpha^{\prime}|<\delta implies sup|x|C0,|q|C|SαμG(t)(x,q)SαμG(t)(x,q)|ϵ/2\sup_{|x|\leq C_{0},|q|\leq C}|S_{\alpha}^{\mu_{G}(t)}(x,q)-S_{\alpha^{\prime}}^{\mu_{G}(t)}(x,q)|\leq\epsilon/2, and moreover,

sup|x|C0|SαμG(t)(x,Vxα(t,x))SαμG(t)(x,Vxα(t,x))|ϵ2\sup_{|x|\leq C_{0}}|S_{\alpha}^{\mu_{G}(t)}(x,V_{x}^{\alpha}(t,x))-S_{\alpha}^{\mu_{G}(t)}(x,V_{x}^{\alpha^{\prime}}(t,x))|\leq\frac{\epsilon}{2}

in view of Lemma 3.4 (i). Therefore (3.32) holds. ∎

We proceed to show the existence of a solution to the GMFG equations (3.2) and (3.15) in terms of {(Vα,μα())|α[0,1]}\{(V^{\alpha},\mu_{\alpha}(\cdot))|\alpha\in[0,1]\}. For μG[0,T]\mu_{G}\in{\mathcal{M}}_{[0,T]}, denote the mapping

(ϕα)α[0,1]Γ(μG()),(\phi_{\alpha})_{\alpha\in[0,1]}\coloneqq\Gamma(\mu_{G}(\cdot)),

where the left hand side is given by (3.22) as the set of best responses with respect to μG()\mu_{G}(\cdot). Next, we combine (ϕα)α[0,1](\phi_{\alpha})_{\alpha\in[0,1]} with μG()\mu_{G}(\cdot) to determine the distribution mαm_{\alpha} of the closed-loop state process

dxα(t)=f~[xα(t),ϕα(t,xα(t)|μG()),μG(t);gα]dt+σdwα(t),\displaystyle dx_{\alpha}(t)=\widetilde{f}[x_{\alpha}(t),\phi_{\alpha}(t,x_{\alpha}(t)|\mu_{G}(\cdot)),\mu_{G}(t);g_{\alpha}]dt+\sigma dw_{\alpha}(t),

where xα(0)x_{\alpha}(0) has distribution μ0x\mu_{0}^{x}. The choice of the Brownian motion for xαx_{\alpha} is immaterial. For mαm_{\alpha} above, denote the mapping from [0,T]{\mathcal{M}}_{[0,T]} to 𝐌TG{\bf M}_{T}^{G}:

(mα)α[0,1]=Γ^(μG()).(m_{\alpha})_{\alpha\in[0,1]}=\widehat{\Gamma}(\mu_{G}(\cdot)).

Define the set

𝐌TG1Γ^([0,T])𝐌TG.\displaystyle{\bf M}_{T}^{G1}\coloneqq\widehat{\Gamma}({\mathcal{M}}_{[0,T]})\subset{\bf M}_{T}^{G}.

Now the existence analysis may be formulated as the problem of finding a fixed point of the form

(3.33) mG=Γ^Marg(mG),\displaystyle m_{G}=\widehat{\Gamma}\circ{\rm Marg}(m_{G}),

in case mG𝐌TG1m_{G}\in{\bf M}_{T}^{G1}. Note that Marg(mG)={(Margt(mα))α[0,1],0tT}{\rm Marg}(m_{G})=\{({\rm Marg}_{t}(m_{\alpha}))_{\alpha\in[0,1]},0\leq t\leq T\}.

Remark 3.6.

The fixed point problem requires mGm_{G} to be from the subset 𝐌TG1{\bf M}_{T}^{G1} of 𝐌TG{\bf M}_{T}^{G}. If one simply looks for mG𝐌TGm_{G}\in{\bf M}_{T}^{G}, the resulting μG()=Marg(mG)\mu_{G}(\cdot)={{\rm Marg}}(m_{G}) lacks required properties such as Hölder continuity in (C2), and this will cause difficulties in establishing Lemma 3.3 for the HJB equation.

Lemma 3.7.

Under (H1)–(H6), the following assertions hold:

(i) 𝐌TG1𝐌TG0{\bf M}_{T}^{G1}\subset{\bf M}_{T}^{G0}.

(ii) For any mG𝐌TG1m_{G}\in{\bf M}_{T}^{G1}, μG()Marg(mG)[0,T]\mu_{G}(\cdot)\coloneqq{{\rm Marg}}(m_{G})\in{\mathcal{M}}_{[0,T]}.

(iii) The best response ϕα(t,x|μG())\phi_{\alpha}(t,x|\mu_{G}(\cdot)) with μG()\mu_{G}(\cdot) given in (ii) is Lipschitz continuous in xx, uniformly with respect to α[0,1]\alpha\in[0,1] and mG𝐌TG1m_{G}\in{\bf M}_{T}^{G1}.

Proof.

(i) and (ii) For mG𝐌TG1m_{G}\in{\bf M}_{T}^{G1}, there exists μG[0,T]\mu^{\prime}_{G}\in{\mathcal{M}}_{[0,T]} such that mG=Γ^(μG()).m_{G}=\widehat{\Gamma}(\mu^{\prime}_{G}(\cdot)). To estimate DT(mα,mα¯)D_{T}(m_{\alpha},m_{\bar{\alpha}}) and W1(μα(t),μα¯(t))W_{1}(\mu_{\alpha}(t),\mu_{\bar{\alpha}}(t)), let xαx_{\alpha} and xα¯x_{\bar{\alpha}} be state processes generated by (3.10) with μG\mu_{G}^{\prime}, the same initial state and Brownian motion under the control laws ϕα(t,x|μG())\phi_{\alpha}(t,x|\mu_{G}^{\prime}(\cdot)) and ϕα¯(t,x|μG())\phi_{\bar{\alpha}}(t,x|\mu_{G}^{\prime}(\cdot)), respectively. Then DT(mα,mα¯)EsuptT|xα(t)xα¯(t)|D_{T}(m_{\alpha},m_{\bar{\alpha}})\leq E\sup_{t\leq T}|x_{\alpha}(t)-x_{\bar{\alpha}}(t)| and W1(μα(t),μα¯(t))E|xα(t)xα¯(t)|W_{1}(\mu_{\alpha}(t),\mu_{\bar{\alpha}}(t))\leq E|x_{\alpha}(t)-x_{\bar{\alpha}}(t)|. Fixing α¯\bar{\alpha}, we have

(3.34) |xα(t)xα¯(t)|\displaystyle|x_{\alpha}(t)-x_{\bar{\alpha}}(t)|\leq 0t|f~[xα(s),ϕα(s,xα(s)|μG()),μG(s);gα]\displaystyle\int_{0}^{t}|\widetilde{f}[x_{\alpha}(s),\phi_{\alpha}(s,x_{\alpha}(s)|\mu^{\prime}_{G}(\cdot)),\mu^{\prime}_{G}(s);g_{\alpha}]
f~[xα¯(s),ϕα¯(s,xα¯(s)|μG()),μG(s);gα¯]|ds.\displaystyle\quad-\widetilde{f}[x_{\bar{\alpha}}(s),\phi_{\bar{\alpha}}(s,x_{\bar{\alpha}}(s)|\mu^{\prime}_{G}(\cdot)),\mu^{\prime}_{G}(s);g_{\bar{\alpha}}]|ds.

Denote

δ1=|f0[xα¯(s),ϕα¯(s,xα¯(s)|μG()),μα(s)]f0[xα¯(s),ϕα¯(s,xα¯(s)|μG()),μα¯(s)]|,\displaystyle\delta_{1}=|f_{0}[x_{\bar{\alpha}}(s),\phi_{\bar{\alpha}}(s,x_{\bar{\alpha}}(s)|\mu^{\prime}_{G}(\cdot)),\mu^{\prime}_{\alpha}(s)]-f_{0}[x_{\bar{\alpha}}(s),\phi_{\bar{\alpha}}(s,x_{\bar{\alpha}}(s)|\mu^{\prime}_{G}(\cdot)),\mu^{\prime}_{\bar{\alpha}}(s)]|,
δ2=|f[xα¯(s),ϕα¯(s,xα¯(s)|μG()),μG(s);gα]f[xα¯(s),ϕα¯(s,xα¯(s)|μG()),μG(s);gα¯]|.\displaystyle\delta_{2}=|f[x_{\bar{\alpha}}(s),\phi_{\bar{\alpha}}(s,x_{\bar{\alpha}}(s)|\mu^{\prime}_{G}(\cdot)),\mu^{\prime}_{G}(s);g_{\alpha}]-f[x_{\bar{\alpha}}(s),\phi_{\bar{\alpha}}(s,x_{\bar{\alpha}}(s)|\mu^{\prime}_{G}(\cdot)),\mu^{\prime}_{G}(s);g_{\bar{\alpha}}]|.

Then by (3.34) and the Lipschitz continuity in xx of ϕα\phi_{\alpha} in Lemma 3.3 (ii), we obtain

(3.35) |xα(t)xα¯(t)|C10t|xα(s)xα¯(s)|𝑑s\displaystyle|x_{\alpha}(t)-x_{\bar{\alpha}}(t)|\leq C_{1}\int_{0}^{t}|x_{\alpha}(s)-x_{\bar{\alpha}}(s)|ds
+C20t{|ϕα(s,xα¯(s)|μG())ϕα¯(s,xα¯(s)|μG())|+δ1(s)+δ2(s)}ds,\displaystyle+C_{2}\int_{0}^{t}\{|\phi_{\alpha}(s,x_{\bar{\alpha}}(s)|\mu^{\prime}_{G}(\cdot))-\phi_{\bar{\alpha}}(s,x_{\bar{\alpha}}(s)|\mu^{\prime}_{G}(\cdot))|+\delta_{1}(s)+\delta_{2}(s)\}ds,

where C2C_{2} depends only on the Lipschitz constants of f0,ff_{0},f; and C1C_{1} does not change with α\alpha for the fixed μG\mu_{G}^{\prime}. Since W1(μα(s),μα¯(s))0W_{1}(\mu^{\prime}_{\alpha}(s),\mu^{\prime}_{\bar{\alpha}}(s))\to 0 as αα¯\alpha\to\bar{\alpha}, by (H2) Eδ1(s)0E\delta_{1}(s)\to 0 as αα¯\alpha\to\bar{\alpha}. By (H5), we have Eδ2(s)0E\delta_{2}(s)\to 0 as αα¯\alpha\to\bar{\alpha}. Then using Lemma 3.5 and boundedness of the integrand below, we obtain

limαα¯E0T{|ϕα(s,xα¯(s)|μG())ϕα¯(s,xα¯(s)|μG())|+δ1(s)+δ2(s)}ds=0.\lim_{\alpha\to\bar{\alpha}}E\int_{0}^{T}\{|\phi_{\alpha}(s,x_{\bar{\alpha}}(s)|\mu^{\prime}_{G}(\cdot))-\phi_{\bar{\alpha}}(s,x_{\bar{\alpha}}(s)|\mu^{\prime}_{G}(\cdot))|+\delta_{1}(s)+\delta_{2}(s)\}ds=0.

By Gronwall’s lemma and (3.35), it follows that

(3.36) limαα¯Esup0tT|xα(t)xα¯(t)|=0.\displaystyle\lim_{\alpha\to\bar{\alpha}}E\sup_{0\leq t\leq T}|x_{\alpha}(t)-x_{\bar{\alpha}}(t)|=0.

Subsequently, as αα¯\alpha\to\bar{\alpha}, we obtain DT(mα,mα¯)0D_{T}(m_{\alpha},m_{\bar{\alpha}})\to 0, which implies (i); in addition, W1(μα(t),μα¯(t))0W_{1}(\mu_{\alpha}(t),\mu_{\bar{\alpha}}(t))\to 0, which verifies condition (C1) of [0,T]{\mathcal{M}}_{[0,T]} for μG\mu_{G}. Since each mαm_{\alpha} is the distribution of xαx_{\alpha}, for μG()\mu_{G}(\cdot) we take the Hölder parameter η=1/2\eta=1/2 and a constant ChC_{h} independent of μG\mu_{G}^{\prime} for (C2). So (ii) holds.

(iii) Due to the choice of η\eta and ChC_{h} for μG()\mu_{G}(\cdot) in (ii), we may select a fixed constant C5C_{5} in (3.26), which does not change with (α,μG())(\alpha,\mu_{G}(\cdot)). Subsequently the upper bound C6C_{6} in (3.28) for |Vxxα||V_{xx}^{\alpha}| does not change with α[0,1],μG()Marg(Γ^([0,T]))\alpha\in[0,1],\mu_{G}(\cdot)\in{\rm Marg}(\widehat{\Gamma}({\mathcal{M}}_{[0,T]})). This ensures a uniform bound for the Lipschitz constant for xx in ϕα\phi_{\alpha}. ∎

We introduce the sensitivity condition.

(H7) For mG,m¯G𝐌TG1=Γ^([0,T])m_{G},\bar{m}_{G}\in{\bf M}_{T}^{G1}=\widehat{\Gamma}({\mathcal{M}}_{[0,T]}), there exists a constant c1c_{1} such that

(3.37) supt,x,α|ϕα(t,x|μG())ϕ¯α(t,x|μ¯G())|c1d(mG,m¯G),\displaystyle\sup_{t,x,\alpha}|\phi_{\alpha}(t,x|\mu_{G}(\cdot))-\bar{\phi}_{\alpha}(t,x|\bar{\mu}_{G}(\cdot))|\leq c_{1}d(m_{G},\bar{m}_{G}),

where the set of control laws {ϕα(t,x|μG()),α[0,1]}\{\phi_{\alpha}(t,x|\mu_{G}(\cdot)),\alpha\in[0,1]\} (resp., {ϕ¯α(t,x|μ¯G()),α[0,1]}\{\bar{\phi}_{\alpha}(t,x|\bar{\mu}_{G}(\cdot)),\alpha\in[0,1]\}) is determined by use of μG=Marg(mG)\mu_{G}={\rm Marg}(m_{G}) (resp., μ¯G=Marg(m¯G)\bar{\mu}_{G}={\rm Marg}(\bar{m}_{G})) in the optimal control problem specified by (3.10) and (3.12) with the graphon section gαg_{\alpha}.

Assumption (H7) is a generalization from the finite type model in [25] where an illustration via a linear model is presented. Related sensitivity conditions are studied in [28].

Let (ϕα)α[0,1](\phi_{\alpha})_{\alpha\in[0,1]} in (3.22) be applied by all agents, where μG()[0,T]\mu_{G}(\cdot)\in{\mathcal{M}}_{[0,T]}. We consider the following generalized McKean-Vlasov equation

(3.38) dxα(t)=f~[xα(t),ϕα(t,xα(t)|μG),νG(t);gα]dt+σdwα(t),\displaystyle dx_{\alpha}(t)=\widetilde{f}[x_{\alpha}(t),\phi_{\alpha}(t,x_{\alpha}(t)|\mu_{G}),\nu_{G}(t);g_{\alpha}]dt+\sigma dw_{\alpha}(t),

where xα(0)x_{\alpha}(0) is given with distribution μ0x\mu_{0}^{x}. For this equation, νG\nu_{G} is part of the solution. If νG\nu_{G} is determined, we have a unique solution xαx_{\alpha} on [0,T][0,T] which further determines its law as the measure mαm_{\alpha} on (CT,T)(C_{T},{\mathcal{F}}_{T}). Note that mαm_{\alpha} does not depend on the choice of the standard Brownian motion wαw_{\alpha}. We look for νG[0,T]\nu_{G}\in{\mathcal{M}}_{[0,T]} to satisfy the condition:

(3.39) Margt(mα)=να(t),α[0,1],t[0,T],\displaystyle{\rm Marg}_{t}(m_{\alpha})=\nu_{\alpha}(t),\quad\forall\alpha\in[0,1],\ t\in[0,T],

i.e., να(t)\nu_{\alpha}(t) is the law of xα(t)x_{\alpha}(t) for all α,t\alpha,t (and we say (xα)0α1(x_{\alpha})_{0\leq\alpha\leq 1} is consistent with νG\nu_{G}).

Lemma 3.8.

Assume (H1)–(H6). For the best response control law ϕα(t,xα|μG())\phi_{\alpha}(t,x_{\alpha}|\mu_{G}(\cdot)) in (3.22), where μG()[0,T]\mu_{G}(\cdot)\in{\mathcal{M}}_{[0,T]}, there exists a unique νG()\nu_{G}(\cdot) for (3.38) satisfying (3.39).

Proof.

In order to solve (xα,νG)(x_{\alpha},\nu_{G}) in (3.38), we specify the law of the process xαx_{\alpha} instead of just its marginal να(t)\nu_{\alpha}(t). This extends the fixed point idea for treating standard McKean-Vlasov equations [41].

For (mα)α[0,1]𝐌TG0(m_{\alpha})_{\alpha\in[0,1]}\in{\bf M}_{T}^{G0} , we determine νG1\nu_{G}^{1} according to να1(t)=Margt(mα)\nu^{1}_{\alpha}(t)={\rm Marg}_{t}(m_{\alpha}), which is used in (3.38) by taking νG=νG1\nu_{G}=\nu_{G}^{1} to solve xαx_{\alpha} on [0,T][0,T]. Let mαnewm_{\alpha}^{\rm new} denote the law of xαx_{\alpha}. It in general does not satisfy Margt(mαnew)=να(t){\rm Marg}_{t}(m_{\alpha}^{\rm new})=\nu_{\alpha}(t) for all tt. Denote the mapping

(mαnew)α[0,1]=Φ𝐌TG0((mα)α[0,1]).(m_{\alpha}^{\rm new})_{\alpha\in[0,1]}=\Phi_{{\bf M}_{T}^{G0}}((m_{\alpha})_{\alpha\in[0,1]}).

By (H5) and Lemma 3.5, Φ𝐌TG0\Phi_{{\bf M}_{T}^{G0}} is a mapping from 𝐌TG0{\bf M}_{T}^{G0} to itself. Similarly, from (m¯α)α[0,1]𝐌TG0(\bar{m}_{\alpha})_{\alpha\in[0,1]}\in{{\bf M}_{T}^{G0}} we determine ν¯G1\bar{\nu}_{G}^{1} for (3.38) and solve x¯α\bar{x}_{\alpha} with its law m¯αnew\bar{m}^{\rm new}_{\alpha}. Denote

(m¯αnew)α[0,1]=Φ𝐌TG0((m¯α)α[0,1]).(\bar{m}_{\alpha}^{\rm new})_{\alpha\in[0,1]}=\Phi_{{\bf M}_{T}^{G0}}((\bar{m}_{\alpha})_{\alpha\in[0,1]}).

If h(x,y)h(x,y) is a bounded Lipschitz continuous function with |h(x,y)h(x¯,y¯)|C1|xx¯|+C2(|yy¯|1)|h(x,y)-h(\bar{x},\bar{y})|\leq C_{1}|x-\bar{x}|+C_{2}(|y-\bar{y}|\wedge 1), we have

|h(x,y)g(α,β)νβ1(t,dy)𝑑βh(x¯,y¯)g(α,β)νβ2(t,dy¯)𝑑β|\displaystyle\Big{|}\int h(x,y)g(\alpha,\beta)\nu_{\beta}^{1}(t,dy)d\beta-\int h(\bar{x},\bar{y})g(\alpha,\beta)\nu_{\beta}^{2}(t,d\bar{y})d\beta\Big{|}
\displaystyle\leq C1|xx¯|+supβ|h(x¯,y)νβ1(t,dy)h(x¯,y¯)νβ2(t,dy¯)|\displaystyle C_{1}|x-\bar{x}|+\sup_{\beta}\Big{|}\int h(\bar{x},y)\nu_{\beta}^{1}(t,dy)-\int h(\bar{x},\bar{y})\nu_{\beta}^{2}(t,d\bar{y})\Big{|}
=\displaystyle= C1|xx¯|+supβ|CTh(x¯,Xt(ω))𝑑mβ(ω)CTh(x¯,Xt(ω¯))𝑑m¯β(ω¯)|\displaystyle C_{1}|x-\bar{x}|+\sup_{\beta}\Big{|}\int_{C_{T}}h(\bar{x},X_{t}(\omega))dm_{\beta}(\omega)-\int_{C_{T}}h(\bar{x},X_{t}(\bar{\omega}))d\bar{m}_{\beta}(\bar{\omega})\Big{|}
\displaystyle\leq C1|xx¯|+C2supβCT×CT(|Xt(ω)Xt(ω¯)|1)𝑑m^β(ω,ω¯),\displaystyle C_{1}|x-\bar{x}|+C_{2}\sup_{\beta}\int_{C_{T}\times C_{T}}(|X_{t}(\omega)-X_{t}(\bar{\omega})|\wedge 1)d\widehat{m}_{\beta}(\omega,\bar{\omega}),

where XX is the canonical process, ω,ω¯CT\omega,\bar{\omega}\in C_{T}, and m^β\widehat{m}_{\beta} is any coupling of mβm_{\beta} and m¯β\bar{m}_{\beta}. Hence

|h(x,y)g(α,β)νβ1(t,dy)𝑑βh(x¯,y¯)g(α,β)νβ2(t,dy¯)𝑑β|\displaystyle|\int h(x,y)g(\alpha,\beta)\nu_{\beta}^{1}(t,dy)d\beta-\int h(\bar{x},\bar{y})g(\alpha,\beta)\nu_{\beta}^{2}(t,d\bar{y})d\beta|
(3.40) \displaystyle\leq\ C1|xx¯|+C2supβDt(mβ,m¯β).\displaystyle C_{1}|x-\bar{x}|+C_{2}\sup_{\beta}D_{t}(m_{\beta},\bar{m}_{\beta}).

By (H2), (H3), the uniform Lipschitz continuity of ϕα\phi_{\alpha} in xx by Lemma 3.3 (ii), and (3.40), we obtain

|f~[xα,ϕα(t,xα|μG),νG1(t);gα]f~[x¯α,ϕα(t,x¯α|μG),νG2(t);gα]|\displaystyle|\widetilde{f}[x_{\alpha},\phi_{\alpha}(t,x_{\alpha}|\mu_{G}),\nu_{G}^{1}(t);g_{\alpha}]-\widetilde{f}[\bar{x}_{\alpha},\phi_{\alpha}(t,\bar{x}_{\alpha}|\mu_{G}),\nu_{G}^{2}(t);g_{\alpha}]|
\displaystyle\leq C1(|xαx¯α|1)+C2supβDt(mβ,m¯β).\displaystyle C_{1}(|x_{\alpha}-\bar{x}_{\alpha}|\wedge 1)+C_{2}\sup_{\beta}D_{t}(m_{\beta},\bar{m}_{\beta}).

Hence by (3.38),

supst|xα(s)x¯α(s)|\displaystyle\sup_{s\leq t}|x_{\alpha}(s)-\bar{x}_{\alpha}(s)| C10t|xα(s)x¯α(s)|1ds\displaystyle\leq C_{1}\int_{0}^{t}|x_{\alpha}(s)-\bar{x}_{\alpha}(s)|\wedge 1ds
+C30tsupβ|Ds(mβ,m¯β)|ds.\displaystyle+C_{3}\int_{0}^{t}\sup_{\beta}|D_{s}(m_{\beta},\bar{m}_{\beta})|ds.

Therefore, by Gronwall’s lemma,

supst|xα(s)x¯α(s)|1C40tsupβ|Ds(mβ,m¯β)|ds,\displaystyle\sup_{s\leq t}|x_{\alpha}(s)-\bar{x}_{\alpha}(s)|\wedge 1\leq C_{4}\int_{0}^{t}\sup_{\beta}|D_{s}(m_{\beta},\bar{m}_{\beta})|ds,

which combined with the definition of the Wasserstein metric Dt(,)D_{t}(\cdot,\cdot) implies that

(3.41) supβ|Dt(mβnew,m¯βnew)|C40tsupβ|Ds(mβ,m¯β)|ds.\displaystyle\sup_{\beta}|D_{t}(m^{\rm new}_{\beta},\bar{m}^{\rm new}_{\beta})|\leq C_{4}\int_{0}^{t}\sup_{\beta}|D_{s}(m_{\beta},\bar{m}_{\beta})|ds.

By iterating (3.41) as in [41, p. 174], we can show that for a sufficiently large k0k_{0}, Φ𝐌TG0k0\Phi_{{\bf M}_{T}^{G0}}^{k_{0}} is a contraction. We can further show that {Φ𝐌TG0k(mG),k1}\{\Phi_{{\bf M}_{T}^{G0}}^{k}(m_{G}),k\geq 1\} is a Cauchy sequence, and we obtain a unique fixed point mGm^{*}_{G} for Φ𝐌TG0\Phi_{{\bf M}_{T}^{G0}}. Then we obtain a solution of (3.38) by taking να(t)=Margt(mα)\nu_{\alpha}(t)={\rm Marg}_{t}(m^{*}_{\alpha}). If there are two different solutions with νGνG\nu_{G}\neq\nu_{G}^{\prime}, we can derive a contradiction by using uniqueness of the fixed point of Φ𝐌TG0\Phi_{{\mathbf{M}}_{T}^{G0}}. ∎

Now we consider two sets of best response control laws (ϕα(t,xα|μG))α[0,1](\phi_{\alpha}(t,x_{\alpha}|\mu_{G}))_{\alpha\in[0,1]} and (ϕ¯α(t,xα|μ¯G))α[0,1](\bar{\phi}_{\alpha}(t,x_{\alpha}|\bar{\mu}_{G}))_{\alpha\in[0,1]}, where μG=Marg(mG),μ¯G=Marg(m¯G)\mu_{G}={\rm Marg}(m_{G}),\bar{\mu}_{G}={\rm Marg}(\bar{m}_{G}) for mG,m¯G𝐌TG1m_{G},\bar{m}_{G}\in{\bf M}_{T}^{G1} (then clearly μG,μ¯G[0,T]\mu_{G},\bar{\mu}_{G}\in{\mathcal{M}}_{[0,T]}), and use Lemma 3.8 to solve (xα,νG)(x_{\alpha},\nu_{G}) and (xα,ν¯G)(x^{\prime}_{\alpha},\bar{\nu}_{G}) from the generalized MV-SDEs

(3.42) dxα=f~[xα,ϕα(t,xα|μG),νG(t);gα]dt+σdwα(t),\displaystyle dx_{\alpha}=\widetilde{f}[x_{\alpha},\phi_{\alpha}(t,x_{\alpha}|\mu_{G}),\nu_{G}(t);g_{\alpha}]dt+\sigma dw_{\alpha}(t),
(3.43) dxα=f~[xα,ϕ¯α(t,xα|μ¯G),ν¯G(t);gα]dt+σdwα(t),\displaystyle dx_{\alpha}^{\prime}=\widetilde{f}[x_{\alpha}^{\prime},\bar{\phi}_{\alpha}(t,x^{\prime}_{\alpha}|\bar{\mu}_{G}),\bar{\nu}_{G}(t);g_{\alpha}]dt+\sigma dw_{\alpha}(t),

where xα(0)=xα(0)x^{\prime}_{\alpha}(0)=x_{\alpha}(0) is given. Let mαmvm_{\alpha}^{\rm mv} (resp., m¯αmv\bar{m}_{\alpha}^{\rm mv}) denote the law of xαx_{\alpha} (resp., xαx^{\prime}_{\alpha}). The following lemma is a generalization of [25, Lemma 9] to the graphon network case.

Lemma 3.9.

For (3.42) and (3.43) there exists a constant c2c_{2} independent of (mG,m¯G)(m_{G},\bar{m}_{G}) such that

supαDT(mαmv,m¯αmv)c2supt,x,α|ϕα(t,x|μG())ϕ¯α(t,x|μ¯G())|.\displaystyle\sup_{\alpha}D_{T}(m_{\alpha}^{\rm mv},\bar{m}_{\alpha}^{\rm mv})\leq c_{2}\sup_{t,x,\alpha}|\phi_{\alpha}(t,x|\mu_{G}(\cdot))-\bar{\phi}_{\alpha}(t,x|\bar{\mu}_{G}(\cdot))|.
Proof.

For (3.42)–(3.43), denote

Δs=\displaystyle\Delta_{s}=\ f~[xα(s),ϕα(s,xα(s)|μG),νG(s);gα]f~[xα(s),ϕ¯α(s,xα(s)|μ¯G),ν¯G(s);gα].\displaystyle\widetilde{f}[x_{\alpha}(s),\phi_{\alpha}(s,x_{\alpha}(s)|\mu_{G}),\nu_{G}(s);g_{\alpha}]-\widetilde{f}[x_{\alpha}^{\prime}(s),\bar{\phi}_{\alpha}(s,x^{\prime}_{\alpha}(s)|\bar{\mu}_{G}),\bar{\nu}_{G}(s);g_{\alpha}].

We have

(3.44) xα(t)xα(t)=0tΔs𝑑s.\displaystyle x_{\alpha}(t)-x_{\alpha}^{\prime}(t)=\int_{0}^{t}\Delta_{s}ds.

Noting να(t)=Margt(mαmv)\nu_{\alpha}(t)={\rm Marg}_{t}(m_{\alpha}^{\rm mv}) and ν¯α(t)=Margt(m¯αmv)\bar{\nu}_{\alpha}(t)={\rm Marg}_{t}(\bar{m}_{\alpha}^{\rm mv}), we have

|Δs|\displaystyle|\Delta_{s}|\leq |f~[xα(s),ϕα(s,xα(s)|μG),νG(s);gα]f~[xα(s),ϕα(s,xα(s)|μG),ν¯G(s);gα]|\displaystyle|\widetilde{f}[x_{\alpha}(s),\phi_{\alpha}(s,x_{\alpha}(s)|\mu_{G}),\nu_{G}(s);g_{\alpha}]-\widetilde{f}[x_{\alpha}^{\prime}(s),\phi_{\alpha}(s,x^{\prime}_{\alpha}(s)|\mu_{G}),\bar{\nu}_{G}(s);g_{\alpha}]|
+|f~[xα(s),ϕα(s,xα(s)|μG),ν¯G(s);gα]f~[xα(s),ϕ¯α(s,xα(s)|μ¯G),ν¯G(s);gα]|\displaystyle\hskip-14.22636pt+|\widetilde{f}[x_{\alpha}^{\prime}(s),\phi_{\alpha}(s,x^{\prime}_{\alpha}(s)|\mu_{G}),\bar{\nu}_{G}(s);g_{\alpha}]-\widetilde{f}[x_{\alpha}^{\prime}(s),\bar{\phi}_{\alpha}(s,x^{\prime}_{\alpha}(s)|\bar{\mu}_{G}),\bar{\nu}_{G}(s);g_{\alpha}]|
\displaystyle\leq C1|xα(s)xα(s)|+C2supβDs(mβmv,m¯βmv)\displaystyle C_{1}|x_{\alpha}(s)-x_{\alpha}^{\prime}(s)|+C_{2}\sup_{\beta}D_{s}(m_{\beta}^{\rm mv},\bar{m}^{\rm mv}_{\beta})
(3.45) +C3supt,x|ϕα(t,x|μG())ϕ¯α(t,x|μ¯G())|,\displaystyle+C_{3}\sup_{t,x}|\phi_{\alpha}(t,x|\mu_{G}(\cdot))-\bar{\phi}_{\alpha}(t,x|\bar{\mu}_{G}(\cdot))|,

where C1C_{1}, C2C_{2} and C3C_{3} do not depend on (α,mG,m¯G)(\alpha,m_{G},\bar{m}_{G}). The difference term on the first line is estimated by the method in (3.40). We have used the fact that ϕα\phi_{\alpha} is uniformly Lipschitz in xx by Lemma 3.7 (iii). Therefore, by (3.44)–(3.45),

|xα(t)xα(t)|\displaystyle|x_{\alpha}(t)-x_{\alpha}^{\prime}(t)|\leq 0t[C1|xα(s)xα(s)|+C2supβDs(mβmv,m¯βmv)]𝑑s\displaystyle\int_{0}^{t}\Big{[}C_{1}|x_{\alpha}(s)-x_{\alpha}^{\prime}(s)|+C_{2}\sup_{\beta}D_{s}(m^{\rm mv}_{\beta},\bar{m}^{\rm mv}_{\beta})\Big{]}ds
+C3tsupt,x|ϕα(t,x|μG())ϕ¯α(t,x|μ¯G())|.\displaystyle+C_{3}t\sup_{t,x}|\phi_{\alpha}(t,x|\mu_{G}(\cdot))-\bar{\phi}_{\alpha}(t,x|\bar{\mu}_{G}(\cdot))|.

By Gronwall’s lemma, we obtain

sup0st|xα(s)xα(s)|1\displaystyle\sup_{0\leq s\leq t}|x_{\alpha}(s)-x_{\alpha}^{\prime}(s)|\wedge 1 eC1tC20tsupβDs(mβmv,m¯βmv)ds\displaystyle\leq e^{C_{1}t}C_{2}\int_{0}^{t}\sup_{\beta}D_{s}(m^{\rm mv}_{\beta},\bar{m}^{\rm mv}_{\beta})ds
+eC1tC3tsupt,x|ϕα(t,x|μG())ϕ¯α(t,x|μ¯G())|,\displaystyle+e^{C_{1}t}C_{3}t\sup_{t,x}|\phi_{\alpha}(t,x|\mu_{G}(\cdot))-\bar{\phi}_{\alpha}(t,x|\bar{\mu}_{G}(\cdot))|,

which again by the definition of the metric Dt(,)D_{t}(\cdot,\cdot) leads to

(3.46) supαDt(mαmv,m¯αmv)\displaystyle\sup_{\alpha}D_{t}(m^{\rm mv}_{\alpha},\bar{m}^{\rm mv}_{\alpha})\leq eC1tC20tsupαDs(mαmv,m¯αmv)ds\displaystyle e^{C_{1}t}C_{2}\int_{0}^{t}\sup_{\alpha}D_{s}(m^{\rm mv}_{\alpha},\bar{m}^{\rm mv}_{\alpha})ds
+eC1tC3tsupt,x,α|ϕα(t,x|μG())ϕ¯α(t,x|μ¯G())|.\displaystyle+e^{C_{1}t}C_{3}t\sup_{t,x,\alpha}|\phi_{\alpha}(t,x|\mu_{G}(\cdot))-\bar{\phi}_{\alpha}(t,x|\bar{\mu}_{G}(\cdot))|.

The lemma follows from applying Gronwall’s lemma to (3.46). ∎

3.4. Existence Theorem

We state the main result on the existence and uniqueness of solutions to the GMFG equation system. We introduce a contraction condition:

(H8) c1c2<1c_{1}c_{2}<1, where c1c_{1} is the constant in the sensitivity condition (H7) and c2c_{2} is specified in Lemma 3.9.

Remark 3.10.

By SDE estimates, one can obtain refined bound information on c2c_{2}. When the coupling effect is weak or TT is small, a small value for c2c_{2} can be obtained.

Remark 3.11.

For linear models, a verification of the contraction condition can be done under reasonable model parameters, as in [25].

Theorem 3.12.

Under (H1)–(H8), there exists a unique solution (Vα,μα())α[0,1](V^{\alpha},\mu_{\alpha}(\cdot))_{\alpha\in[0,1]} to the GMFG equations (3.2) and (3.15), which (i) gives the feedback control best response (BR) strategy φ(t,xα|μG();gα)\varphi(t,x_{\alpha}|{\mu}_{G}(\cdot);g_{\alpha}) depending only upon the agent’s state and the ensemble μG\mu_{G} of local mean fields (i.e. (xα,μG)(x_{\alpha},{\mu}_{G})), and (ii) generates a Nash equilibrium.

Proof.

Step 1 – We return to the fixed point equation (3.33), which is redisplayed below:

(3.47) mG=Γ^Marg(mG),\displaystyle m_{G}=\widehat{\Gamma}\circ{\rm Marg}(m_{G}),

where mG=(mα)α[0,1]𝐌TG1m_{G}=(m_{\alpha})_{\alpha\in[0,1]}\in{\bf M}_{T}^{G1}. For mG𝐌TG1m_{G}\in{\bf M}_{T}^{G1}, the Hölder continuity in tt of the regenerated μG()=Marg(mG)\mu_{G}(\cdot)={\rm Marg}(m_{G}) can be checked by elementary SDE estimates by adapting the proof of [25, Lemma 7].

Step 2 – Take a general mG𝐌TG1m_{G}\in{\bf M}_{T}^{G1} to determine μG=Marg(mG)\mu_{G}={\rm Marg}(m_{G}) and ϕα(t,xα|μG())\phi_{\alpha}(t,x_{\alpha}|\mu_{G}(\cdot)). When m¯G𝐌TG1\bar{m}_{G}\in{\bf M}_{T}^{G1} is used, we determine μ¯G\bar{\mu}_{G} and ϕ¯α(t,xα|μ¯G())\bar{\phi}_{\alpha}(t,x_{\alpha}|\bar{\mu}_{G}(\cdot)). Once the set of strategies (ϕα)α[0,1](\phi_{\alpha})_{\alpha\in[0,1]} is applied to the generalized MV equation (3.38), by Lemma 3.8, we may solve for (xα,νG())(x_{\alpha},\nu_{G}(\cdot)) such that xαx_{\alpha} has the law mαnewm_{\alpha}^{\rm new} and Margt(mαnew)=να(t){\rm Marg}_{t}(m_{\alpha}^{\rm new})=\nu_{\alpha}(t). This is done in parallel for m¯G\bar{m}_{G} to generate m¯αnew\bar{m}_{\alpha}^{\rm new}. We accordingly determine mGnewm_{G}^{\rm new} and m¯Gnew\bar{m}_{G}^{\rm new}.

Step 3 – By (3.37) and Lemma 3.9, we obtain

supαDT(mαnew,m¯αnew)c1c2d(mG,m¯G),\displaystyle\sup_{\alpha}D_{T}(m_{\alpha}^{\rm new},\bar{m}_{\alpha}^{\rm new})\leq c_{1}c_{2}d(m_{G},\bar{m}_{G}),

which implies

d(mGnew,m¯Gnew)c1c2d(mG,m¯G).d(m_{G}^{\rm new},\bar{m}_{G}^{\rm new})\leq c_{1}c_{2}d(m_{G},\bar{m}_{G}).

Based on the above contraction property, we construct a Cauchy sequence in the complete metric space 𝐌TG{\bf M}_{T}^{G} by iterating with mGm_{G} and establish existence of a solution to the GMFG equation system. To show uniqueness, suppose mGm_{G} and m~G\tilde{m}_{G} are two fixed points to (3.47). We obtain d(mG,m~G)c1c2d(mG,m~G)d(m_{G},\tilde{m}_{G})\leq c_{1}c_{2}d(m_{G},\tilde{m}_{G}), which implies mG=m~Gm_{G}=\tilde{m}_{G}.

The Nash equilibrium property follows from the best response property of ϕα\phi_{\alpha} for a given vertex α\alpha. ∎

3.5. An Example on Lipschitz feedback

The main analysis in Section 3 relies on (H4) to ensure Lipschitz feedback. We provide a concrete model to check this assumption.

Example 3.13.

The dynamics and cost have

f0(x,u,y)=f0(x,y)u,f(x,u,y)=f(x,y)u,\displaystyle f_{0}(x,u,y)=f_{0}(x,y)u,\quad f(x,u,y)=f(x,y)u,
l0(x,u,y)=l1(x,y)+l2(x,y)u2,l(x,u,y)=l3(x,y)+l4(x,y)u2,\displaystyle l_{0}(x,u,y)=l_{1}(x,y)+{l}_{2}(x,y)u^{2},\quad l(x,u,y)=l_{3}(x,y)+l_{4}(x,y)u^{2},

where x,yx,y\in\mathbb{R} and uU=[a,b]u\in U=[a,b]. The functions f0f_{0}, ff, l1l_{1}, l2l_{2}, l3l_{3}, l4l_{4} satisfy (H1)–(H3), and there exists c0>0c_{0}>0 such that l2,l4c0l_{2},l_{4}\geq c_{0} for all x,yx,y.

Given νGC([0,1],𝒫1())\nu_{G}\in C([0,1],{\mathcal{P}}_{1}({\mathbb{R}})), we check the minimizer of

SανG(x,q)=argminuU{q(f0[x,να]+f[x,νG;gα])u+(l2[x,να]+l4[x,νG;gα])u2},\displaystyle S_{\alpha}^{\nu_{G}}(x,q)=\arg\min_{u\in U}\{q(f_{0}[x,\nu_{\alpha}]+f[x,\nu_{G};g_{\alpha}])u+(l_{2}[x,\nu_{\alpha}]+l_{4}[x,\nu_{G};g_{\alpha}])u^{2}\},

where x,qx,q\in\mathbb{R}.

Proposition 3.14.

Given any compact interval {\mathcal{I}}, SανG(x,q)S_{\alpha}^{\nu_{G}}(x,q) in Example 3.13 is a singleton and Lipschitz continuous in (x,q)(x,q), where xx\in\mathbb{R} and qq\in{\mathcal{I}}, uniformly with respect to (νG,α)(\nu_{G},\alpha).

Proof.

Consider the function Φ(u)=u22su\Phi(u)=u^{2}-2su, where uUu\in U and ss is a parameter. Its minimum is attained at the unique point

u^=Θ(s){aifsa,sifa<s<b,bifsb.\hat{u}=\Theta(s)\coloneqq\begin{cases}a&\mbox{if}\quad s\leq a,\\ s&\mbox{if}\quad a<s<b,\\ b&\mbox{if}\quad s\geq b.\end{cases}

Denote the function

hα,νG(x)=f0[x,μα]+f[x,νG;gα]2(l2[x,μα]+l4[x,νG;gα]).h_{\alpha,\nu_{G}}(x)=-\frac{f_{0}[x,\mu_{\alpha}]+f[x,\nu_{G};g_{\alpha}]}{2(l_{2}[x,\mu_{\alpha}]+l_{4}[x,\nu_{G};g_{\alpha}])}.

By elementary estimates we can show

|hα,νG(x)hα,νG(y)|C0|xy|,\displaystyle|h_{\alpha,\nu_{G}}(x)-h_{\alpha,\nu_{G}}(y)|\leq C_{0}|x-y|,

where C0C_{0} does not depend on (νG,α)(\nu_{G},\alpha). We have

SανG(x,q)\displaystyle S_{\alpha}^{\nu_{G}}(x,q) =argminu(u22qhα,νG(x)u)\displaystyle=\arg\min_{u}(u^{2}-2qh_{\alpha,\nu_{G}}(x)u)
=Θ(qhα,νG(x)).\displaystyle=\Theta({qh_{\alpha,\nu_{G}}(x)}).

It is clear that SανG(x,q)S_{\alpha}^{\nu_{G}}(x,q) is a continuous function of (x,q)(x,q). For (xi,qi)×(x_{i},q_{i})\in\mathbb{R}\times{\mathcal{I}}, i=1,2i=1,2,

|SανG(x1,q1)SανG(x2,q2)|\displaystyle|S_{\alpha}^{\nu_{G}}(x_{1},q_{1})-S_{\alpha}^{\nu_{G}}(x_{2},q_{2})|
Lip(Θ)|q1hα,νG(x1)q2hα,νG(x2)|\displaystyle\leq{\rm Lip}(\Theta)|q_{1}h_{\alpha,\nu_{G}}(x_{1})-q_{2}h_{\alpha,\nu_{G}}(x_{2})|
Lip(Θ)(|q1q2|supx|hα,νG(x)|+C0|x1x2||q2|).\displaystyle\leq{\rm Lip}(\Theta)\Big{(}|q_{1}-q_{2}|\sup_{x}|h_{\alpha,\nu_{G}}(x)|+C_{0}|x_{1}-x_{2}||q_{2}|\Big{)}.

In fact, the Lipschitz constant Lip(Θ)=1{\rm Lip}(\Theta)=1. Note that there exists a fixed constant CC such that |hα,νG(x)|C|h_{\alpha,\nu_{G}}(x)|\leq C for all α,νG\alpha,\nu_{G}. This proves the proposition. ∎

If (H1)–(H3) and (H5) hold for Example 3.13, they further imply (H4) and (H6) so that the best response is Lipschitz continuous in xx by Lemma 3.3 and Proposition 3.14.

4. Performance Analysis

In the MFG case it is shown [25, 8] that the joint strategy {uio(t)=φi(t,xi(t)|μ),1iN}\{u^{o}_{i}(t)=\varphi_{i}(t,x_{i}(t)|\mu_{\cdot}),1\leq i\leq N\} yields an ϵ\epsilon-Nash equilibrium, i.e. for all ϵ>0\epsilon>0, there exists N(ϵ)N(\epsilon) such that for all NN(ϵ)N\geq N(\epsilon)

(4.1) JiN(ui,ui)ϵinfui𝒰iJiN(ui,ui)JiN(ui,ui).J_{i}^{N}({u}_{i}^{\circ},{u}_{-i}^{\circ})-\epsilon\leq\inf_{u_{i}\in\mathcal{U}_{i}}J_{i}^{N}(u_{i},{u}_{-i}^{\circ})\leq J_{i}^{N}({u}_{i}^{\circ},{u}_{-i}^{\circ}).

This form of approximate Nash equilibrium is a principal result of the MFG analyses in the sequence [25, 8, 40] and in many other studies. The importance of (4.1) is that it states that the cost function of any agent in a finite population can be reduced by at most ϵ\epsilon if it changes unilaterally from the infinite population MFG feedback law while all other agents remain with the infinite population based control strategies. The main result of this section is that the same property holds for GMFG systems.

Throughout this section, let μG()\mu_{G}(\cdot) be solved from the GMFG equations (3.2) and (3.15).

4.1. The ϵ\epsilon-Nash Equilibrium

The analysis of GMFG systems as limits of finite objects necessarily involves the consideration of graph limits and double limits in population and graph order. A corresponding set of assumptions is given below.

(H9) MkM_{k}\to\infty and min1lMk|𝒞l|\min_{1\leq l\leq M_{k}}|\mathcal{C}_{l}|\to\infty as kk\to\infty.

(H10) All agents have i.i.d. initial states with distribution μ0x\mu_{0}^{x} and E|xi(0)|C0E|x_{i}(0)|\leq C_{0}.

Remark 4.1.

(H10) is a simplifying assumption to keep further notation light. It may be generalized to α\alpha dependent initial distributions.

(H11) The sequence {Gk;1k<}\{G_{k};1\leq k<\infty\} and the graphon limit satisfy

limkmaxij=1Mk|1Mkg𝒞i𝒞jkβIjgIi,β𝑑β|=0,\lim_{k\to\infty}\max_{i}\sum_{j=1}^{M_{k}}\Big{|}\frac{1}{M_{k}}g_{\mathcal{C}_{i}\mathcal{C}_{j}}^{k}-\int_{\beta\in I_{j}}g_{I_{i}^{*},\beta}d\beta\Big{|}=0,

where IiI^{*}_{i} is the midpoint of the subinterval Ii{I1,,IMk}I_{i}\in\{I_{1},\ldots,I_{M_{k}}\} of length 1/Mk1/{M_{k}}.

Remark 4.2.

Assumption (H11) specifies the nature of the approximation error between gkg^{k} for the finite graph and the graphon function gg.

The next proposition shows that under (H5) and (H11), the limit gg is well determined.

Proposition 4.3.

For the given sequence {gk,k1}\{g^{k},k\geq 1\} under (H9), if there exists a graphon gg satisfying (H5) and (H11), then it is unique.

Proof.

Assume there is another graphon g^\hat{g} satisfying (H5) and (H11). Fix any ϵ>0\epsilon>0 and any 𝒮×𝒯[0,1]×[0,1]\mathcal{S}\times\mathcal{T}\subset[0,1]\times[0,1]. By Lemma A.2, there exists a sufficiently large k0k_{0} (depending on ϵ\epsilon, 𝒮{\mathcal{S}} and 𝒯{\mathcal{T}}), such that for both gg and g^\hat{g} we have

|𝒮×𝒯(gk0g)𝑑x𝑑y|ϵ,|𝒮×𝒯(gk0g^)𝑑x𝑑y|ϵ.\displaystyle\Big{|}\int_{\mathcal{S}\times\mathcal{T}}(g^{k^{0}}-g)dxdy\Big{|}\leq\epsilon,\qquad\Big{|}\int_{\mathcal{S}\times\mathcal{T}}(g^{k^{0}}-\hat{g})dxdy\Big{|}\leq\epsilon.

Hence

|𝒮×𝒯(gg^)𝑑x𝑑y|2ϵ.\displaystyle\Big{|}\int_{\mathcal{S}\times\mathcal{T}}(g-\hat{g})dxdy\Big{|}\leq 2\epsilon.

Since 𝒮×𝒯\mathcal{S}\times\mathcal{T} is arbitrary, we have gg^2ϵ.\|g-\hat{g}\|_{\Box}\leq 2\epsilon. Since ϵ\epsilon is arbitrary, we have gg^=0.\|g-\hat{g}\|_{\Box}=0. But the cut norm is a norm, so we have g=g^g=\hat{g}. ∎

For the ϵ\epsilon-Nash equilibrium analysis, we consider a sequence of games each defined on a finite graph GkG_{k}. Recall that there is a total of N=l=1Mk|𝒞l|N=\sum_{l=1}^{M_{k}}|\mathcal{C}_{l}| agents.

Suppose the cluster 𝒞(i)\mathcal{C}(i) of agent 𝒜i{\mathcal{A}}_{i} corresponds to the subinterval I(i){I1,,IMk}I(i)\in\{I_{1},\ldots,I_{M_{k}}\}. The agent 𝒜i{\mathcal{A}}_{i} takes the midpoint I(i)I^{*}(i) of the subinterval I(i)I(i) and uses the GMFG equations to determine its control law

(4.2) u^i=φ(t,xi|μG();gI(i)),1iN,\displaystyle\hat{u}_{i}=\varphi(t,x_{i}|\mu_{G}(\cdot);g_{I^{*}(i)}),\quad 1\leq i\leq N,

which we simply write as φ(t,xi,gI(i))\varphi(t,x_{i},g_{I^{*}(i)}). Denote the resulting state process by x^i\hat{x}_{i}, 1iN1\leq i\leq N. Recall that

f0(xiN,uiN,𝒞(i))=1|𝒞(i)|j𝒞(i)f(xiN,uiN,xjN),\displaystyle f_{0}(x_{i}^{N},u_{i}^{N},{\mathcal{C}}(i))=\frac{1}{|\mathcal{C}(i)|}\sum_{j\in\mathcal{C}(i)}f(x_{i}^{N},u_{i}^{N},x_{j}^{N}),
fGk(xiN,uiN,g𝒞(i)k)=1Mkl=1Mkg𝒞(i)𝒞lk1|𝒞l|j𝒞lf(xiN,uiN,xjN),\displaystyle f_{G_{k}}(x_{i}^{N},u_{i}^{N},g_{\mathcal{C}(i)}^{k})=\frac{1}{M_{k}}\sum_{l=1}^{M_{k}}g^{k}_{\mathcal{C}(i)\mathcal{C}_{l}}\frac{1}{|\mathcal{C}_{l}|}\sum_{j\in\mathcal{C}_{l}}f(x_{i}^{N},u_{i}^{N},x_{j}^{N}),

where the superscript NN is added to indicate the population size. The closed-loop system of NN agents on the finite graph GkG_{k} under the set of strategies (4.2) is given by

System A:dx^iN=\displaystyle\mbox{\it System A:}\quad d\hat{x}_{i}^{N}= f0(x^iN,φ(t,x^iN,gI(i)),𝒞(i))dt\displaystyle f_{0}(\hat{x}_{i}^{N},\varphi(t,\hat{x}_{i}^{N},g_{I^{*}(i)}),{\mathcal{C}}(i))dt
(4.3) +fGk(x^iN,φ(t,x^iN,gI(i)),g𝒞(i)k)dt+σdwi,\displaystyle+f_{G_{k}}(\hat{x}_{i}^{N},\varphi(t,\hat{x}_{i}^{N},g_{I^{*}(i)}),g_{\mathcal{C}(i)}^{k})dt+\sigma dw_{i},

where 1iN1\leq i\leq N and x^iN(0)=xiN(0)\hat{x}_{i}^{N}(0)=x_{i}^{N}(0). Note that g𝒞(i)kg_{\mathcal{C}(i)}^{k} appears in fGkf_{G_{k}} as determined by the finite population system dynamics. We state the following main result.

Theorem 4.4.

(ϵ\epsilon-Nash equilibrium) Assume (H1)–(H11) hold. Then when the strategies (4.2) determined by the GMFG equations (3.2) and (3.15) are applied to a sequence of finite graph systems {Gk;1k<}\{G_{k};1\leq k<\infty\}, the ϵ\epsilon-Nash equilibrium property holds where ϵ0\epsilon\to 0 as kk\to\infty, and where the unilateral agent 𝒜i{\mathcal{A}}_{i} uses a centralized Lipschitz feedback strategy ψ(t,xi,xi)\psi(t,x_{i},x_{-i}), where xix_{-i} denotes the set of states of all other agents.

We first explain the basic idea for the demonstration of the ϵ\epsilon-Nash equilibrium property. Suppose all other players, except agent 𝒜ι{\mathcal{A}}_{\iota}, employ the control strategies based on the GMFG equation system. When 𝒜ι{\mathcal{A}}_{\iota} employs a different strategy, the resulting change in its performance can be measured using a limiting stochastic control problem where both the system dynamics and the cost are subject to small perturbation due to the mean field approximation of the effects of all other agents. The proof is technical and preceded by some lemmas.

4.2. Proof of Theorem 4.4

Suppose xιNx_{\iota}^{N} is determined from a general feedback control law uιNu_{\iota}^{N} instead of the GMFG best response. With the exception of agent 𝒜ι{\mathcal{A}}_{\iota} with its unilateral strategy, all other agents 𝒜j{\mathcal{A}}_{j}, jιj\neq{\iota}, still have strategies determined by (4.2). We introduce the system:

(4.4) System B:{dxιN=f0(xιN,uιN,𝒞(ι))dt+fGk(xιN,uιN,g𝒞(ι)k)dt+σdwι,dxjN=f0(xjN,φ(t,xjN,gI(j)),𝒞(j))dt+fGk(xjN,φ(t,xjN,gI(j)),g𝒞(j)k)dt+σdwj,jι,1jN.\displaystyle\mbox{\it System B:}\quad\begin{cases}dx_{\iota}^{N}=f_{0}(x_{\iota}^{N},u_{\iota}^{N},\mathcal{C}(\iota))dt+f_{G_{k}}(x_{\iota}^{N},u_{\iota}^{N},g_{\mathcal{C}(\iota)}^{k})dt+\sigma dw_{\iota},\\ dx_{j}^{N}=f_{0}(x_{j}^{N},\varphi(t,x_{j}^{N},g_{I^{*}(j)}),\mathcal{C}(j))dt\\ \qquad\qquad+f_{G_{k}}(x_{j}^{N},\varphi(t,x_{j}^{N},g_{I^{*}(j)}),g_{\mathcal{C}(j)}^{k})dt+\sigma dw_{j},\\ \qquad\qquad j\neq{\iota},\quad 1\leq j\leq N.\end{cases}

We note that xjNx_{j}^{N} is affected by the unilateral choice of strategy by 𝒜ι{\mathcal{A}}_{\iota} due to the coupling in f0f_{0} and fGkf_{G_{k}}. For this reason, xjNx_{j}^{N} differs from x^jN\hat{x}_{j}^{N} in (4.3) although the control law of 𝒜j{\mathcal{A}}_{j}, jιj\neq\iota, remains the same. The central task is to estimate by how much 𝒜ι{\mathcal{A}}_{\iota} can reduce its cost.

To facilitate the performance estimate in System BB, we introduce two auxiliary systems below. Consider

System C:dyiN=\displaystyle\mbox{\it System C:}\quad dy_{i}^{N}=\ f0(yiN,φ(t,yiN,gI(i)),z)myiN(dz)𝑑t\displaystyle\int_{\mathbb{R}}f_{0}(y_{i}^{N},\varphi(t,y_{i}^{N},g_{I^{*}(i)}),z)m_{y_{i}^{N}}(dz)dt
+1Mkl=1Mkg𝒞(i)𝒞lk1|𝒞l|j𝒞lf(yiN,φ(t,yiN,gI(i)),z)myjN(dz)𝑑t\displaystyle+\frac{1}{M_{k}}\sum_{l=1}^{M_{k}}g^{k}_{\mathcal{C}(i)\mathcal{C}_{l}}\frac{1}{|\mathcal{C}_{l}|}\sum_{j\in\mathcal{C}_{l}}\int_{\mathbb{R}}f(y_{i}^{N},\varphi(t,y_{i}^{N},g_{I^{*}(i)}),z)m_{y_{j}^{N}}(dz)dt
+σdwi\displaystyle+\sigma dw_{i}
=\displaystyle=\ f0(yiN,φ(t,yiN,gI(i)),z)myiN(dz)𝑑t\displaystyle\int_{\mathbb{R}}f_{0}(y_{i}^{N},\varphi(t,y_{i}^{N},g_{I^{*}(i)}),z)m_{y_{i}^{N}}(dz)dt
+1Mkl=1Mkg𝒞(i)𝒞lkf(yiN,φ(t,yiN,gI(i)),z)mlN(t,dz)𝑑t\displaystyle+\frac{1}{M_{k}}\sum_{l=1}^{M_{k}}g^{k}_{\mathcal{C}(i)\mathcal{C}_{l}}\int f(y_{i}^{N},\varphi(t,y_{i}^{N},g_{I^{*}(i)}),z)m_{l}^{N}(t,dz)dt
(4.5) +σdwi,\displaystyle+\sigma dw_{i},

where 1iN1\leq i\leq N and yiN(0)=xiN(0)y_{i}^{N}(0)=x_{i}^{N}(0), and myjN(t)m_{y_{j}^{N}(t)} denotes the law of yjN(t)y_{j}^{N}(t). Each Brownian motion wiw_{i} is the same as in (4.3). The second equality holds since all processes in cluster 𝒞l\mathcal{C}_{l} have the same distribution denoted by mlN(t,dz)m_{l}^{N}(t,dz) at time tt. It is clear that the processes y1N,,yNNy_{1}^{N},\ldots,y_{N}^{N} are independent, and {yjN,j𝒞l}\{y_{j}^{N},j\in\mathcal{C}_{l}\} are i.i.d. for any given ll.

Next we introduce

(4.6) System D:dyi(t)\displaystyle\mbox{\it System D:}\quad dy_{i}^{\infty}(t) =f~[yi(t),φ(t,yi(t),gI(i)),μG(t);gI(i)]dt+σdwi(t),\displaystyle=\widetilde{f}[y_{i}^{\infty}(t),\varphi(t,y_{i}^{\infty}(t),g_{I^{*}(i)}),\mu_{G}(t);g_{I^{*}(i)\ }]dt+\sigma dw_{i}(t),

where 1iN1\leq i\leq N and yi(0)=xiN(0)y_{i}^{\infty}(0)=x_{i}^{N}(0). Here wiw_{i} is the same as in (4.3). The process yiy_{i}^{\infty} is generated by the closed-loop dynamics for an agent at the node I(i)I^{*}(i) associated with the cluster 𝒞(i)\mathcal{C}(i) using the GMFG based control law (4.2) while situated in an infinite population represented by the ensemble μG()\mu_{G}(\cdot) of local mean fields. We view (4.6) as an instance of the generic equation (3.10) under the control law (4.2). By Theorem 3.12, yi(t)y^{\infty}_{i}(t) has the law μI(i)(t)\mu_{I^{*}(i)}(t). Note that if j𝒞(i)j\in\mathcal{C}(i), yiy_{i}^{\infty} and yjy_{j}^{\infty} are two processes of the same distribution.

We shall denote the AA to CC system deviation by ϵ1,N\epsilon_{1,N}, the CC to DD deviation by ϵ2,N\epsilon_{2,N} and the (non-unilateral agent) BB to DD deviation by ϵ3,N\epsilon_{3,N}. Specifically, we set

ϵ1,N=supiN,tE|x^iN(t)yiN(t)|,ϵ2,N=supiN,tE|yiN(t)yi(t)|,\displaystyle\epsilon_{1,N}=\sup_{i\leq N,t}E|\hat{x}_{i}^{N}(t)-y_{i}^{N}(t)|,\qquad\epsilon_{2,N}=\sup_{i\leq N,t}E|y_{i}^{N}(t)-y_{i}^{\infty}(t)|,
ϵ3,N=supuιN,t,ιjNE|xjN(t)yj(t)|,\displaystyle\epsilon_{3,N}=\sup_{u_{\iota}^{N},t,\iota\neq j\leq N}E|x_{j}^{N}(t)-y_{j}^{\infty}(t)|,

where xjNx_{j}^{N} is given by (4.4).

Lemma 4.5.

The SDE system (4.5) has a unique solution (y1N,,yNN)(y_{1}^{N},\ldots,y_{N}^{N}).

Proof.

The proof is similar to [25, Theorem 6]. ∎

Lemma 4.6.

ϵ1,N0\epsilon_{1,N}\to 0 as NN\to\infty (due to kk\to\infty).

Proof.

We write

(4.7) x^iN(t)yiN(t)=\displaystyle\hat{x}_{i}^{N}(t)-y_{i}^{N}(t)= 0t1|𝒞(i)|j𝒞(i)ξij0(s)ds\displaystyle\int_{0}^{t}\frac{1}{|\mathcal{C}(i)|}\sum_{j\in\mathcal{C}(i)}\xi_{ij}^{0}(s)ds
+0t1Mkl=1Mkg𝒞(i)𝒞lk1|𝒞l|j𝒞lξij(s)ds,\displaystyle+\int_{0}^{t}\frac{1}{M_{k}}\sum_{l=1}^{M_{k}}g_{\mathcal{C}(i)\mathcal{C}_{l}}^{k}\frac{1}{|\mathcal{C}_{l}|}\sum_{j\in\mathcal{C}_{l}}\xi_{ij}(s)ds,

where

ξij0(s)=f0(x^iN,φ(s,x^iN,gI(i)),x^jN)f0(yiN,φ(s,yiN,gI(i)),z)myjN(s)(dz),\displaystyle\xi_{ij}^{0}(s)=f_{0}(\hat{x}_{i}^{N},\varphi(s,\hat{x}_{i}^{N},g_{I^{*}(i)}),\hat{x}_{j}^{N})-\int_{\mathbb{R}}f_{0}(y_{i}^{N},\varphi(s,y_{i}^{N},g_{I^{*}(i)}),z)m_{y_{j}^{N}(s)}(dz),
ξij(s)=f(x^iN,φ(s,x^iN,gI(i)),x^jN)f(yiN,φ(s,yiN,gI(i)),z)myjN(s)(dz).\displaystyle\xi_{ij}(s)=f(\hat{x}_{i}^{N},\varphi(s,\hat{x}_{i}^{N},g_{I^{*}(i)}),\hat{x}_{j}^{N})-\int_{\mathbb{R}}f(y_{i}^{N},\varphi(s,y_{i}^{N},g_{I^{*}(i)}),z)m_{y_{j}^{N}(s)}(dz).

We check the second line of (4.7) first. Write

ξij(s)=\displaystyle\xi_{ij}(s)= f(x^iN,φ(s,x^iN,gI(i)),x^jN)f(yiN,φ(s,yiN,gI(i)),yjN)\displaystyle f(\hat{x}_{i}^{N},\varphi(s,\hat{x}_{i}^{N},g_{I^{*}(i)}),\hat{x}_{j}^{N})-f(y_{i}^{N},\varphi(s,y_{i}^{N},g_{I^{*}(i)}),y_{j}^{N})
+f(yiN,φ(s,yiN,gI(i)),yjN)f(yiN,φ(s,yiN,gI(i)),z)myjN(s)(dz).\displaystyle+f(y_{i}^{N},\varphi(s,y_{i}^{N},g_{I^{*}(i)}),y_{j}^{N})-\int_{\mathbb{R}}f(y_{i}^{N},\varphi(s,y_{i}^{N},g_{I^{*}(i)}),z)m_{y_{j}^{N}(s)}(dz).

Denote

ζij=f(yiN,φ(s,yiN,gI(i)),yjN)f(yiN,φ(s,yiN,gI(i)),z)myjN(s)(dz).\zeta_{ij}=f(y_{i}^{N},\varphi(s,y_{i}^{N},g_{I^{*}(i)}),y_{j}^{N})-\int_{\mathbb{R}}f(y_{i}^{N},\varphi(s,y_{i}^{N},g_{I^{*}(i)}),z)m_{y_{j}^{N}(s)}(dz).

By the Lipschitz conditions (H2), (H3) and the best response’s uniform Lipschitz continuity in xx by Lemma 3.7, we obtain

|1Mkl=1Mkg𝒞(i)𝒞lk1|𝒞l|j𝒞lξij(s)|\displaystyle\Big{|}\frac{1}{M_{k}}\sum_{l=1}^{M_{k}}g_{\mathcal{C}(i)\mathcal{C}_{l}}^{k}\frac{1}{|\mathcal{C}_{l}|}\sum_{j\in\mathcal{C}_{l}}\xi_{ij}(s)\Big{|}
\displaystyle\leq C|x^iNyiN|+CMkl=1Mkg𝒞(i)𝒞lk1|𝒞l|j𝒞l|x^jNyjN|\displaystyle C|\hat{x}_{i}^{N}-y_{i}^{N}|+\frac{C}{M_{k}}\sum_{l=1}^{M_{k}}g_{\mathcal{C}(i)\mathcal{C}_{l}}^{k}\frac{1}{|\mathcal{C}_{l}|}\sum_{j\in\mathcal{C}_{l}}|\hat{x}_{j}^{N}-y_{j}^{N}|
+|1Mkl=1Mkg𝒞(i)𝒞lk1|𝒞l|j𝒞lζij|.\displaystyle+\Big{|}\frac{1}{M_{k}}\sum_{l=1}^{M_{k}}g_{\mathcal{C}(i)\mathcal{C}_{l}}^{k}\frac{1}{|\mathcal{C}_{l}|}\sum_{j\in\mathcal{C}_{l}}\zeta_{ij}\Big{|}.

Then by independence of yiNy_{i}^{N}, 1iN1\leq i\leq N,

E|1Mkl=1Mkg𝒞(i)𝒞lk1|𝒞l|j𝒞lζij|2\displaystyle E\Big{|}\frac{1}{M_{k}}\sum_{l=1}^{M_{k}}g_{\mathcal{C}(i)\mathcal{C}_{l}}^{k}\frac{1}{|\mathcal{C}_{l}|}\sum_{j\in\mathcal{C}_{l}}\zeta_{ij}\Big{|}^{2} Cl=1Mkj𝒞l|g𝒞(i)𝒞lk|2Mk2|𝒞l|2\displaystyle\leq C\sum_{l=1}^{M_{k}}\sum_{j\in\mathcal{C}_{l}}\frac{|g_{\mathcal{C}(i)\mathcal{C}_{l}}^{k}|^{2}}{M_{k}^{2}|\mathcal{C}_{l}|^{2}}
CMkminl|𝒞l|.\displaystyle\leq\frac{C}{M_{k}\min_{l}|\mathcal{C}_{l}|}.

The estimate for 1|𝒞(i)|j𝒞(i)ξij0(s)\frac{1}{|\mathcal{C}(i)|}\sum_{j\in\mathcal{C}(i)}\xi_{ij}^{0}(s) can be obtained similarly. Now it follows from (4.7) that

E|x^iN(t)yiN(t)|C0tE|x^iN(s)yiN(s)|𝑑s\displaystyle E|\hat{x}_{i}^{N}(t)-y_{i}^{N}(t)|\leq C\int_{0}^{t}E|\hat{x}_{i}^{N}(s)-y_{i}^{N}(s)|ds
+CMkl=1Mkg𝒞(i)𝒞lk|𝒞l|j𝒞l0tE|x^jN(s)yjN(s)|𝑑s\displaystyle+\frac{C}{M_{k}}\sum_{l=1}^{M_{k}}\frac{g_{\mathcal{C}(i)\mathcal{C}_{l}}^{k}}{|\mathcal{C}_{l}|}\sum_{j\in\mathcal{C}_{l}}\int_{0}^{t}E|\hat{x}_{j}^{N}(s)-y_{j}^{N}(s)|ds
+C|𝒞(i)|j𝒞(i)0tE|x^jN(s)yjN(s)|𝑑s+C1Mkminl|𝒞l|+C|𝒞(i)|\displaystyle+\frac{C}{|\mathcal{C}(i)|}\sum_{j\in\mathcal{C}(i)}\int_{0}^{t}E|\hat{x}_{j}^{N}(s)-y_{j}^{N}(s)|ds+\frac{C_{1}}{\sqrt{M_{k}\min_{l}|\mathcal{C}_{l}|}}+\frac{C}{\sqrt{|\mathcal{C}(i)|}}
C20tΔN(s)𝑑s+C3minl|𝒞l|,\displaystyle\leq C_{2}\int_{0}^{t}\Delta^{N}(s)ds+\frac{C_{3}}{\sqrt{\min_{l}|\mathcal{C}_{l}|}},

where ΔN(t)=max1iNE|x^iN(t)yiN(t)|.\Delta^{N}(t)=\max_{1\leq i\leq N}E|\hat{x}_{i}^{N}(t)-y_{i}^{N}(t)|. The above further implies

ΔN(t)C20tΔN(s)𝑑s+C3minl|𝒞l|.\displaystyle\Delta^{N}(t)\leq C_{2}\int_{0}^{t}\Delta^{N}(s)ds+\frac{C_{3}}{\sqrt{\min_{l}|\mathcal{C}_{l}|}}.

The lemma follows from (H9) and Gronwall’s lemma. ∎

Lemma 4.7.

We have ϵ2,N0\epsilon_{2,N}\to 0 as NN\to\infty.

Proof.

For System DD and 1iN1\leq i\leq N, yi(t)y_{i}^{\infty}(t) has the law μI(i)(t)\mu_{I^{*}(i)}(t) and we write

(4.8) dyi=\displaystyle dy_{i}^{\infty}=\ f0(yi,φ(t,yi,gI(i)),z)μI(i)(t,dz)𝑑t+σdwi\displaystyle\int_{\mathbb{R}}f_{0}(y_{i}^{\infty},\varphi(t,y_{i}^{\infty},g_{I^{*}(i)}),z)\mu_{I^{*}(i)}(t,dz)dt+\sigma dw_{i}
+01f(yi,φ(t,yi,gI(i)),z)g(I(i),β)μβ(t,dz)𝑑β𝑑t.\displaystyle+\int_{0}^{1}\int_{\mathbb{R}}f(y_{i}^{\infty},\varphi(t,y_{i}^{\infty},g_{I^{*}(i)}),z)g(I^{*}(i),\beta)\mu_{\beta}(t,dz)d\beta\ dt.

Set

01f(yi,φ(t,yi,gI(i)),z)g(I(i),β)μβ(t,dz)𝑑β\displaystyle\int_{0}^{1}\int_{\mathbb{R}}f(y_{i}^{\infty},\varphi(t,y_{i}^{\infty},g_{I^{*}(i)}),z)g(I^{*}(i),\beta)\mu_{\beta}(t,dz)d\beta
=\displaystyle=\ l=1MkβIlf(yi,φ(t,yi,gI(i)),z)g(I(i),β)μβ(t,dz)𝑑β\displaystyle\sum_{l=1}^{M_{k}}\int_{\beta\in I_{l}}\int_{\mathbb{R}}f(y_{i}^{\infty},\varphi(t,y_{i}^{\infty},g_{I^{*}(i)}),z)g(I^{*}(i),\beta)\mu_{\beta}(t,dz)d\beta
\displaystyle\eqqcolon ξki+ζki,\displaystyle\ \xi_{k}^{i}+\zeta_{k}^{i},

where

ξki=l=1MkβIlg(I(i),β)𝑑βf(yi,φ(t,yi,gI(i)),z)μIl(t,dz),\displaystyle\xi_{k}^{i}=\sum_{l=1}^{M_{k}}\int_{\beta\in I_{l}}g(I^{*}(i),\beta)d\beta\int_{\mathbb{R}}f(y_{i}^{\infty},\varphi(t,y_{i}^{\infty},g_{I^{*}(i)}),z)\mu_{I_{l}^{*}}(t,dz),
ζki=l=1Mkζkli,\displaystyle\zeta_{k}^{i}=\sum_{l=1}^{M_{k}}\zeta^{i}_{kl},
(4.9) ζkliβIlf(yi,φ(t,yi,gI(i)),z)g(I(i),β)[μβ(t,dz)μIl(t,dz)]𝑑β.\displaystyle\zeta_{kl}^{i}\coloneqq\int_{\beta\in I_{l}}\int_{\mathbb{R}}f(y_{i}^{\infty},\varphi(t,y_{i}^{\infty},g_{I^{*}(i)}),z)g(I^{*}(i),\beta)[\mu_{\beta}(t,dz)-\mu_{I_{l}^{*}}(t,dz)]d\beta.

We rewrite

ξki=\displaystyle\xi_{k}^{i}= l=1Mkg𝒞(i)𝒞lkMkf(yi,φ(t,yi,gI(i)),z)μIl(t,dz)\displaystyle\sum_{l=1}^{M_{k}}\frac{g^{k}_{\mathcal{C}(i)\mathcal{C}_{l}}}{M_{k}}\int_{\mathbb{R}}f(y_{i}^{\infty},\varphi(t,y_{i}^{\infty},g_{I^{*}(i)}),z)\mu_{I_{l}^{*}}(t,dz)
+l=1Mk[βIlg(I(i),β)𝑑βg𝒞(i)𝒞lkMk]f(yi,φ(t,yi,gI(i)),z)μIl(t,dz)\displaystyle+\sum_{l=1}^{M_{k}}\left[\int_{\beta\in I_{l}}g(I^{*}(i),\beta)d\beta-\frac{g^{k}_{\mathcal{C}(i)\mathcal{C}_{l}}}{M_{k}}\right]\int_{\mathbb{R}}f(y_{i}^{\infty},\varphi(t,y_{i}^{\infty},g_{I^{*}(i)}),z)\mu_{I_{l}^{*}}(t,dz)
\displaystyle\eqqcolon\ ξk,1i+ξk,2i.\displaystyle\xi^{i}_{k,1}+\xi^{i}_{k,2}.

By (H11) and boundedness of ff, we have limksupt,ωmax1iN|ξk,2i|=0\lim_{k\to\infty}\sup_{t,\omega}\max_{1\leq i\leq N}|\xi^{i}_{k,2}|=0 so that

(4.10) limkmaxi0TE|ξk,2i(t)|𝑑t=0.\displaystyle\lim_{k\to\infty}\max_{i}\int_{0}^{T}E|\xi^{i}_{k,2}(t)|dt=0.

Now (4.8) may be rewritten in the form

dyi=\displaystyle dy_{i}^{\infty}=\ f0(yi,φ(t,yi,gI(i)),z)μI(i)(t,dz)𝑑t+σdwi\displaystyle\int_{\mathbb{R}}f_{0}(y_{i}^{\infty},\varphi(t,y_{i}^{\infty},g_{I^{*}(i)}),z)\mu_{I^{*}(i)}(t,dz)dt+\sigma dw_{i}
+(ξk,1i+ξk,2i+ζki)dt.\displaystyle+(\xi^{i}_{k,1}+\xi^{i}_{k,2}+\zeta^{i}_{k})dt.

In view of (4.5), we have

yi(t)yiN(t)\displaystyle y_{i}^{\infty}(t)-y_{i}^{N}(t)
=\displaystyle= 0t[f0(yi,φ(s,yi,gI(i)),z)μI(i)(s,dz)f0(yiN,φ(s,yiN,gI(i)),z)myiN(s)(dz)]𝑑s\displaystyle\int_{0}^{t}\int_{\mathbb{R}}[f_{0}(y_{i}^{\infty},\varphi(s,y_{i}^{\infty},g_{I^{*}(i)}),z)\mu_{I^{*}(i)}(s,dz)-f_{0}(y_{i}^{N},\varphi(s,y_{i}^{N},g_{I^{*}(i)}),z)m_{y^{N}_{i}(s)}(dz)]ds
+1Mkl=1Mkg𝒞(i)𝒞lk0tf(yi,φ(s,yi,gI(i)),z)μIl(s,dz)𝑑s\displaystyle+\frac{1}{M_{k}}\sum_{l=1}^{M_{k}}g^{k}_{\mathcal{C}(i)\mathcal{C}_{l}}\int_{0}^{t}\int_{\mathbb{R}}f(y_{i}^{\infty},\varphi(s,y_{i}^{\infty},g_{I^{*}(i)}),z)\mu_{I_{l}^{*}}(s,dz)ds
1Mkl=1Mkg𝒞(i)𝒞lk0tf(yiN,φ(s,yiN,gI(i)),z)mlN(s,dz)𝑑s\displaystyle-\frac{1}{M_{k}}\sum_{l=1}^{M_{k}}g^{k}_{\mathcal{C}(i)\mathcal{C}_{l}}\int_{0}^{t}\int_{\mathbb{R}}f(y_{i}^{N},\varphi(s,y_{i}^{N},g_{I^{*}(i)}),z)m_{l}^{N}(s,dz)ds
+0t(ξk,2i+ζki)𝑑s.\displaystyle+\int_{0}^{t}(\xi^{i}_{k,2}+\zeta^{i}_{k})ds.

Denote

Δil(s)=\displaystyle\Delta_{il}(s)= |f(yi,φ(s,yi,gI(i)),z)μIl(s,dz)\displaystyle\Big{|}\int_{\mathbb{R}}f(y_{i}^{\infty},\varphi(s,y_{i}^{\infty},g_{I^{*}(i)}),z)\mu_{I_{l}^{*}}(s,dz)
f(yiN,φ(s,yiN,gI(i)),z)mlN(s,dz)|.\displaystyle-\int_{\mathbb{R}}f(y_{i}^{N},\varphi(s,y_{i}^{N},g_{I^{*}(i)}),z)m_{l}^{N}(s,dz)\Big{|}.

It follows that

Δil(s)\displaystyle\Delta_{il}(s)\leq |f(yi,φ(s,yi,gI(i)),z)μIl(s,dz)\displaystyle\Big{|}\int_{\mathbb{R}}f(y_{i}^{\infty},\varphi(s,y_{i}^{\infty},g_{I^{*}(i)}),z)\mu_{I_{l}^{*}}(s,dz)
f(yiN,φ(s,yiN,gI(i)),z)μIl(s,dz)|\displaystyle-\int f(y_{i}^{N},\varphi(s,y_{i}^{N},g_{I^{*}(i)}),z)\mu_{I_{l}^{*}}(s,dz)\Big{|}
+|f(yiN,φ(s,yiN,gI(i)),z)μIl(s,dz)\displaystyle+\Big{|}\int_{\mathbb{R}}f(y_{i}^{N},\varphi(s,y_{i}^{N},g_{I^{*}(i)}),z)\mu_{I_{l}^{*}}(s,dz)
f(yiN,φ(s,yiN,gI(i)),z)mlN(s,dz)|\displaystyle-\int_{\mathbb{R}}f(y_{i}^{N},\varphi(s,y_{i}^{N},g_{I^{*}(i)}),z)m_{l}^{N}(s,dz)\Big{|}
\displaystyle\eqqcolon Δil1(s)+Δil2(s).\displaystyle\ \Delta_{il1}(s)+\Delta_{il2}(s).

By the Lipschitz condition (H2), for any fixed yy\in\mathbb{R}, we have

|f(y,φ(s,y,gI(i)),z)μIl(s,dz)f(y,φ(s,y,gI(i)),z)mlN(s,dz)|\displaystyle\Big{|}\int_{\mathbb{R}}f(y,\varphi(s,y,g_{I^{*}(i)}),z)\mu_{I_{l}^{*}}(s,dz)-\int_{\mathbb{R}}f(y,\varphi(s,y,g_{I^{*}(i)}),z)m_{l}^{N}(s,dz)\Big{|}
=\displaystyle= |Ef(y,φ(s,y,gI(i)),yj)Ef(y,φ(s,y,gI(i)),yjN)|\displaystyle|Ef(y,\varphi(s,y,g_{I^{*}(i)}),y_{j}^{\infty})-Ef(y,\varphi(s,y,g_{I^{*}(i)}),y_{j}^{N})|
\displaystyle\leq CE|yj(s)yjN(s)|,\displaystyle CE|y_{j}^{\infty}(s)-y_{j}^{N}(s)|,

where j𝒞lj\in\mathcal{C}_{l} and we have used the fact that yi(t)y_{i}^{\infty}(t) in (4.8) has the law μI(i)(t)\mu_{I^{*}(i)}(t) and that yjN(t)y_{j}^{N}(t) has the law mlN(t)m_{l}^{N}(t). Consequently, we have for j𝒞lj\in\mathcal{C}_{l}, with probability one,

(4.11) Δil2(s)CE|yj(s)yjN(s)|.\displaystyle\Delta_{il2}(s)\leq CE|y_{j}^{\infty}(s)-y_{j}^{N}(s)|.

We estimate Δkl1\Delta_{kl1} using the Lipschitz property of ff and φI(i)\varphi_{I^{*}(i)}. Now it follows that

EΔil(s)CE|yi(s)yiN(s)|+CE|yj(s)yjN(s)|,j𝒞l.\displaystyle E\Delta_{il}(s)\leq CE|y_{i}^{\infty}(s)-y_{i}^{N}(s)|+CE|y_{j}^{\infty}(s)-y_{j}^{N}(s)|,\quad j\in\mathcal{C}_{l}.

We similarly estimate the difference term involving f0f_{0}. Therefore,

E|yi(t)yiN(t)|\displaystyle E|y_{i}^{\infty}(t)-y_{i}^{N}(t)|\leq\ C0tE|yiyiN|𝑑s+0tE(|ξk,2i|+|ζki|)𝑑s\displaystyle C\int_{0}^{t}E|y_{i}^{\infty}-y_{i}^{N}|ds+\int_{0}^{t}E(|\xi^{i}_{k,2}|+|\zeta^{i}_{k}|)ds
+1Mkl=1Mkg𝒞(i)𝒞lk0tEΔil𝑑s\displaystyle+\frac{1}{M_{k}}\sum_{l=1}^{M_{k}}g^{k}_{\mathcal{C}(i)\mathcal{C}_{l}}\int_{0}^{t}E\Delta_{il}ds
\displaystyle\leq C10tmaxiE|yiyiN|ds+0tE(|ξk,2i|+|ζki|)𝑑s\displaystyle C_{1}\int_{0}^{t}\max_{i}E|y_{i}^{\infty}-y_{i}^{N}|ds+\int_{0}^{t}E(|\xi^{i}_{k,2}|+|\zeta^{i}_{k}|)ds
+CMkl=1Mkg𝒞(i)𝒞lk0tmaxjE|yjyjN|ds\displaystyle+\frac{C}{M_{k}}\sum_{l=1}^{M_{k}}g^{k}_{\mathcal{C}(i)\mathcal{C}_{l}}\int_{0}^{t}\max_{j}E|y_{j}^{\infty}-y_{j}^{N}|ds
\displaystyle\leq 2C20tmaxiE|yiyiN|ds+0tE(|ξk,2i|+|ζki|)𝑑s.\displaystyle 2C_{2}\int_{0}^{t}\max_{i}E|y_{i}^{\infty}-y_{i}^{N}|ds+\int_{0}^{t}E(|\xi^{i}_{k,2}|+|\zeta^{i}_{k}|)ds.

Consequently,

maxiE|yi(t)yiN(t)|\displaystyle\max_{i}E|y_{i}^{\infty}(t)-y_{i}^{N}(t)|\leq\ 2C20tmaxiE|yiyiN|ds+maxi0tE(|ξk,2i|+|ζki|)𝑑s.\displaystyle 2C_{2}\int_{0}^{t}\max_{i}E|y_{i}^{\infty}-y_{i}^{N}|ds+\max_{i}\int_{0}^{t}E(|\xi^{i}_{k,2}|+|\zeta^{i}_{k}|)ds.

By Gronwall’s lemma,

(4.12) sup0tTmaxiE|yi(t)yiN(t)|Cmaxi0TE(|ξk,2i|+|ζki|)𝑑s.\displaystyle\sup_{0\leq t\leq T}\max_{i}E|y_{i}^{\infty}(t)-y_{i}^{N}(t)|\leq C\max_{i}\int_{0}^{T}E(|\xi^{i}_{k,2}|+|\zeta^{i}_{k}|)ds.

To estimate (4.9), by (H2) we derive

ζkl,βi\displaystyle\zeta^{i}_{kl,\beta}\coloneqq |f(yi,φ(t,yi,gI(i)),z)[μβ(t,dz)μIl(t,dz)]|\displaystyle\Big{|}\int_{\mathbb{R}}f(y_{i}^{\infty},\varphi(t,y_{i}^{\infty},g_{I^{*}(i)}),z)[\mu_{\beta}(t,dz)-\mu_{I_{l}^{*}}(t,dz)]\Big{|}
=\displaystyle= |2[f(yi,φ(t,yi,gI(i)),z1)f(yi,φ(t,yi,gI(i)),z2)]γ^(dz1,dz2)|\displaystyle\Big{|}\int_{\mathbb{R}^{2}}[f(y_{i}^{\infty},\varphi(t,y_{i}^{\infty},g_{I^{*}(i)}),z_{1})-f(y_{i}^{\infty},\varphi(t,y_{i}^{\infty},g_{I^{*}(i)}),z_{2})]\widehat{\gamma}(dz_{1},dz_{2})\Big{|}
\displaystyle\leq C2|z1z2|γ^(dz1,dz2),\displaystyle C\int_{\mathbb{R}^{2}}|z_{1}-z_{2}|\widehat{\gamma}(dz_{1},dz_{2}),

where the probability measure γ^\widehat{\gamma} is any coupling of μβ(t)\mu_{\beta}(t) and μIl(t)\mu_{I_{l}^{*}}(t) and CC is the Lipschitz constant of ff. Since the coupling γ^\widehat{\gamma} is arbitrary, we have ζkl,βiCW1(μβ(t),μI(i)(t))\zeta^{i}_{kl,\beta}\leq CW_{1}(\mu_{\beta}(t),\mu_{I^{*}(i)}(t)). Denote δkμ=suplMksupβIl,tTW1(μβ(t),μIl(t)).\delta_{k}^{\mu}=\sup_{l\leq M_{k}}\sup_{\beta\in I_{l},t\leq T}W_{1}(\mu_{\beta}(t),\mu_{I_{l}^{*}}(t)). Then with probability one,

|ζkli(t)|Cδkμ/Mk\displaystyle|\zeta^{i}_{kl}(t)|\leq C\delta^{\mu}_{k}/M_{k}

in view of (4.9), and therefore maxi|ζki(t)|Cδkμ\max_{i}|\zeta^{i}_{k}(t)|\leq C\delta_{k}^{\mu}. Note that δkμ0\delta^{\mu}_{k}\to 0 as kk\to\infty by Lemma A.1. Recalling (4.10), the right hand side of (4.12) tends to 0 as kk\to\infty. This completes the proof. ∎

Lemma 4.8.

limNsupt,iNE|x^iNyi|=0\lim_{N\to\infty}\sup_{t,i\leq N}E|\hat{x}_{i}^{N}-y_{i}^{\infty}|=0.

Proof.

The lemma follows from Lemmas 4.6 and 4.7. ∎

Lemma 4.9.

limNϵ3,N=0.\lim_{N\to\infty}\epsilon_{3,N}=0.

Proof.

For (x^1N,,x^NN)(\hat{x}_{1}^{N},\ldots,\hat{x}_{N}^{N}) in System AA and (x1N,,xNN)(x_{1}^{N},\ldots,x_{N}^{N}) in System BB, we compare the SDEs of x^jN\hat{x}_{j}^{N} and xjNx_{j}^{N} and apply Gronwall’s lemma to obtain

supuιN,t,jι|xjNx^jN|Cminl|𝒞l|.\sup_{u_{\iota}^{N},t,j\neq\iota}|x_{j}^{N}-\hat{x}_{j}^{N}|\leq\frac{C}{{\min_{l}|\mathcal{C}_{l}|}}.

Next by Lemma 4.8, we obtain the desired estimate. ∎

Consider the limiting optimal control problem with dynamics and cost

(4.13) dxι=f~[xι,uι,μG;gI(ι)]dt+σdwι,\displaystyle dx_{\iota}^{\infty}=\widetilde{f}[x_{\iota}^{\infty},u_{\iota},\mu_{G};g_{I^{*}({\iota})}]dt+\sigma dw_{\iota},
(4.14) Jι=E0Tl~[xι,uι,μG;gI(ι)]𝑑t,\displaystyle J_{\iota}^{*}=E\int_{0}^{T}\widetilde{l}[x_{\iota}^{\infty},u_{\iota},\mu_{G};g_{I^{*}({\iota})}]dt,

where xι(0)=xιN(0)x_{\iota}^{\infty}(0)=x_{\iota}^{N}(0) and μG()\mu_{G}(\cdot) is given by the GMFG equation system.

To establish the ϵ\epsilon-Nash equilibrium property, the cost of agent 𝒜ι{\mathcal{A}}_{\iota} within the NN agents can be written using the mean field limit dynamics and cost, both involving μG()\mu_{G}(\cdot), up to a small error term that can be bounded uniformly with respect to uιNu_{\iota}^{N}, while 𝒜ι{\mathcal{A}}_{\iota} chooses its control uιNu_{\iota}^{N}. It can further have little improvement due to the best response property of φ(t,xι|μG();gI(ι))\varphi(t,x_{\iota}|\mu_{G}(\cdot);g_{I^{*}({\iota})}) within the mean field limit. We rewrite the first equation in (4.4) of System BB as

(4.15) dxιN=f~[xιN,uιN,μG;gI(ι)]dt+(δf0k(t)+δfk(t))dt+σdwι,\displaystyle dx_{\iota}^{N}=\widetilde{f}[x_{\iota}^{N},u_{\iota}^{N},\mu_{G};g_{I^{*}({\iota})}]dt+(\delta_{f_{0}}^{k}(t)+\delta_{f}^{k}(t))dt+\sigma dw_{\iota},

where δf0k=f0(xιN,uιN,𝒞(ι))f0[xιN,uιN,μI(ι)]\delta_{f_{0}}^{k}=f_{0}(x_{\iota}^{N},u_{\iota}^{N},{\mathcal{C}({\iota})})-f_{0}[x_{\iota}^{N},u_{\iota}^{N},\mu_{I^{*}({\iota})}] and δfk=fGk(xιN,uιN,g𝒞(ι)k)f[xιN,uιN,μG;gI(ι)]\delta_{f}^{k}=f_{G_{k}}(x_{\iota}^{N},u_{\iota}^{N},g_{\mathcal{C}({\iota})}^{k})-f[x_{\iota}^{N},u_{\iota}^{N},\mu_{G};g_{I^{*}({\iota})}]. Similarly the cost of 𝒜ι{\mathcal{A}}_{\iota} in System BB is written as

JιN(uιN)=E0T(l~[xιN,uιN,μG;gI(ι)]+δl0k(t)+δlk(t))𝑑t,\displaystyle J_{\iota}^{N}(u_{\iota}^{N})=E\int_{0}^{T}(\widetilde{l}[x_{\iota}^{N},u_{\iota}^{N},\mu_{G};g_{I^{*}({\iota})}]+\delta_{l_{0}}^{k}(t)+\delta_{l}^{k}(t))dt,

where we have δl0k=l0(xιN,uιN,𝒞(ι))l0[xιN,uιN,μI(ι)]\delta_{l_{0}}^{k}=l_{0}(x_{\iota}^{N},u_{\iota}^{N},{\mathcal{C}({\iota})})-l_{0}[x_{\iota}^{N},u_{\iota}^{N},\mu_{I^{*}({\iota})}] and δlk=lGk(xιN,uιN,g𝒞(ι)k)l[xιN,uιN,μG;gI(ι)]\delta_{l}^{k}=l_{G_{k}}(x_{\iota}^{N},u_{\iota}^{N},g_{\mathcal{C}({\iota})}^{k})-l[x_{\iota}^{N},u_{\iota}^{N},\mu_{G};g_{I^{*}({\iota})}]. Note that all other agents have applied the control laws φ(t,xjN,gI(j))\varphi(t,x_{j}^{N},g_{I^{*}(j)}), jιj\neq{\iota}. So we only indicate uιNu_{\iota}^{N} within JιNJ_{\iota}^{N}. It is clear that δf0k\delta_{f_{0}}^{k}, δfk\delta_{f}^{k}, δl0k\delta_{l_{0}}^{k}, and δlk\delta_{l}^{k} are all affected by the control law uιNu_{\iota}^{N}. Let \mathboldyt=(y1(t),,yN(t)){\mathbold y}^{\infty}_{t}=(y_{1}^{\infty}(t),\ldots,y_{N}^{\infty}(t)) for System DD. Our next step is to derive a uniform upper bounded for E|δfk|E|\delta_{f}^{k}| and E|δlk|E|\delta_{l}^{k}| with respect to uιNu_{\iota}^{N}.

Define the two random variables

Δfk(z,u,\mathboldyt)=1Mkl=1Mkg𝒞(ι)𝒞lk1|𝒞l|j𝒞lf(z,u,yj(t))f[z,u,μG(t);gI(ι)],\displaystyle\Delta_{f}^{k}(z,u,{\mathbold y}^{\infty}_{t})=\frac{1}{M_{k}}\sum_{l=1}^{M_{k}}g^{k}_{{\mathcal{C}}({\iota}){\mathcal{C}}_{l}}\frac{1}{|{\mathcal{C}}_{l}|}\sum_{j\in{\mathcal{C}}_{l}}f(z,u,y_{j}^{\infty}(t))-f[z,u,\mu_{G}(t);g_{I^{*}({\iota})}],
Δlk(z,u,\mathboldyt)=1Mkl=1Mkg𝒞(ι)𝒞lk1|𝒞l|j𝒞ll(z,u,yj(t))l[z,u,μG(t);gI(ι)],\displaystyle\Delta_{l}^{k}(z,u,{\mathbold y}^{\infty}_{t})=\frac{1}{M_{k}}\sum_{l=1}^{M_{k}}g^{k}_{{\mathcal{C}}({\iota}){\mathcal{C}}_{l}}\frac{1}{|{\mathcal{C}}_{l}|}\sum_{j\in{\mathcal{C}}_{l}}l(z,u,y_{j}^{\infty}(t))-{l}[z,u,\mu_{G}(t);g_{I^{*}({\iota})}],

where zz\in\mathbb{R} and uUu\in U are deterministic and fixed.

Lemma 4.10.

We have

(4.16) limksupz,u,tE(|Δfk(z,u,\mathboldyt)|2+|Δlk(z,u,\mathboldyt)|2)=0.\displaystyle\lim_{k\to\infty}\sup_{z,u,t}E(|\Delta_{f}^{k}(z,u,{\mathbold y}^{\infty}_{t})|^{2}+|\Delta_{l}^{k}(z,u,{\mathbold y}^{\infty}_{t})|^{2})=0.
Proof.

As in the proof of Lemma 4.7, we approximate μβ\mu_{\beta}, β[0,1]\beta\in[0,1], by using a finite number of points of β\beta, and next expand the two quadratic terms in (4.16). The estimate is carried out using (H11) and Lemma A.1. ∎

Lemma 4.11.

For any given constant Cz>0C_{z}>0 and any ϵ(0,1)\epsilon\in(0,1),

limkinftP((z,u)[Cz,Cz]×U{|Δfk(z,u,\mathboldyt)|ϵ})=1,\displaystyle\lim_{k\to\infty}\inf_{t}P(\cap_{(z,u)\in[-C_{z},C_{z}]\times U}\{|\Delta_{f}^{k}(z,u,{\mathbold y}^{\infty}_{t})|\leq\epsilon\})=1,
limkinftP((z,u)[Cz,Cz]×U{|Δlk(z,u,\mathboldyt)|ϵ})=1.\displaystyle\lim_{k\to\infty}\inf_{t}P(\cap_{(z,u)\in[-C_{z},C_{z}]\times U}\{|\Delta_{l}^{k}(z,u,{\mathbold y}^{\infty}_{t})|\leq\epsilon\})=1.
Proof.

We establish the first limit, and may deal with the second one in the same way. Note that the event

(4.17) fCzk(z,u)[Cz,Cz]×U{|Δfk(z,u,\mathboldyt)|ϵ}\displaystyle{\mathcal{E}}_{fC_{z}}^{k}\coloneqq\cap_{(z,u)\in[-C_{z},C_{z}]\times U}\{|\Delta_{f}^{k}(z,u,{\mathbold y}^{\infty}_{t})|\leq\epsilon\}

is well defined since Δfk\Delta_{f}^{k} is continuous in (z,u)(z,u) and the intersection may be equivalently expressed using only a countable number of values of (z,u)(z,u) in [Cz,Cz]×U[-C_{z},C_{z}]\times U.

Take any ϵ(0,1)\epsilon\in(0,1). By (H2) and (H3), we can find δϵ>0\delta_{\epsilon}>0 such that |Δfk(z,u,\mathboldyt)Δfk(z,u,\mathboldyt)|ϵ/2|\Delta_{f}^{k}(z,u,{\mathbold y}^{\infty}_{t})-\Delta_{f}^{k}(z^{\prime},u^{\prime},{\mathbold y}^{\infty}_{t})|\leq\epsilon/2 whenever |zz|+|uu|δϵ|z-z^{\prime}|+|u-u^{\prime}|\leq\delta_{\epsilon}. For the selected δϵ\delta_{\epsilon}, we can find a fixed p0p_{0} and (zj,uj)[Cz,Cz]×U(z^{j},u^{j})\in[-C_{z},C_{z}]\times U, j=1,,p0j=1,\ldots,p_{0} such that for any (z,u)[Cz,Cz]×U(z,u)\in[-C_{z},C_{z}]\times U, there exists some j0j_{0} ensuring |zzj0|+|uuj0|δϵ|z-z^{j_{0}}|+|u-u^{j_{0}}|\leq\delta_{\epsilon}.

By Lemma 4.10 and Markov’s inequality, for any δ>0\delta>0, there exists Kδ,p0K_{\delta,p_{0}} such that for all kKδ,p0k\geq K_{\delta,p_{0}}, we have

(4.18) P({|Δfk(zj,uj,\mathboldyt)|ϵ/2})1δ/p0,j,t.\displaystyle P(\{|\Delta_{f}^{k}(z^{j},u^{j},{\mathbold y}^{\infty}_{t})|\leq\epsilon/2\})\geq 1-\delta/p_{0},\quad\forall j,t.

Let jk{\mathcal{E}}^{k}_{j} denote the event {|Δfk(zj,uj,\mathboldyt)|ϵ/2}\{|\Delta_{f}^{k}(z^{j},u^{j},{\mathbold y}^{\infty}_{t})|\leq\epsilon/2\}. By (4.18), P(j=1p0jk)1δP(\cap_{j=1}^{p_{0}}{\mathcal{E}}_{j}^{k})\geq 1-\delta for kKδ,p0k\geq K_{\delta,p_{0}}. Now if ωkj=1p0jk\omega\in{\mathcal{E}}^{k}\coloneqq\cap_{j=1}^{p_{0}}{\mathcal{E}}_{j}^{k}, kKδ,p0k\geq K_{\delta,p_{0}}, then for any (z,u)[Cz,Cz]×U(z,u)\in[-C_{z},C_{z}]\times U, we have |Δfk(z,u,\mathboldyt)|ϵ.|\Delta_{f}^{k}(z,u,{\mathbold y}^{\infty}_{t})|\leq\epsilon. Hence kfCzk.{\mathcal{E}}^{k}\subset{\mathcal{E}}_{fC_{z}}^{k}. It follows that for all kKδ,p0k\geq K_{\delta,p_{0}}, P(fCzk)1δ.P({\mathcal{E}}_{fC_{z}}^{k})\geq 1-\delta. Since δ(0,1)\delta\in(0,1) is arbitrary and Kδ,p0K_{\delta,p_{0}} does not depend on tt, the first limit follows. ∎

Lemma 4.12.

We have

limksupt,uιNE(|Δfk(xιN(t),uιN(t),\mathboldyt)|+|Δlk(xιN(t),uιN(t),\mathboldyt)|)=0.\displaystyle\lim_{k\to\infty}\sup_{t,u_{\iota}^{N}}E(|\Delta_{f}^{k}(x_{\iota}^{N}(t),u_{\iota}^{N}(t),{\mathbold y}^{\infty}_{t})|+|\Delta_{l}^{k}(x_{\iota}^{N}(t),u_{\iota}^{N}(t),{\mathbold y}^{\infty}_{t})|)=0.
Proof.

Fix any ϵ(0,1)\epsilon\in(0,1). By (H1) and (H2) we can find a sufficiently large CzC_{z}, independent of (k,N)(k,N), such that for all uιN()u_{\iota}^{N}(\cdot),

P(sup0tT|xιN(t)|Cz)1ϵ.\displaystyle P\Big{(}\sup_{0\leq t\leq T}|x_{\iota}^{N}(t)|\leq C_{z}\Big{)}\geq 1-\epsilon.

Denote x={sup0tT|xιN(t)|Cz}{\mathcal{E}}_{x}=\{\sup_{0\leq t\leq T}|x_{\iota}^{N}(t)|\leq C_{z}\}. By Lemma 4.11, for the above ϵ\epsilon and fCzk{\mathcal{E}}^{k}_{fC_{z}} given by (4.17), there exists K0K_{0} independent of tt such that for all kK0k\geq K_{0},

P(fCzk)1ϵ.P({\mathcal{E}}^{k}_{fC_{z}})\geq 1-\epsilon.

Now if ωxfCzk\omega\in{\mathcal{E}}_{x}\cap{\mathcal{E}}^{k}_{fC_{z}}, then |Δfk(xιN(t),uιN(t),\mathboldyt)|ϵ.|\Delta_{f}^{k}(x_{\iota}^{N}(t),u_{\iota}^{N}(t),{\mathbold y}^{\infty}_{t})|\leq\epsilon. We have P(xfCzk)12ϵP({\mathcal{E}}_{x}\cap{\mathcal{E}}^{k}_{fC_{z}})\geq 1-2\epsilon, and so

P(|Δfk(xιN(t),uιN(t),\mathboldyt)|ϵ)P(xfCzk)12ϵ.P(|\Delta_{f}^{k}(x_{\iota}^{N}(t),u_{\iota}^{N}(t),{\mathbold y}^{\infty}_{t})|\leq\epsilon)\geq P({\mathcal{E}}_{x}\cap{\mathcal{E}}^{k}_{fC_{z}})\geq 1-2\epsilon.

It follows that for all kK0k\geq K_{0},

E|Δfk(xιN(t),uιN(t),\mathboldyt)|ϵ+2ϵC,\displaystyle E|\Delta_{f}^{k}(x_{\iota}^{N}(t),u_{\iota}^{N}(t),{\mathbold y}^{\infty}_{t})|\leq\epsilon+2\epsilon C,

where CC does not depend on (uιN(),t)(u_{\iota}^{N}(\cdot),t). The bound for Δlk\Delta_{l}^{k} is similarly obtained. ∎

Lemma 4.13.

We have

limksupt,uιN()E(|δfk|+|δlk|)=0.\displaystyle\lim_{k\to\infty}\sup_{t,u_{\iota}^{N}(\cdot)}E(|\delta_{f}^{k}|+|\delta_{l}^{k}|)=0.
Proof.

By Lipschitz continuity of (f,l)(f,l), we estimate E|δfkΔfk(xιN,uιN,\mathboldyt)|E|\delta_{f}^{k}-\Delta_{f}^{k}(x_{\iota}^{N},u_{\iota}^{N},{\mathbold y}^{\infty}_{t})| and E|δlkΔlk(xιN,uιN,\mathboldyt)|E|\delta_{l}^{k}-\Delta_{l}^{k}(x_{\iota}^{N},u_{\iota}^{N},{\mathbold y}^{\infty}_{t})|, and next apply Lemma 4.9 to show that they converge to zero as kk\to\infty. Recalling Lemma 4.12, we complete the proof. ∎

Lemma 4.14.

We have

limksupt,uιN()E(|δf0k|+|δl0k|)=0.\displaystyle\lim_{k\to\infty}\sup_{t,u_{\iota}^{N}(\cdot)}E(|\delta_{f_{0}}^{k}|+|\delta_{l_{0}}^{k}|)=0.
Proof.

The proof is similar to that of Lemma 4.13 and the details are omitted. ∎

Denote

ϵflk=supt,uιN()E(|δf0k|+|δl0k|+|δfk|+|δlk|).\epsilon_{fl}^{k}=\sup_{t,u_{\iota}^{N}(\cdot)}E(|\delta_{f_{0}}^{k}|+|\delta_{l_{0}}^{k}|+|\delta_{f}^{k}|+|\delta_{l}^{k}|).
Lemma 4.15.

For any admissible control uιNu_{\iota}^{N} in System B and JιJ_{\iota}^{*} in (4.14),

JιN(uιN)infuιJι(uι)Cϵflk,\displaystyle J_{\iota}^{N}(u_{\iota}^{N})\geq\inf_{u_{\iota}}J_{\iota}^{*}(u_{\iota})-C\epsilon_{fl}^{k},

where the constant CC does not depend on uιNu_{\iota}^{N}.

Proof.

Take any full state based Lipschitz feedback control uιNu_{\iota}^{N}. It together with the other agents’s control laws generates the closed-loop state processes x1N(t),,xNN(t)x_{1}^{N}(t),\ldots,x_{N}^{N}(t). Let uιN(t,ω)u_{\iota}^{N}(t,\omega) denote the realization as a non-anticipative process. Now we take uˇι=uιN(t,ω)\check{u}_{\iota}=u_{\iota}^{N}(t,\omega) in (4.13) and let xˇι\check{x}_{\iota}^{\infty} be the resulting state process. It is clear from (4.14) that

(4.19) Jι(uˇι)infuιJι(uι).\displaystyle J_{\iota}^{*}(\check{u}_{\iota})\geq\inf_{u_{\iota}}J_{\iota}^{*}(u_{\iota}).

Recalling (4.15) and applying Gronwall’s lemma to estimate the difference xˇιxιN\check{x}_{\iota}^{\infty}-x_{\iota}^{N}, we can show there exists CC independent of uιNu_{\iota}^{N} such that |JιN(uιN)Jι(uˇι)|Cϵflk,|J_{\iota}^{N}(u_{\iota}^{N})-J_{\iota}^{*}(\check{u}_{\iota})|\leq C\epsilon_{fl}^{k}, which combined with (4.19) completes the proof. ∎

Lemma 4.16.

Let φI(ι)=φ(t,x,gI(ι))\varphi_{I^{*}({\iota})}=\varphi(t,x,g_{I^{*}({\iota})}) be the GMFG based control law (4.2). We have

JιN(φI(ι))infuιJι(uι)+Cϵflk.\displaystyle J^{N}_{\iota}(\varphi_{I^{*}({\iota})})\leq\inf_{u_{\iota}}J_{\iota}^{*}(u_{\iota})+C\epsilon_{fl}^{k}.
Proof.

Let φI(ι)\varphi_{I^{*}({\iota})} be applied to the two systems (4.13) and (4.15). We further use Gronwall’s lemma to estimate E|xιxιN|E|x_{\iota}^{\infty}-x_{\iota}^{N}|. We obtain |JιN(φI(ι))Jι(φI(ι))|Cϵflk|J_{\iota}^{N}(\varphi_{I^{*}({\iota})})-J_{\iota}^{*}(\varphi_{I^{*}({\iota})})|\leq C\epsilon_{fl}^{k}. Note that Jι(φI(ι))=infuιJι(uι)J_{\iota}^{*}(\varphi_{I^{*}({\iota})})=\inf_{u_{\iota}}J_{\iota}^{*}(u_{\iota}). This completes the proof. ∎

Proof of Theorem 4.4. It follows from Lemmas 4.13, 4.14, 4.15 and 4.16. ∎

5. The LQ Case

This section considers a special class of linear-quadratic-Gaussian (LQG) GMFG models. Consider the graph GkG_{k} with vertices 𝒱k={1,,Mk}{\mathcal{V}}_{k}=\{1,\ldots,M_{k}\} and graph adjacency matrix gk=[gjlk]g^{k}=[g^{k}_{jl}]. For agent 𝒜i{\mathcal{A}}_{i} in subpopulation cluster 𝒞q{\mathcal{C}}_{q} situated at node qq, let the intra- and inter-cluster coupling terms be denoted by z0,iz_{0,i} and ziz_{i}, respectively, where

z0,i=1|𝒞q|j𝒞qxj,zi=1|Mk|l𝒱kgqlk1|𝒞l|j𝒞lxj,xj,z0,i,zin.z_{0,i}=\frac{1}{|\mathcal{C}_{q}|}\sum_{j\in\mathcal{C}_{q}}x_{j},\quad z_{i}=\frac{1}{|M_{k}|}\sum_{l\in{\mathcal{V}}_{k}}g^{k}_{ql}\frac{1}{|\mathcal{C}_{l}|}\sum_{j\in\mathcal{C}_{l}}x_{j},\quad x_{j},\ z_{0,i},\ z_{i}\in{\mathbb{R}}^{n}.

The dynamics of 𝒜i{\mathcal{A}}_{i} are given by the linear system

dxi=(Axi+D0z0,i+Dzi+Bui)dt+Σdwi,1iN,\displaystyle dx_{i}=(Ax_{i}+D_{0}z_{0,i}+Dz_{i}+Bu_{i})dt+\Sigma dw_{i},\quad 1\leq i\leq N,

where uinuu_{i}\in{\mathbb{R}}^{n_{u}} is the control input, winww_{i}\in{\mathbb{R}}^{n_{w}} is a standard Brownian motion, and AA, BB, D0D_{0}, DD, Σ\Sigma are conformally dimensioned matrices. Assume Exi(0)=x0Ex_{i}(0)=x_{0} for all ii.

The individual agent’s cost function takes the form

Ji(ui;νi)=\displaystyle J_{i}(u_{i};\nu_{i})= E0T[(xiνi)TQ(xiνi)+uiTRui]𝑑t\displaystyle{E}\int_{0}^{T}\big{[}(x_{i}-\nu_{i})^{T}Q(x_{i}-\nu_{i})+u_{i}^{T}Ru_{i}\big{]}dt
+E[(xi(T)νi(T))TQT(xi(T)νi(T))],1iN,\displaystyle+E\big{[}(x_{i}(T)-\nu_{i}(T))^{T}Q_{T}(x_{i}(T)-\nu_{i}(T))\big{]},\quad 1\leq i\leq N,

where QQ, QT0Q_{T}\geq 0, R>0R>0, and νi=γ0z0,i+γzi+η\nu_{i}=\gamma_{0}z_{0,i}+\gamma z_{i}+\eta is the process tracked by 𝒜i{\mathcal{A}}_{i}. Here ηn\eta\in\mathbb{R}^{n} and γ0,γ\gamma_{0},\gamma\in\mathbb{R}.

In the infinite population and graphon limit case, denote the local mean nxμα(dx)\int_{{\mathbb{R}}^{n}}x\mu_{\alpha}(dx) at tt for an α\alpha-agent situated at vertex α\alpha by x¯α\bar{x}_{\alpha}, and the graphon weighted mean 01g(α,β)x¯β𝑑β\int_{0}^{1}g(\alpha,\beta)\bar{x}_{\beta}d\beta by zαz_{\alpha}. The α\alpha-agent’s state equation is given by

dxα=(Axα+D0x¯α+Dzα+Buα)dt+Σdwα,α[0,1].\displaystyle dx_{\alpha}=(Ax_{\alpha}+D_{0}\bar{x}_{\alpha}+Dz_{\alpha}+Bu_{\alpha})dt+\Sigma dw_{\alpha},\quad\alpha\in[0,1].

The α\alpha-agent’s cost function is

Jα(uα;να)=\displaystyle J_{\alpha}(u_{\alpha};\nu_{\alpha})= E0T[(xανα)TQ(xανα)+uαTRuα]𝑑t\displaystyle{E}\int_{0}^{T}\big{[}(x_{\alpha}-\nu_{\alpha})^{T}Q(x_{\alpha}-\nu_{\alpha})+u_{\alpha}^{T}Ru_{\alpha}\big{]}dt
+E[(xα(T)να(T))TQT(xα(T)να(T))],\displaystyle+E\big{[}(x_{\alpha}(T)-\nu_{\alpha}(T))^{T}Q_{T}(x_{\alpha}(T)-\nu_{\alpha}(T))\big{]},

where να=γ0x¯α+γzα+η\nu_{\alpha}=\gamma_{0}\bar{x}_{\alpha}+\gamma z_{\alpha}+\eta.

Consider the Riccati equation

0=Π˙t+ATΠt+ΠtAΠtBR1BTΠt+Q,\displaystyle 0=\dot{\Pi}_{t}+A^{T}\Pi_{t}+\Pi_{t}A-\Pi_{t}BR^{-1}B^{T}\Pi_{t}+Q,

where ΠT=QT\Pi_{T}=Q_{T}, and

0=s˙α(t)+(ABR1BTΠt)Tsα(t)+Πt(D0x¯α(t)+Dzα(t))Qνα(t),\displaystyle 0=\dot{s}_{\alpha}(t)+(A-BR^{-1}B^{T}\Pi_{t})^{T}{s}_{\alpha}(t)+{\Pi}_{t}(D_{0}\bar{x}_{\alpha}(t)+Dz_{\alpha}(t))-Q\nu_{\alpha}(t),

where sα(T)=QTνα(T)s_{\alpha}(T)=-Q_{T}\nu_{\alpha}(T). The best response for the α\alpha-agent is given by

uα(t)=R1BT[Πtxα(t)+sα(t)].u_{\alpha}(t)=-R^{-1}B^{T}[\Pi_{t}x_{\alpha}(t)+s_{\alpha}(t)].

Now the mean state process of xα{x}_{\alpha} is

x¯˙α=(ABR1BTΠt+D0)x¯α+DzαBR1BsαT,α[0,1].\dot{\bar{x}}_{\alpha}=(A-BR^{-1}B^{T}\Pi_{t}+D_{0})\bar{x}_{\alpha}+Dz_{\alpha}-BR^{-1}B{{}^{T}}s_{\alpha},\quad\alpha\in[0,1].

The existence analysis reduces to verifying the existence and uniqueness of solutions for the equation system

(5.1) x¯˙α=(ABR1BTΠt+D0)x¯αBR1BTsα+D01g(α,β)x¯β𝑑β,\displaystyle\dot{\bar{x}}_{\alpha}=(A-BR^{-1}B^{T}\Pi_{t}+D_{0})\bar{x}_{\alpha}-BR^{-1}B^{T}s_{\alpha}+D\int_{0}^{1}g(\alpha,\beta)\bar{x}_{\beta}d\beta,
(5.2) s˙α=(ABR1BTΠt)Tsα+(γ0QΠtD0)x¯α\displaystyle\dot{s}_{\alpha}=-(A-BR^{-1}B^{T}\Pi_{t})^{T}s_{\alpha}+(\gamma_{0}Q-\Pi_{t}D_{0})\bar{x}_{\alpha}
+(γQΠtD)01g(α,β)x¯β𝑑β+Qη,\displaystyle\qquad+(\gamma Q-\Pi_{t}D)\int_{0}^{1}g(\alpha,\beta)\bar{x}_{\beta}d\beta+Q\eta,

where x¯α(0)=x0\bar{x}_{\alpha}(0)=x_{0} and sα(T)=QT[γ0x¯α(T)+γ01g(α,β)x¯β(T)𝑑β+η]s_{\alpha}(T)=-Q_{T}[\gamma_{0}\bar{x}_{\alpha}(T)+\gamma\int_{0}^{1}g(\alpha,\beta)\bar{x}_{\beta}(T)d\beta+\eta].

To analyze (5.1)–(5.2), let Φ(t,s)\Phi(t,s) and Ψ(t,s)\Psi(t,s) be the fundamental solution matrix of

x˙=(ABR1BTΠt+D0)x,y˙=(ABR1BTΠt)Ty\displaystyle\dot{x}=(A-BR^{-1}B^{T}\Pi_{t}+D_{0})x,\qquad\dot{y}=-(A-BR^{-1}B^{T}\Pi_{t})^{T}y

for x(t),y(t)nx(t),y(t)\in\mathbb{R}^{n}. For the special case with D0=0D_{0}=0, Ψ(t,s)=ΦT(s,t)\Psi(t,s)=\Phi^{T}(s,t) holds. We convert the existence analysis into a fixed point problem. We view x¯β(t)=x¯(β,t)\bar{x}_{\beta}(t)=\bar{x}(\beta,t) as a function of (β,t)(\beta,t). Below we derive an equation for x¯α(t)\bar{x}_{\alpha}(t) by eliminating sα(t)s_{\alpha}(t). Denote the function space DΛD_{\Lambda} consisting of continuous n\mathbb{R}^{n}-valued functions on [0,1]×[0,T][0,1]\times[0,T] with norm xˇ=supα,t|xˇ(α,t)|\|\check{x}\|=\sup_{\alpha,t}|\check{x}(\alpha,t)|. We use |||\cdot| to denote the Frobenius norm of a vector or matrix. Define the operator Λ\Lambda as follows: for xˇDΛ\check{x}\in D_{\Lambda},

(Λxˇ)(α,t)=\displaystyle({\Lambda}\check{x})(\alpha,t)=\ 0tΦ(t,r)BR1BT{rTΨ(r,τ)[(γ0QΠτD0)xˇ(α,τ)\displaystyle\int_{0}^{t}\Phi(t,r)BR^{-1}B^{T}\Big{\{}\int_{r}^{T}\Psi(r,\tau)\Big{[}(\gamma_{0}Q-\Pi_{\tau}D_{0})\check{x}(\alpha,\tau)
+(γQΠτD)01g(α,β)xˇ(β,τ)dβ]dτ\displaystyle\qquad\qquad+(\gamma Q-\Pi_{\tau}D)\int_{0}^{1}g(\alpha,\beta)\check{x}(\beta,\tau)d\beta\Big{]}d\tau
+Ψ(r,T)QT[γ0xˇ(α,T)+γ01g(α,β)xˇ(β,T)dβ]}dr\displaystyle+\Psi(r,T)Q_{T}\Big{[}\gamma_{0}\check{x}(\alpha,T)+\gamma\int_{0}^{1}g(\alpha,\beta)\check{x}(\beta,T)d\beta\Big{]}\Big{\}}dr
+0tΦ(t,r)D01g(α,β)xˇ(β,r)𝑑β𝑑r.\displaystyle+\int_{0}^{t}\Phi(t,r)D\int_{0}^{1}g(\alpha,\beta)\check{x}(\beta,r)d\beta dr.

If (H5) holds, Λ\Lambda is from DΛD_{\Lambda} to itself.

The solution of the LQG GMFG reduces to finding a fixed point xˇ\check{x} to the equation

xˇ(α,t)=\displaystyle\check{x}(\alpha,t)= (Λxˇ)(α,t)+Φ(t,0)x0\displaystyle(\Lambda\check{x})(\alpha,t)+\Phi(t,0)x_{0}
+0tΦ(t,r)BR1BT[rTΨ(r,τ)Q𝑑τ+Ψ(r,T)QT]η𝑑r.\displaystyle+\int_{0}^{t}\Phi(t,r)BR^{-1}B^{T}\Big{[}\!\int_{r}^{T}\Psi(r,\tau)Qd\tau+\Psi(r,T)Q_{T}\Big{]}\eta dr.

Denote cg=maxα01g(α,β)𝑑βc_{g}=\max_{\alpha}\int_{0}^{1}g(\alpha,\beta)d\beta. We have the bound for the operator norm:

ΛcΛ\displaystyle\|\Lambda\|\leq c_{\Lambda}\coloneqq supt[0,T]{0trT|Φ(t,r)BR1BTΨ(r,τ)|(|γ0QΠτD0|\displaystyle\sup_{t\in[0,T]}\Big{\{}\int_{0}^{t}\int_{r}^{T}|\Phi(t,r)BR^{-1}B^{T}\Psi(r,\tau)|\cdot(|\gamma_{0}Q-\Pi_{\tau}D_{0}|
+cg|γQΠτD|)dτdr\displaystyle\qquad\qquad+c_{g}|\gamma Q-\Pi_{\tau}D|)d\tau dr
+0t[|Φ(t,r)BR1BTΨ(r,T)QT|(|γ0|+cg|γ|)+cg|Φ(t,r)D|]dr}.\displaystyle+\int_{0}^{t}\Big{[}|\Phi(t,r)BR^{-1}B^{T}\Psi(r,T)Q_{T}|\cdot(|\gamma_{0}|+c_{g}|\gamma|)+c_{g}|\Phi(t,r)D|\Big{]}dr\Big{\}}.

If cΛ<1,c_{\Lambda}<1, Λ\Lambda is a contraction and (5.1)–(5.2) has a unique solution.

As an example for illustration, we assume the graphon weighted mean at vertex α\alpha arises from an underlying uniform attachment graphon, and consequently

zα=01(1max(α,β))nxμβ(dx)𝑑β,α,β[0,1],z_{\alpha}=\int_{0}^{1}(1-\max(\alpha,\beta))\int_{\mathbb{R}^{n}}x\mu_{\beta}(dx)d\beta,\quad\alpha,\beta\in[0,1],

where it is readily verified that the uniform attachment graphon satisfies (H5).

Appendix

Lemma A.1.

Assume (H1)–(H8). Let φα\varphi_{\alpha} be the GMFG based best response (4.2) and μα(t)\mu_{\alpha}(t) the distribution of the closed-loop process xα(t)x_{\alpha}(t), α[0,1]\alpha\in[0,1], in (3.15) with initial distribution μ0x\mu_{0}^{x}. Then we have

limr0sup|tt|+|ββ|<rW1(μβ(t),μβ(t))=0,\lim_{r\to 0}\sup_{|t-t^{*}|+|\beta-\beta^{*}|<r}W_{1}(\mu_{\beta}(t),\mu_{\beta^{*}}(t^{*}))=0,

where t,t[0,T]t,t^{*}\in[0,T] and β,β[0,1]\beta,\beta^{*}\in[0,1].

Proof.

Step 1. Take any β,β[0,1]\beta,\beta^{*}\in[0,1]. For μG()\mu_{G}(\cdot) determined from the GMFG equations (3.2) and (3.15), define two processes

dyβ=f~[yβ,φ(t,yβ,gβ),μG;gβ]dt+σdwβ,\displaystyle dy_{\beta^{*}}=\widetilde{f}[y_{\beta^{*}},\varphi(t,y_{\beta^{*}},g_{\beta^{*}}),\mu_{G};g_{\beta^{*}}]dt+\sigma dw_{\beta^{*}},
dyβ=f~[yβ,φ(t,yβ,gβ),μG;gβ]dt+σdwβ,\displaystyle dy_{\beta}=\widetilde{f}[y_{\beta},\varphi(t,y_{\beta},g_{\beta}),\mu_{G};g_{\beta}]dt+\sigma dw_{\beta^{*}},

where yβ(0)=yβ(0)=xiN(0)y_{\beta^{*}}(0)=y_{\beta}(0)=x_{i}^{N}(0) and the same Brownian motion is used. Then the distributions of yβ(t)y_{\beta^{*}}(t) and yβ(t)y_{\beta}(t) are μβ(t)\mu_{\beta^{*}}(t) and μβ(t)\mu_{\beta}(t), respectively. We obtain

yβ(t)yβ(t)\displaystyle y_{\beta}(t)-y_{\beta^{*}}(t)
=\displaystyle= 0tΔβ,β0(s)𝑑s+0t01Δβ,β(s,z,λ)μλ(s,dz)𝑑λ𝑑s,\displaystyle\int_{0}^{t}\Delta_{\beta,\beta^{*}}^{0}(s)ds+\int_{0}^{t}\int_{0}^{1}\int_{\mathbb{R}}\Delta_{\beta,\beta^{*}}(s,z,\lambda)\mu_{\lambda}(s,dz)d\lambda ds,

where

Δβ,β0(s)=f0(yβ,φ(s,yβ,gβ),z)μβ(s,dz)f0(yβ,φ(s,yβ,gβ),z)μβ(s,dz),\displaystyle\Delta_{\beta,\beta^{*}}^{0}(s)=\int_{\mathbb{R}}f_{0}(y_{\beta},\varphi(s,y_{\beta},g_{\beta}),z)\mu_{\beta}(s,dz)-\int_{\mathbb{R}}f_{0}(y_{\beta^{*}},\varphi(s,y_{\beta^{*}},g_{\beta^{*}}),z)\mu_{\beta^{*}}(s,dz),
Δβ,β(s,z,λ)=f(yβ,φ(s,yβ,gβ),z)g(β,λ)\displaystyle\Delta_{\beta,\beta^{*}}(s,z,\lambda)=f(y_{\beta},\varphi(s,y_{\beta},g_{\beta}),z)g(\beta,\lambda)
f(yβ,φ(s,yβ,gβ),z)g(β,λ).\displaystyle\qquad\qquad\qquad-f(y_{\beta^{*}},\varphi(s,y_{\beta^{*}},g_{\beta^{*}}),z)g(\beta^{*},\lambda).

We will simply write μλ(s,dz)\mu_{\lambda}(s,dz) as μλ(dz)\mu_{\lambda}(dz) if the time argument is clear, where λ\lambda is the vertex index. Denote κβ,β(s)=|φ(s,yβ,gβ)φ(s,yβ,gβ)|\kappa_{\beta,\beta^{*}}(s)=|\varphi(s,y_{\beta^{*}},g_{\beta})-\varphi(s,y_{\beta^{*}},g_{\beta^{*}})|, where the time argument ss in yβy_{\beta} and yβy_{\beta^{*}} has been suppressed. It follows that

|Δβ,β0(s)|\displaystyle|\Delta_{\beta,\beta^{*}}^{0}(s)|\leq
|f0(yβ,φ(s,yβ,gβ),z)μβ(s,dz)f0(yβ,φ(s,yβ,gβ),z)μβ(s,dz)|\displaystyle\Big{|}\int_{\mathbb{R}}f_{0}(y_{\beta},\varphi(s,y_{\beta},g_{\beta}),z)\mu_{\beta}(s,dz)-\int_{\mathbb{R}}f_{0}(y_{\beta},\varphi(s,y_{\beta},g_{\beta}),z)\mu_{\beta^{*}}(s,dz)\Big{|}
+|f0(yβ,φ(s,yβ,gβ),z)μβ(s,dz)f0(yβ,φ(s,yβ,gβ),z)μβ(s,dz)|\displaystyle+\Big{|}\int_{\mathbb{R}}f_{0}(y_{\beta},\varphi(s,y_{\beta},g_{\beta}),z)\mu_{\beta^{*}}(s,dz)-\int_{\mathbb{R}}f_{0}(y_{\beta^{*}},\varphi(s,y_{\beta^{*}},g_{\beta^{*}}),z)\mu_{\beta^{*}}(s,dz)\Big{|}
CE|yβyβ|+C|yβyβ|+C|φ(s,yβ,gβ)φ(s,yβ,gβ)|\displaystyle\leq CE|y_{\beta}-y_{\beta^{*}}|+C|y_{\beta}-y_{\beta^{*}}|+C|\varphi(s,y_{\beta},g_{\beta})-\varphi(s,y_{\beta^{*}},g_{\beta^{*}})|
CE|yβyβ|+C1|yβyβ|+Cκβ,β(s),\displaystyle\leq CE|y_{\beta}-y_{\beta^{*}}|+C_{1}|y_{\beta}-y_{\beta^{*}}|+C\kappa_{\beta,\beta^{*}}(s),

where the second inequality is obtained using (H2), (H3), and the method in (4.11). The last inequality has used the uniform Lipschitz continuity of φβ\varphi_{\beta} in the space variable (see Lemma 3.7). It follows that

(A.1) E|Δβ,β0(s)|C2E|yβ(s)yβ(s)|+CEκβ,β(s).\displaystyle E|\Delta_{\beta,\beta^{*}}^{0}(s)|\leq C_{2}E|y_{\beta}(s)-y_{\beta^{*}}(s)|+CE\kappa_{\beta,\beta^{*}}(s).

Next, we have

|01Δβ,β(s,z,λ)μλ(dz)𝑑λ|\displaystyle\Big{|}\int_{0}^{1}\int_{\mathbb{R}}\Delta_{\beta,\beta^{*}}(s,z,\lambda)\mu_{\lambda}(dz)d\lambda\Big{|}
(A.2) \displaystyle\leq |01[f(yβ,φ(s,yβ,gβ),z)f(yβ,φ(s,yβ,gβ),z)]g(β,λ)μλ(dz)𝑑λ|\displaystyle\Big{|}\int_{0}^{1}\int_{\mathbb{R}}[f(y_{\beta},\varphi(s,y_{\beta},g_{\beta}),z)-f(y_{\beta^{*}},\varphi(s,y_{\beta^{*}},g_{\beta^{*}}),z)]g(\beta,\lambda)\mu_{\lambda}(dz)d\lambda\Big{|}
+|01f(yβ,φ(s,yβ,gβ),z)[g(β,λ)g(β,λ)]μλ(dz)𝑑λ|\displaystyle+\Big{|}\int_{0}^{1}\int_{\mathbb{R}}f(y_{\beta^{*}},\varphi(s,y_{\beta^{*}},g_{\beta^{*}}),z)[g(\beta,\lambda)-g(\beta^{*},\lambda)]\mu_{\lambda}(dz)d\lambda\Big{|}
=:\displaystyle=: If(s)+Ig(s).\displaystyle I_{f}(s)+I_{g}(s).

We have

If(s)\displaystyle I_{f}(s) 01C(|yβyβ|+κβ,β)g(β,λ)μλ(dz)𝑑λ\displaystyle\leq\int_{0}^{1}\int_{\mathbb{R}}C(|y_{\beta}-y_{\beta^{*}}|+\kappa_{\beta,\beta^{*}})g(\beta,\lambda)\mu_{\lambda}(dz)d\lambda
C(|yβyβ|+κβ,β)(s),\displaystyle\leq C(|y_{\beta}-y_{\beta^{*}}|+\kappa_{\beta,\beta^{*}})(s),

where we have used the Lipschitz property of ff and φβ\varphi_{\beta}. Therefore,

(A.3) EIf(s)C(E|yβ(s)yβ(s)|+Eκβ,β(s)).\displaystyle EI_{f}(s)\leq C(E|y_{\beta}(s)-y_{\beta^{*}}(s)|+E\kappa_{\beta,\beta^{*}}(s)).

For any fixed value yβ(s,ω)y_{\beta^{*}}(s,\omega), denote

ξβ,s,ω(λ)=f(yβ,φ(s,yβ,gβ),z)μλ(dz).\displaystyle\xi_{\beta^{*},s,\omega}(\lambda)=\int_{\mathbb{R}}f(y_{\beta^{*}},\varphi(s,y_{\beta^{*}},g_{\beta^{*}}),z)\mu_{\lambda}(dz).

We have

Ig(s)=|01ξβ,s,ω(λ)g(β,λ)𝑑λ01ξβ,s,ω(λ)g(β,λ)𝑑λ|.\displaystyle I_{g}(s)=\Big{|}\int_{0}^{1}\xi_{\beta^{*},s,\omega}(\lambda)g(\beta,\lambda)d\lambda-\int_{0}^{1}\xi_{\beta^{*},s,\omega}(\lambda)g(\beta^{*},\lambda)d\lambda\Big{|}.

Hence, by (H5), Ig(s)0I_{g}(s)\to 0 (ω,s)(\omega,s)-a.e. as ββ\beta\to\beta^{*}. It is clear Ig(s)I_{g}(s) is bounded by a fixed constant since ff is a bounded function. For the fixed β\beta^{*}, by Lemma 3.5, the random variable κβ,β(s)\kappa_{\beta,\beta^{*}}(s) is bounded and converges to zero with probability one. Denote δg=0TEIg(s)𝑑s\delta_{g}=\int_{0}^{T}EI_{g}(s)ds and δκ=0TEκβ,β(s)𝑑s\delta_{\kappa}=\int_{0}^{T}E\kappa_{\beta,\beta^{*}}(s)ds. By dominated convergence, we have

limββ(δg+δκ)=0.\lim_{\beta\to\beta^{*}}(\delta_{g}+\delta_{\kappa})=0.

By (A.1)–(A.3), it follows that

E|yβ(t)yβ(t)|C0tE|yβ(s)yβ(s)|𝑑s+C(δκ+δg).\displaystyle E|y_{\beta}(t)-y_{\beta^{*}}(t)|\leq C\int_{0}^{t}E|y_{\beta}(s)-y_{\beta^{*}}(s)|ds+C(\delta_{\kappa}+\delta_{g}).

By Gronwall’s lemma, we have

sup0tTE|yβ(t)yβ(t)|CeCT(δκ+δg).\displaystyle\sup_{0\leq t\leq T}E|y_{\beta}(t)-y_{\beta^{*}}(t)|\leq Ce^{CT}(\delta_{\kappa}+\delta_{g}).

Since W1(μβ(t),μβ(t))E|yβ(t)yβ(t)|W_{1}(\mu_{\beta}(t),\mu_{\beta^{*}}(t))\leq E|y_{\beta}(t)-y_{\beta^{*}}(t)|, then

(A.4) suptW1(μβ(t),μβ(t))C1(δκ+δg),\displaystyle\sup_{t}W_{1}(\mu_{\beta}(t),\mu_{\beta^{*}}(t))\leq C_{1}(\delta_{\kappa}+\delta_{g}),

where δκ\delta_{\kappa} and δg\delta_{g} depend on β\beta^{*}.

Step 2. Now we consider given (β,t)[0,1]×[0,T](\beta^{*},t^{*})\in[0,1]\times[0,T]. By use of the SDE of yβy_{\beta} and elementary estimates, we obtain

(A.5) lim|tt|0supβW1(μβ(t),μβ(t))=0.\displaystyle\lim_{|t-t^{*}|\to 0}\sup_{\beta}W_{1}(\mu_{\beta}(t^{*}),\mu_{\beta}(t))=0.

We have

W1(μβ(t),μβ(t))W1(μβ(t),μβ(t))+W1(μβ(t),μβ(t)).\displaystyle W_{1}(\mu_{\beta}(t),\mu_{\beta^{*}}(t^{*}))\leq W_{1}(\mu_{\beta}(t),\mu_{\beta}(t^{*}))+W_{1}(\mu_{\beta}(t^{*}),\mu_{\beta^{*}}(t^{*})).

Given any ϵ>0\epsilon>0, by (A.4) and (A.5) there exists δϵ,β>0\delta_{\epsilon,\beta^{*}}>0 such that whenever |tt|+|ββ|δϵ,β|t-t^{*}|+|\beta-\beta^{*}|\leq\delta_{\epsilon,\beta^{*}}, we have

W1(μβ(t),μβ(t))ϵ2,W1(μβ(t),μβ(t))ϵ2.W_{1}(\mu_{\beta}(t),\mu_{\beta}(t^{*}))\leq\frac{\epsilon}{2},\qquad W_{1}(\mu_{\beta}(t^{*}),\mu_{\beta^{*}}(t^{*}))\leq\frac{\epsilon}{2}.

Therefore, W1(μβ(t),μβ(t))ϵW_{1}(\mu_{\beta}(t),\mu_{\beta^{*}}(t^{*}))\leq\epsilon. We conclude that μβ(t)\mu_{\beta}(t) as a mapping from the compact space [0,1]×[0,T][0,1]\times[0,T] to 𝒫1(){\mathcal{P}}_{1}({\mathbb{R}}) with the metric W1(,)W_{1}(\cdot,\cdot) is continuous and hence must be uniformly continuous. The lemma follows. ∎

Lemma A.2.

Suppose the graphon gg satisfies (H5) and (H11). Then for any given measurable sets 𝒮,𝒯[0,1]\mathcal{S},\mathcal{T}\subset[0,1], under (H9) we have

(A.6) limk|𝒮×𝒯(gkg)𝑑x𝑑y|=0.\displaystyle\lim_{k\to\infty}\Big{|}\int_{{\mathcal{S}}\times{\mathcal{T}}}(g^{k}-g)dxdy\Big{|}=0.
Proof.

Step 1. We approximate 𝒮,𝒯{\mathcal{S}},{\mathcal{T}} by open sets. Let μL\mu_{\rm L} denote the Lebesgue measure on d\mathbb{R}^{d}, where the dimension dd will be clear from the context. Consider the given sets 𝒮,𝒯{\mathcal{S}},{\mathcal{T}}, and choose an arbitrary ϵ>0\epsilon>0. Note that for any measurable set A1dA_{1}\subset\mathbb{R}^{d} and any δ0>0\delta_{0}>0, there exists an open set A2A1A_{2}\supset A_{1} such that μL(A2\A1)δ0\mu_{\rm L}(A_{2}\backslash A_{1})\leq\delta_{0} (see e.g. [36]). So there exist open sets 𝒮o\mathcal{S}^{o}\subset\mathbb{R} and 𝒯o{\mathcal{T}}^{o}\subset\mathbb{R} such that 𝒮𝒮o\mathcal{S}\subset\mathcal{S}^{o}, 𝒯𝒯o{\mathcal{T}}\subset{\mathcal{T}}^{o} and μL(𝒮o\𝒮)ϵ\mu_{\rm L}(\mathcal{S}^{o}\backslash\mathcal{S})\leq\epsilon, μL(𝒯o\𝒯)ϵ\mu_{\rm L}({\mathcal{T}}^{o}\backslash{\mathcal{T}})\leq\epsilon.

Define the new open sets 𝒮1o=𝒮o(0,1)\mathcal{S}_{1}^{o}=\mathcal{S}^{o}\cap(0,1) and 𝒯1o=𝒯o(0,1){\mathcal{T}}^{o}_{1}={\mathcal{T}}^{o}\cap(0,1). Each open set in \mathbb{R} may be written as the union of at most countable disjoint open intervals [36]; among such a union for 𝒮1o\mathcal{S}_{1}^{o}, we may find a finite integer ss^{*} (depending on (𝒮,ϵ)({\mathcal{S}},\epsilon)) and constituent disjoint open intervals Ii𝒮[0,1]I_{i}^{\mathcal{S}}\subset[0,1], 1is1\leq i\leq s^{*}, such that Usi=1sIi𝒮𝒮1oU_{s^{*}}\coloneqq\cup_{i=1}^{s^{*}}I_{i}^{\mathcal{S}}\subset{\mathcal{S}}_{1}^{o} and μL(𝒮1o\Us)ϵ\mu_{\rm L}(\mathcal{S}_{1}^{o}\backslash U_{s^{*}})\leq\epsilon. Similarly, we find a finite integer tt^{*} and disjoint open intervals Ii𝒯[0,1]I_{i}^{\mathcal{T}}\subset[0,1] such that Utj=1tIj𝒯𝒯1oU_{t^{*}}\coloneqq\cup_{j=1}^{t^{*}}I_{j}^{\mathcal{T}}\subset{\mathcal{T}}_{1}^{o} and μL(𝒯1o\Ut)ϵ\mu_{\rm L}({\mathcal{T}}_{1}^{o}\backslash U_{t^{*}})\leq\epsilon. Here the choice of (s,t)(s^{*},t^{*}) depends on (𝒮,𝒯,ϵ)(\mathcal{S},\mathcal{T},\epsilon).

By the construction of UsU_{s^{*}} and UtU_{t^{*}}, we have the bound for the measure of the following symmetric differences:

μL(𝒮ΔUs)2ϵ,μL(𝒯ΔUt)2ϵ,\displaystyle\mu_{\rm L}(\mathcal{S}\Delta U_{s^{*}})\leq 2{\epsilon},\quad\mu_{\rm L}(\mathcal{T}\Delta U_{t^{*}})\leq 2{\epsilon},

which implies μL((𝒮×𝒯)Δ(Us×Ut))6ϵ\mu_{\rm L}((\mathcal{S}\times\mathcal{T})\Delta(U_{s^{*}}\times U_{t^{*}}))\leq 6\epsilon. Since |gkg|1|g^{k}-g|\leq 1 for any x,yx,y, we have

(A.7) |𝒮×𝒯(gkg)𝑑x𝑑yηk|6ϵ,\displaystyle\Big{|}\int_{\mathcal{S}\times\mathcal{T}}(g^{k}-g)dxdy-\eta_{k}\Big{|}\leq{6\epsilon},

where

ηk\displaystyle\eta_{k} |Us×Ut(gkg)𝑑x𝑑y|.\displaystyle\coloneqq\Big{|}\int_{U_{s^{*}}\times U_{t^{*}}}(g^{k}-g)dxdy\Big{|}.

Step 2. Blow we estimate ηk\eta_{k}. Under (H9) we take a sufficiently large K0K_{0}, depending on s{s^{*}} (and so on (𝒮,ϵ)({\mathcal{S}},\epsilon)), such that for all kK0k\geq K_{0},

sMkϵ.\frac{s^{*}}{M_{k}}\leq{\epsilon}.

Consider kK0k\geq K_{0}. We select from the subintervals I1k,,IMkkI^{k}_{1},\ldots,I_{M_{k}}^{k} of equal length 1/Mk1/M_{k} in the partition of [0,1][0,1] such that a subinterval is selected whenever its interior is contained in UsU_{s^{*}}. The method here is to fill UsU_{s^{*}} as much as possible from inside by these subintervals. This procedure determines a subcollection denoted by IirkI^{k}_{i_{r}}, r=1,,rkr=1,\ldots,r_{k}. Denote U^s=r=1rkIirk\hat{U}_{s^{*}}=\cup_{r=1}^{r_{k}}I^{k}_{i_{r}}. Then the interior of U^s\hat{U}_{s^{*}} is contained in UsU_{s^{*}}. We need to estimate the measure for the part of UsU_{s^{*}} not covered by U^s\hat{U}_{s^{*}}. We check Ii𝒮I_{i}^{\mathcal{S}}, 1is1\leq i\leq s^{*}, to obtain two cases: (i) Ii𝒮U^sI_{i}^{\mathcal{S}}\subset\hat{U}_{s^{*}}, (ii) Ii𝒮I_{i}^{\mathcal{S}} has a portion (allowed to be equal to its whole) of positive measure staying outside U^s\hat{U}_{s^{*}}. For case (ii), the portion of Ii𝒮I_{i}^{\mathcal{S}} that is not covered by U^s\hat{U}_{s^{*}} consists of either one interval, as part or the whole of Ii𝒮I_{i}^{\mathcal{S}}, or two intervals each having an endpoint of Ii𝒮I_{i}^{\mathcal{S}} as its boundary; hence the measure of that portion is less than 2/Mk2/M_{k}. It follows that

(A.8) μL(Us\U^s)2sMk2ϵ.\displaystyle\mu_{\rm L}(U_{s^{*}}\backslash\hat{U}_{s^{*}})\leq\frac{2s^{*}}{M_{k}}\leq 2{\epsilon}.

By (A.8), for all kK0k\geq K_{0}, we have

(A.9) |Us×Ut(gkg)𝑑x𝑑yU^s×Ut(gkg)𝑑x𝑑y|2ϵ.\displaystyle\Big{|}\int_{U_{s^{*}}\times U_{t^{*}}}(g^{k}-g)dxdy-\int_{\hat{U}_{s^{*}}\times U_{t^{*}}}(g^{k}-g)dxdy\Big{|}\leq{2\epsilon}.

Step 3. Now for kK0k\geq K_{0} we check

η^k|U^s×Ut(gkg)𝑑x𝑑y|.\displaystyle\hat{\eta}_{k}\coloneqq\Big{|}\int_{\hat{U}_{s^{*}}\times U_{t^{*}}}(g^{k}-g)dxdy\Big{|}.

By (H5), for the selected UtU_{t^{*}}, Utg(x,y)𝑑y\int_{U_{t^{*}}}g(x,y)dy as a function of xx is uniformly continuous on [0,1][0,1]. So for ϵ\epsilon chosen in Step 1, there exists δ>0\delta>0 (depending on gg, ϵ\epsilon and UtU_{t^{*}}) such that

(A.10) |Utg(x,y)𝑑yUtg(x,y)𝑑y|ϵ\displaystyle\Big{|}\int_{U_{t^{*}}}g(x,y)dy-\int_{U_{t^{*}}}g(x^{\prime},y)dy\Big{|}\leq{\epsilon}

whenever |xx|δ|x-x^{\prime}|\leq\delta. For the above δ\delta, we fix K1K0K_{1}\geq K_{0} such that for all kK1k\geq K_{1}, we have 1/Mk2δ1/{M_{k}}\leq 2\delta. Note that we use (Iirk)(I_{i_{r}}^{k})^{*} to denote the midpoint of the interval IirkI_{i_{r}}^{k}. Now for kK1k\geq K_{1}, we have

η^k\displaystyle\hat{\eta}_{k} =|r=1rkIirkUt[gk(x,y)g(x,y)]𝑑y𝑑x|\displaystyle=\Big{|}\sum_{r=1}^{r_{k}}\int_{I^{k}_{i_{r}}}\int_{U_{t^{*}}}[g^{k}(x,y)-g(x,y)]dydx\Big{|}
|r=1rkIirkUt[gk((Iirk),y)g((Iirk),y)]𝑑y𝑑x|+ϵ\displaystyle\leq\Big{|}\sum_{r=1}^{r_{k}}\int_{I^{k}_{i_{r}}}\int_{U_{t^{*}}}[g^{k}((I^{k}_{i_{r}})^{*},y)-g((I^{k}_{i_{r}})^{*},y)]dydx\Big{|}+{\epsilon}
=|r=1rk1MkUt[gk((Iirk),y)g((Iirk),y)]𝑑y|+ϵ\displaystyle=\Big{|}\sum_{r=1}^{r_{k}}\frac{1}{M_{k}}\int_{U_{t^{*}}}[g^{k}((I^{k}_{i_{r}})^{*},y)-g((I^{k}_{i_{r}})^{*},y)]dy\Big{|}+{\epsilon}
1Mkr=1rkζk+ϵ,\displaystyle\leq\frac{1}{M_{k}}\sum_{r=1}^{r_{k}}\zeta_{k}+{\epsilon},

where

ζk|Ut[gk((Iirk),y)g((Iirk),y)]𝑑y|.\displaystyle\zeta_{k}\coloneqq\Big{|}\int_{U_{t^{*}}}[g^{k}((I^{k}_{i_{r}})^{*},y)-g((I^{k}_{i_{r}})^{*},y)]dy\Big{|}.

The first inequality follows from (A.10) and μL(r=1rkIirk)1\mu_{\rm L}(\cup_{r=1}^{r_{k}}I^{k}_{i_{r}})\leq 1.

Step 4. Now we estimate ζk\zeta_{k}. As in Step 2, we take a sufficiently large K2K1K_{2}\geq K_{1}, depending on (t,ϵ)({t^{*},\epsilon}), such that for all kK2k\geq K_{2}, t/Mkϵ.{t^{*}}/{M_{k}}\leq{\epsilon}. For kK2k\geq K_{2} and the subintervals I1k,,IMkkI^{k}_{1},\ldots,I_{M_{k}}^{k}, as in Step 2, we select a subcollection denoted by IjτkI^{k}_{j_{\tau}}, τ=1,,τk\tau=1,\ldots,\tau_{k}, each of which is selected whenever its interior is contained in UtU_{t^{*}}. Then it follows that

(A.11) μL(Ut\τ=1τkIjτk)2tMk2ϵ.\displaystyle\mu_{\rm L}(U_{t^{*}}\backslash\cup_{\tau=1}^{\tau_{k}}I^{k}_{j_{\tau}})\leq\frac{2t^{*}}{M_{k}}\leq 2\epsilon.

By (A.11), we have for all kK2k\geq K_{2},

ζk\displaystyle\zeta_{k}\leq |τ=1τkIjτk[gk((Iirk),y)g((Iirk),y)]𝑑y|+2ϵ\displaystyle\Big{|}\int_{\cup_{\tau=1}^{\tau_{k}}I^{k}_{j_{\tau}}}[g^{k}((I^{k}_{i_{r}})^{*},y)-g((I^{k}_{i_{r}})^{*},y)]dy\Big{|}+2{\epsilon}
\displaystyle\leq τ=1τk|girjτkMkβIjτkg(Iirk),β𝑑β|+2ϵ.\displaystyle\sum_{\tau=1}^{\tau_{k}}\Big{|}\frac{g^{k}_{i_{r}j_{\tau}}}{M_{k}}-\int_{\beta\in I^{k}_{j_{\tau}}}g_{(I^{k}_{i_{r}})^{*},\beta}d\beta\Big{|}+2{\epsilon}.

We write g(α,β)g(\alpha,\beta) as gα,βg_{\alpha,\beta}.

Step 5. Note that rk,τkMkr_{k},\tau_{k}\leq M_{k}. Subsequently, by Step 3 and Step 4, we have for kK2k\geq K_{2},

η^k\displaystyle\hat{\eta}_{k} 1Mkr=1rk[τ=1τk|girjτkMkβIjτkg(Iirk),β𝑑β|+2ϵ]+ϵ\displaystyle\leq\frac{1}{M_{k}}\sum_{r=1}^{r_{k}}\Big{[}\sum_{\tau=1}^{\tau_{k}}\Big{|}\frac{g^{k}_{i_{r}j_{\tau}}}{M_{k}}-\int_{\beta\in I^{k}_{j_{\tau}}}g_{(I^{k}_{i_{r}})^{*},\beta}d\beta\Big{|}+2{\epsilon}\Big{]}+{\epsilon}
1Mkr=1rkτ=1τk|girjτkMkβIjτkg(Iirk),β𝑑β|+3ϵ\displaystyle\leq\frac{1}{M_{k}}\sum_{r=1}^{r_{k}}\sum_{\tau=1}^{\tau_{k}}\Big{|}\frac{g^{k}_{i_{r}j_{\tau}}}{M_{k}}-\int_{\beta\in I^{k}_{j_{\tau}}}g_{(I^{k}_{i_{r}})^{*},\beta}d\beta\Big{|}+{3\epsilon}
(A.12) maxij=1Mk|g𝒞i𝒞jkMkβIjkg(Iik),β𝑑β|+3ϵ.\displaystyle\leq\max_{i}\sum_{j=1}^{M_{k}}\Big{|}\frac{g^{k}_{{\mathcal{C}}_{i}{\mathcal{C}}_{j}}}{M_{k}}-\int_{\beta\in I_{j}^{k}}g_{(I^{k}_{i})^{*},\beta}d\beta\Big{|}+{3\epsilon}.

By (A.7), (A.9) and (A.12), we obtain for all kK2k\geq K_{2} depending on (𝒮,𝒯,ϵ)({\mathcal{S}},{\mathcal{T}},\epsilon),

|𝒮×𝒯(gkg)𝑑x𝑑y|\displaystyle\Big{|}\int_{\mathcal{S}\times\mathcal{T}}(g^{k}-g)dxdy\Big{|} maxij=1Mk|g𝒞i𝒞jkMkβIjkg(Ijk),β𝑑β|+11ϵ.\displaystyle\leq\max_{i}\sum_{j=1}^{M_{k}}\Big{|}\frac{g^{k}_{{\mathcal{C}}_{i}{\mathcal{C}}_{j}}}{M_{k}}-\int_{\beta\in I_{j}^{k}}g_{(I^{k}_{j})^{*},\beta}d\beta\Big{|}+11\epsilon.

The lemma follows. ∎

References

  • [1] Erhan Bayraktar, Suman Chakraborty, and Ruoyu Wu, Graphon mean field systems, arXiv:2003.13180 (2020).
  • [2] Christian Borgs, Jennifer Chayes, László Lovász, Vera T Sós, Balázs Szegedy, and Katalin Vesztergombi, Graph limits and parameter testing, Proc. the thirty-eighth annual ACM symposium on Theory of computing, 2006, pp. 261–270.
  • [3] Christian Borgs, Jennifer T Chayes, László Lovász, Vera T Sós, and Katalin Vesztergombi, Convergent sequences of dense graphs I: Subgraph frequencies, metric properties and testing, Advances in Mathematics 219 (2008), no. 6, 1801–1851.
  • [4] by same author, Convergent sequences of dense graphs II. multiway cuts and statistical physics, Annals of Mathematics 176 (2012), no. 1, 151–219.
  • [5] Peter E Caines, Mean field games, Encyclopedia of Systems and Control (2015), 706–712.
  • [6] Peter E Caines and Minyi Huang, Graphon mean field games and the GMFG equations, Proc. 57th IEEE CDC (Miami Beach, FL, USA), 2018, pp. 4129–4134.
  • [7] by same author, Graphon mean field games and the GMFG equations: ϵ\epsilon-Nash equilibria, Proc. the 58th IEEE CDC (Nice, France), 2019, pp. 286–292.
  • [8] Peter E Caines, Minyi Huang, and Roland P Malhamé, Mean field games, Handbook of Dynamic Game Theory (Tamer Başar and Georges Zaccour, eds.), Springer, Berlin, 2017, pp. 345–372.
  • [9] Rene Carmona and Francois Delarue, Probabilistic theory of mean field games with applications I, vol. 83, Springer International Publishing, 2018.
  • [10] by same author, Probabilistic theory of mean field games with applications II, vol. 84, Springer International Publishing, 2018.
  • [11] François Delarue, Mean field games: A toy model on an Erdös-Renyi graph, ESAIM: Proceedings and Surveys 60 (2017), 1–26.
  • [12] Joseph L Doob, Stochastic processes, Wiley, New York, 1953.
  • [13] Wendell H. Fleming and Raymond W. Rishel, Deterministic and stochastic optimal control, Springer-Verlag, New York, 1975.
  • [14] Isabelle Gallagher, From Newton to Navier-Stokes, or how to connect fluid mechanics equations from microscopic to macroscopic scales, Bulletin of the American Math. Society 56 (2013), no. 1, 65–85.
  • [15] Isabelle Gallagher, Laure Saint-Raymond, and Benjamin Texier, From Newton to Boltzmann: hard spheres and short-range potentials, European Mathematical Society, 2013.
  • [16] Shuang Gao and Peter E. Caines, The control of arbitrary size networks of linear systems via graphon limits: An initial investigation, Proc. 56th IEEE CDC (Melbourne, Australia), December 2017, pp. 1052–1057.
  • [17] by same author, Controlling complex networks of linear systems via graphon limits, Presented at the Symposium of Controlling Complex Networks of NetSci17, Indianapolis, IN, USA (2017).
  • [18] by same author, Minimum energy control of arbitrary size networks of linear systems via graphon limits, Presented at the SIAM Workshop on Network Science, Pittsburgh, PA, USA (2017).
  • [19] by same author, Graphon linear quadratic regulation of large-scale networks of linear systems, Proc. 57th IEEE Conference on Decision and Control (Miami Beach, FL, USA), December 2018, pp. 5892–5897.
  • [20] by same author, Graphon control of large-scale networks of linear systems, IEEE Transactions on Automatic Control 65 (2020), no. 10, 4090–4105.
  • [21] Olivier Guéant, Existence and uniqueness result for mean field games with congestion effect on graphs, Applied Mathematics & Optimization 72 (2015), no. 2, 291–303.
  • [22] Isom H Herron and Michael R Foster, Partial differential equations in fluid dynamics, Cambridge University Press, 2008.
  • [23] Minyi Huang, Peter E Caines, and Roland P Malhamé, Large-population cost-coupled LQG problems with nonuniform agents: individual-mass behavior and decentralized ε\varepsilon-Nash equilibria, IEEE Transactions on Automatic Control 52 (2007), no. 9, 1560–1571.
  • [24] by same author, The NCE (mean field) principle with locality dependent cost interactions, IEEE Transactions on Automatic Control 55 (2010), no. 12, 2799–2805.
  • [25] Minyi Huang, Roland P Malhamé, and Peter E Caines, Large population stochastic dynamic games: closed-loop McKean-Vlasov systems and the Nash certainty equivalence principle, Communications in Information & Systems 6 (2006), no. 3, 221–252.
  • [26] Dmitry Kaliuzhnyi-Verbovetskyi and Georgi S Medvedev, The semilinear heat equation on sparse random graphs, SIAM Journal on Mathematical Analysis 49 (2017), no. 2, 1333–1355.
  • [27] Ioannis Karatzas and Steven Shreve, Brownian motion and stochastic calculus, vol. 113, Springer Science & Business Media, 2012.
  • [28] Vassili Kolokoltsov and Wei Yang, Sensitivity analysis for HJB equations with an application to coupled backward-forward systems, arXiv preprint arXiv:1303.6234v2 (2015).
  • [29] Olga Aleksandrovna Ladyzhenskaya, NN Ural’ceva, and VA Solonnikov, Linear and quasi-linear equations of parabolic type, American Mathematical Society, 1968.
  • [30] Jean-Michel Lasry and Pierre-Louis Lions, Jeux á champ moyen. I - le cas stationnaire, Comptes Rendus Mathématique 343 (2006), no. 9, 619–625.
  • [31] by same author, Jeux á champ moyen. II horizon fini et controle optimal, Comptes Rendus Mathématique 343 (2006), no. 10, 679–684.
  • [32] László Lovász, Large networks and graph limits, vol. 60, American Mathematical Soc., 2012.
  • [33] László Lovász and Balázs Szegedy, Limits of dense graph sequences, Journal of Combinatorial Theory, Series B 96 (2006), no. 6, 933–957.
  • [34] Georgi S Medvedev, The nonlinear heat equation on dense graphs and graph limits, SIAM J. Math. Anal. 46 (2014), no. 4, 2743–2766.
  • [35] by same author, The nonlinear heat equation on w-random graphs, Archive for Rational Mechanics and Analysis 212 (2014), no. 3, 781–803.
  • [36] I. P. Natanson, Theory of functions of a real variable, vol. I, F. Ungar Publishing Co., 1983, 5th printing.
  • [37] Francesca Parise and Asuman Ozdaglar, Graphon games, arXiv preprint arXiv:1802.00080 (2018).
  • [38] Wolfgang Pauli and Charles P Enz, Thermodynamics and the kinetic theory of gases, vol. 3, Courier Corporation, 2000.
  • [39] Cristobal Quininao and Jonathan Touboul, Limits and dynamics of randomly connected neuronal networks, Acta Appl Math 136 (2015), 167–192.
  • [40] Nevroz Sen and Peter E Caines, Mean field game theory with a partially observed major agent, SIAM Journal on Control and Optimization 54 (2016), no. 6, 3174–3224.
  • [41] Alain-Sol Sznitman, Topics in propagation of chaos, Ecole d’été de probabilités de Saint-Flour XIX—1989, Springer, 1991, pp. 165–251.