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

Quantitative relative entropy estimates on the whole space for convolution interaction forces

Paul Nikolaev Paul Nikolaev, University of Mannheim, Germany pnikolae@mail.uni-mannheim.de  and  David J. Prömel David J. Prömel, University of Mannheim, Germany proemel@uni-mannheim.de
Abstract.

Quantitative estimates are derived, on the whole space, for the relative entropy between the joint law of random interacting particles and the tensorized law at the limiting systeme. The developed method combines the relative entropy method under the moderated interaction scaling introduced by Oeschläger, and the propagation of chaos in probability. The result includes the case that the interaction force does not need to be a potential field. Furthermore, if the interaction force is a potential field, with a convolutional structure, then the developed estimate also provides the modulated energy estimates. Moreover, we demonstrate propagation of chaos for large stochastic systems of interacting particles and discuss possible applications to various interacting particle systems, including the Coulomb interaction case.

Key words: diffusion-aggregation equation, interacting particle systems, McKean–Vlasov equations, non-linear non-local PDE, propagation of chaos, relative entropy, modulated energy.

MSC 2010 Classification: 35D30, 35Q70, 60K35.

1. Introduction

In this article we study NN-particle systems 𝐗N=(X1,,XN)\mathbf{X}^{N}=(X^{1},\ldots,X^{N}) given by stochastic differential equations (SDEs) of the form

dXti=1Nj=1Nk(XtiXtj)dt+σdBti,i=1,,N,𝐗0Ni=1𝑁ρ0,\,\mathrm{d}X_{t}^{i}=-\frac{1}{N}\sum\limits_{j=1}^{N}k(X_{t}^{i}-X_{t}^{j})\,\mathrm{d}t+\sigma\,\mathrm{d}B_{t}^{i},\quad i=1,\ldots,N,\;\;\mathbf{X}_{0}^{N}\sim\overset{N}{\underset{i=1}{\otimes}}\rho_{0},

starting from i.i.d. initial data ρ0\rho_{0}. Such interacting systems arise naturally in various areas of science and engineering, including physics, chemistry, biology, ecology, and social sciences. For instance, they represent the behavior of ion channels, chemotaxis [KS70, HP09, Hor04], angiogenesis on the microscopic level and swarm movement [TBL06], flocking [HL09], opinion dynamics [Lor07, Hos20], cancer invasion [DTGC14] on the macroscopic level. The macroscopic level is often described through the evolution of the density of particles/individuals ρ\rho known to satisfy an aggregation-diffusion equation, which in general is a non-local, non-linear partial differential equation (PDE). Transitioning from microscopic models to continuum descriptions, i.e. NN approaches infinity, entails to explore the mean-field limit, see e.g. [Szn91, CCH14, JW16, Jab14]. It consists of demonstrating the convergence of the empirical measure μtN\mu_{t}^{N} for all t0t\geq 0, where μtN\mu_{t}^{N} is defined as

ωμtN(ω,A):=1Ni=1NδXti(ω)(A),A().\omega\mapsto\mu^{N}_{t}(\omega,A):=\frac{1}{N}\sum\limits_{i=1}^{N}\delta_{X^{i}_{t}(\omega)}(A),\quad\quad A\in\mathcal{B}(\mathbb{R}).

Various topologies are considered for the convergence, such as weak convergence, convergence in Wasserstein distance, convergence in terms of the Boltzmann entropy, and convergence in terms of the Fisher information. A comprehensive analysis can be found in [HM14]. In the present article we will focus on the convergence in entropy.

Main contribution: We present a novel method to derive propagation of chaos in entropy on the whole space for both non-conservative field and potential field possessing a convolution structure. Inspired by Oelschläger [Oel87], the presented method is based on the crucial observation that, under the convolution structure, the expectation of mollified L2L^{2} norm and the modulated energy (also as a weighted L2L^{2}-norm) can be estimated using the dynamics of the underlying systems in conjunction with the propagation of chaos in probability, as demonstrated in [LP17, HLP20, FHS19, CNP23]. The key contribution of the present work lies in the technique of combining propagation of chaos in probability [LP17, HLL19a, LY19, FHS19, HKPZ19, CCS19, HLP20, CLPY20, CNP23] with the underlying entropy structure from [JW16, Ser20, CH23] and the fluctuation estimates in [Oel87]. Consequently, we prove that convergence in probability for an interaction kernel, which is obtained by some type of mollification technique, implies convergence in relative entropy for an algebraic cut-off NβN^{-\beta}. This demonstrates that convergence in probability is actually a quite strong convergence result.

We emphasize that the main quantitative estimate, Theorem 3.3, is presented in a general manner and can easily be extended to a multi-dimensional setting, allowing its application to a wide range of kernels. We refer to Remark 3.4 for more details and to Section 5 for some interesting examples from the fields of chemotaxis and opinion dynamics. In particular, the method can be further applied in handling the attractive and repulsive Coulumb interaction potential in dimension d2d\geq 2, which includes the Keller–Segel model. Finally, we derive an estimate on the supremum norm in time of the relative entropy between the law of the approximated particle system and the chaotic law of the approximated mean-field SDE system of rate greater than 1/21/2. Moreover, the approximation is of algebraic order, which is sharper than the logarithmic cut-off derived from the standard coupling method [Szn91, LY19].

Theorem 3.3 can be considered as an intermediate result on the approximated level. On the one hand, the remaining limit of the regularized mean-field equation to the mean-field equation reduces to a question regarding the convergence on the PDE level. On the other hand, the convergence of the regularized particle system to the particles system is a question about the stability of solutions to the stochastic differential equations. For bounded interaction kernels, we also provide both convergences in the L1L^{1}-norm. Consequently, we prove the L1L^{1}-convergence of the mm-th marginal of the Liouville equation to the mm-th chaotic law of the non-linear diffusion-aggregation equation. This final convergence result is only presented for bounded kernels since, in general, the existence for the linear Liouville equation (2.6) on N\mathbb{R}^{N} is not given, see [BJW19, Proposition 4.2] for the torus setting.

Related literature: The study of propagation of chaos for a globally Lipschitz continuous interaction force kk has already a fairly long history, see e.g. [McK67, Szn91, HM14]. One of the first idea was to utilize the coupling method, i.e. comparing (Xti,t0)(X_{t}^{i},t\geq 0) with their associated McKean–Vlasov SDEs.

Motivated by models, particularly from physics, with bounded measurable or even singular interaction force kernels, extensive efforts have been devoted to investigated propagation of chaos for particle systems with such kernels. Initially, approaches to treat such irregular kernels were often based on compactness methods in combination with the martingale problems associated to the McKean–Vlasov SDEs, see e.g. [Oel84, Osa87, Gär88, FJ17, GQ15, LLY19, LLY19, ORT20]. For general LpL^{p}-interaction force kernels kk, the propagation of chaos was demonstrated for first and second order systems on the torus [BJS22] and on the whole space d\mathbb{R}^{d} [HRZ22, Han22, Lac23]. Another approach, initiated by Lazarovici and Pickl for the Vlasov–Poisson system [LP17], allows to deduce propagation of chaos in probability. This method is well-suited for singular interaction kernels, even when the underlying systems may not be well-defined.

For the moderate interacting system in deriving porous medium equation, Oeschläger has actually a series of contributions many decades ago, for example in [Oel84, Oel87]. Especially, for the fluctuation analysis, a smoothed L2L^{2} estimate with convergence rate o(N1/2)o(N^{-1/2}) has been obtained. The convolution structure of the moderate interaction played important roles. In the estimates proposed in [Oel87], the repulsive moderate interaction provides an essential quantity to absorb the rests from interacting effect. Recently, [Hol23] obtained Oeschläger’s L2L^{2} estimate for the moderate interacting system with attractive potential, under the assumption that the convergence of probability for the moderating interacting particle system holds, which is still an open problem. Recently [CH23], derived also a connection between the relative entropy and the regularized L2L^{2}-norm in the moderate interaction framework by directly citing the estimate from [Oel87]. The novelty of our work is that we do not follow the framework provided by [Oel87], but generate a direct estimation method in a general framework.

Another way to treat singular kernels such as the Coulomb potential x/|x|sx/|x|^{s} for s0s\geq 0 was investigated in the deterministic setting [Ser20, NRS22] (σ=0)\sigma=0) as well as in the random setting [JW18, BJW19, BJW20, RS23] (σ>0\sigma>0). The aforementioned references introduced the modulated free energy, which is a practical quantity suited for the Coulomb case. In particular, it metrize the weak convergence of the empirical measures [RS23]. A drawback of the modulated free energy approach in combination with the relative entropy is the torus domain as well as the requirement of entropy solutions on the particle level (microscopic level), see [BJW19, Proposition 4.2], which is non-trivial outside a setting on the torus. Furthermore, in order to apply the large deviation result in [JW18], strict conditions are required on the initial data and the solution of the Fokker–Planck equation. Recently, Wang and Feng extended these results to the 2D2D-viscous point vortex model on the whole space 2\mathbb{R}^{2}. The idea is to show exponential decay of the solution [FW23, Theorem 4.4] to be able to apply the large deviation result in [JW18]. Again strict restrictions on the initial conditions such as exponential decay of the initial data are necessary.

In the present article we manage to avoid the large deviation principle [JW18] and the strict conditions on the initial data by utilizing the convergence in probability, see (2.15) below. We also can treat general forces such as rotational fields or magnetic fields in physics. We also manage to derive quantitative bounds on singular forces such as attractive Coulomb interaction kernels on the whole space, which to our knowledge require approximation techniques by the nature of their singularities on the level of the Lioville equation. The price we pay lies in the obtained convergence rate. While [JW18] can establish the convergence in the sense of the Boltzmann entropy on the level of the associated Fokker–Planck equations with an order of N1N^{-1}, we achieve a rate of N1/2ϑN^{-1/2-\vartheta} for some ϑ>0\vartheta>0. Nevertheless, the convergence is faster than 1/21/2 and, therefore, we are optimistic that this result can be used as a stepping stone for Gaussian fluctuation.

Organization of the paper: In Section 2 we introduce the notation, the interacting particle systems and their associated diffusion-aggregation equations, give the necessary assumptions, and list the main results of this paper. We present the main ideas and the main estimate (Theorem 3.3) in Section 3. In Section 4, we demonstrate the propagation of chaos in the case of bounded interaction forces for the non-regularized systems by establishing the convergence of the approximated PDEs to the non-approximated counterparts. In Section 5, we showcase the applicability of the developed method by discussing, e.g., the regularized, singular Keller–Segel models and bounded confidence models.

Acknowledgments: P. Nikolaev and D. J. Prömel would like to deeply thank L. Chen for fruitful discussions and suggestions leading to a significant improvement of the present work.

2. Problem setting, preliminaries and main results

In this section we introduce the basic setting, the interacting particle systems, their associated partial differential equations, some preliminary results, as well we the main results of this article.

2.1. Particle systems

In this subsection we introduce the probabilistic setting, in particular, for the NN-particle system and the associated McKean–Vlasov equation. To that end, let (Ω,,(t)t0,)(\Omega,\mathcal{F},(\mathcal{F}_{t})_{t\geq 0},\mathbb{P}) be a complete probability space with right-continuous filtration (t)t0(\mathcal{F}_{t})_{t\geq 0} and (Bti,t0)(B_{t}^{i},t\geq 0), i=1,,Ni=1,\ldots,N, be independent one-dimensional Brownian motions with respect to (t,t0)(\mathcal{F}_{t},t\geq 0). In the following, we use the notation XρX\sim\rho to represent that ρ\rho is the law of random variable XX.

The NN-particle system 𝐗tN:=(Xt1,,XtN)\mathbf{X}_{t}^{N}:=(X_{t}^{1},\ldots,X_{t}^{N}) is given by

(2.1) dXti=1Nj=1Nk(XtiXtj)dt+σdBti,i=1,,N,𝐗0Ni=1𝑁ρ0,\,\mathrm{d}X_{t}^{i}=-\frac{1}{N}\sum\limits_{j=1}^{N}k(X_{t}^{i}-X_{t}^{j})\,\mathrm{d}t+\sigma\,\mathrm{d}B_{t}^{i},\quad i=1,\ldots,N,\;\;\mathbf{X}_{0}^{N}\sim\overset{N}{\underset{i=1}{\otimes}}\rho_{0},

where σ>0\sigma>0 is the diffusion parameter and 𝐗0N\mathbf{X}_{0}^{N} is independent of the Brownian motions (Bti,t0)(B_{t}^{i},t\geq 0), i=1,,Ni=1,\ldots,N. The particle system (2.1) induces in the limiting case NN\to\infty the following i.i.d. sequence 𝐘tN:=(Yt1,,YtN)\mathbf{Y}_{t}^{N}:=(Y_{t}^{1},\ldots,Y_{t}^{N}) of mean-field particles

(2.2) dYti=(kρt)(Yti)dt+σdBti,i=1,,N,𝐘0N=𝐗0N,\,\mathrm{d}Y_{t}^{i}=-(k*\rho_{t})(Y_{t}^{i})\,\mathrm{d}t+\sigma\,\mathrm{d}B_{t}^{i},\quad i=1,\ldots,N,\;\;\mathbf{Y}_{0}^{N}=\mathbf{X}_{0}^{N},

where ρt:=ρ(t,)\rho_{t}:=\rho(t,\cdot) denotes the probability density of the i.i.d. random variable YtiY_{t}^{i}.

To introduce the regularized versions of (2.1) and (2.2), we take the smooth approximation (kε,ε>0)(k^{\varepsilon},\varepsilon>0) of kk and replace the drift term with its approximation. Hence, the regularized microscopic NN-particle system 𝐗tN,ε:=(Xt1,ε,,XtN,ε)\mathbf{X}_{t}^{N,\varepsilon}:=(X_{t}^{1,\varepsilon},\ldots,X_{t}^{N,\varepsilon}) is given by

(2.3) dXti,ε=1Nj=1Nkε(Xti,εXtj,ε)dt+σdBti,i=1,,N,𝐗0N,εi=1𝑁ρ0,\mathrm{d}X_{t}^{i,\varepsilon}=-\frac{1}{N}\sum\limits_{j=1}^{N}k^{\varepsilon}(X_{t}^{i,\varepsilon}-X_{t}^{j,\varepsilon})\,\mathrm{d}t+\sigma\,\mathrm{d}B_{t}^{i},\quad i=1,\ldots,N,\;\;\mathbf{X}_{0}^{N,\varepsilon}\sim\overset{N}{\underset{i=1}{\otimes}}\rho_{0},

and the regularized mean-field trajectories 𝐘tN,ε:=(Yt1,ε,,YtN,ε)\mathbf{Y}_{t}^{N,\varepsilon}:=(Y_{t}^{1,\varepsilon},\ldots,Y_{t}^{N,\varepsilon}) by

(2.4) dYti,ε=(kερtε)(Yti,ε)dt+σdBti,i=1,,N,𝐘0N,ε=𝐗0N,ε,\mathrm{d}Y_{t}^{i,\varepsilon}=-(k^{\varepsilon}*\rho_{t}^{\varepsilon})(Y_{t}^{i,\varepsilon})\,\mathrm{d}t+\sigma\,\mathrm{d}B_{t}^{i},\quad i=1,\ldots,N,\;\;\mathbf{Y}_{0}^{N,\varepsilon}=\mathbf{X}_{0}^{N,\varepsilon},

where ρtε:=ρε(t,)\rho_{t}^{\varepsilon}:=\rho^{\varepsilon}(t,\cdot) denotes the probability density of the i.i.d. random variable Yti,εY_{t}^{i,\varepsilon}.

Finally, let the empirical measure of the regularized interaction system be given by

(2.5) μtN,ε(ω):=1Ni=1NδXti,ε(ω)𝒫(),\mu_{t}^{N,\varepsilon}(\omega):=\frac{1}{N}\sum\limits_{i=1}^{N}\delta_{X_{t}^{i,\varepsilon}(\omega)}\in\mathcal{P}(\mathbb{R}),

where δ\delta is the Dirac measure.

2.2. Associated PDEs

Itô’s formula implies that the associated probability densities of the particle systems, introduced in Subsection 2.1, satisfy partial differential equations (PDEs). Indeed, the interacting particle system (2.1) induces the following Liouville equation on N\mathbb{R}^{N},

(2.6) {tρtN(XN)=σ22i=1NxixiρtN(XN)+i=1Nxi(ρtN(XN)1Nj=1Nk(xixj))ρ0N(XN)=i=1Nρ0(xi)\displaystyle\begin{cases}\partial_{t}\rho_{t}^{N}(\mathrm{X}^{N})&=\frac{\sigma^{2}}{2}\sum\limits_{i=1}^{N}\partial_{x_{i}x_{i}}\rho_{t}^{N}(\mathrm{X}^{N})+\sum\limits_{i=1}^{N}\partial_{x_{i}}\Bigg{(}\rho^{N}_{t}(\mathrm{X}^{N})\frac{1}{N}\sum\limits_{j=1}^{N}k(x_{i}-x_{j})\Bigg{)}\\ \rho_{0}^{N}(\mathrm{X}^{N})&=\prod\limits_{i=1}^{N}\rho_{0}(x_{i})\end{cases}

for XN=(x1,,xN)N\mathrm{X}^{N}=(x_{1},\ldots,x_{N})\in\mathbb{R}^{N}, the system (2.2) induces the non-linear aggregation-diffusion equation

(2.7) {tρt=σ22xxρ+x(ρtkρt)(t,x)[0,T)×ρ(x,0)=ρ0x,\displaystyle\begin{cases}\partial_{t}\rho_{t}=\frac{\sigma^{2}}{2}\partial_{xx}\rho+\partial_{x}(\rho_{t}k*\rho_{t})\quad&\forall(t,x)\in[0,T)\times\mathbb{R}\\ \;\;\,\rho(x,0)=\rho_{0}&\forall x\in\mathbb{R}\end{cases},

the regularized particle system (2.3) the Liouville equation

(2.8) {tρtN,ε(XN)=σ22i=1NxixiρtN,ε(XN)+i=1Nxi(ρtN,ε(XN)1Nj=1Nkε(xixj))ρ0N,ε(XN)=i=1Nρ0(xi)\displaystyle\begin{cases}\partial_{t}\rho_{t}^{N,\varepsilon}(\mathrm{X}^{N})=\frac{\sigma^{2}}{2}\sum\limits_{i=1}^{N}\partial_{x_{i}x_{i}}\rho_{t}^{N,\varepsilon}(\mathrm{X}^{N})+\sum\limits_{i=1}^{N}\partial_{x_{i}}\Bigg{(}\rho^{N,\varepsilon}_{t}(\mathrm{X}^{N})\frac{1}{N}\sum\limits_{j=1}^{N}k^{\varepsilon}(x_{i}-x_{j})\Bigg{)}\\ \;\;\,\,\rho_{0}^{N,\varepsilon}(\mathrm{X}^{N})=\prod\limits_{i=1}^{N}\rho_{0}(x_{i})\end{cases}

and the regularized system  (2.4) the aggregation-diffusion equation

(2.9) {tρtε=σ22xxρtε+x(ρtεkερtε)(t,x)[0,T)×ρε(x,0)=ρ0x.\displaystyle\begin{cases}\partial_{t}\rho_{t}^{\varepsilon}=\frac{\sigma^{2}}{2}\partial_{xx}\rho_{t}^{\varepsilon}+\partial_{x}(\rho_{t}^{\varepsilon}k^{\varepsilon}*\rho_{t}^{\varepsilon})\quad&\forall(t,x)\in[0,T)\times\mathbb{R}\\ \;\;\,\rho^{\varepsilon}(x,0)=\rho_{0}&\forall x\in\mathbb{R}\end{cases}.

Note that we use ρt\rho_{t} and ρtε\rho_{t}^{\varepsilon} for the solutions of the PDEs (2.7) and (2.9) as well as for the probability densities of the particle systems (2.2) and (2.4), respectively, since these objects coincide by the superposition principle, see [BR20], in combination with existence results of densities for considered SDEs, see [Rom18].

Furthermore, we need to define the marginal of the system of rank 1mN1\leq m\leq N,

(2.10) ρtN,m=NmρtN(x1,,xN)dxm+1dxN.\rho_{t}^{N,m}=\int_{\mathbb{R}^{N-m}}\rho_{t}^{N}(x_{1},\ldots,x_{N})\,\mathrm{d}x_{m+1}\ldots\,\mathrm{d}x_{N}.

We remark that the mm-th martingale solves the following Liouville equation

(2.11) tρtN,m=σ22i=1mNmxixiρtN(x1,,xN)+xi(ρtN(x1,,xN)1Nj=1Nk(xixj))dxm+1dxN.\displaystyle\begin{split}\partial_{t}\rho_{t}^{N,m}=&\;\frac{\sigma^{2}}{2}\sum\limits_{i=1}^{m}\int_{\mathbb{R}^{N-m}}\partial_{x_{i}x_{i}}\rho_{t}^{N}(x_{1},\ldots,x_{N})\\ &+\partial_{x_{i}}\Bigg{(}\rho^{N}_{t}(x_{1},\ldots,x_{N})\frac{1}{N}\sum\limits_{j=1}^{N}k(x_{i}-x_{j})\Bigg{)}\,\mathrm{d}x_{m+1}\ldots\,\mathrm{d}x_{N}.\end{split}

Similar to (2.10) we denote by ρtN,m,ε\rho_{t}^{N,m,\varepsilon} the m-th marginal of the approximated Lioville equation, i.e.

ρtN,m,ε:=NmρtN,ε(x1,,xN)dxm+1dxN,\rho_{t}^{N,m,\varepsilon}:=\int_{\mathbb{R}^{N-m}}\rho_{t}^{N,\varepsilon}(x_{1},\ldots,x_{N})\,\mathrm{d}x_{m+1}\ldots\,\mathrm{d}x_{N},

which solves (2.11) with kεk^{\varepsilon} instead of kk. Additionally, we define the chaotic law

ρtm,ε(x1,,xm):=i=1mρtε(xi),\rho_{t}^{\otimes m,\varepsilon}(x_{1},\ldots,x_{m}):=\prod\limits_{i=1}^{m}\rho_{t}^{\varepsilon}(x_{i}),

which solves the following equation

tρtm,ε(x1,,xm)=σ22i=1mxixiρtm,ε(x1,,xm)+i=1mxi((kρtε)(xi)ρtm,ε(x1,,xm))\displaystyle\begin{split}&\;\partial_{t}\rho_{t}^{\otimes m,\varepsilon}(x_{1},\ldots,x_{m})\\ &\quad=\;\frac{\sigma^{2}}{2}\sum\limits_{i=1}^{m}\partial_{x_{i}x_{i}}\rho_{t}^{\otimes m,\varepsilon}(x_{1},\ldots,x_{m})+\sum\limits_{i=1}^{m}\partial_{x_{i}}\Big{(}(k*\rho_{t}^{\varepsilon})(x_{i})\rho_{t}^{\otimes m,\varepsilon}(x_{1},\ldots,x_{m})\Big{)}\end{split}

with initial data ρ0m,ε=ρ0m\rho_{0}^{\otimes m,\varepsilon}=\rho_{0}^{\otimes m}.

2.3. Preliminary results

In this subsection, we gather essential definitions, the requisite function spaces and preliminary results for the well-posedness of the above mentioned SDEs and PDEs.

Throughout the entire paper, we use the generic constant CC for inequalities, which may change from line to line. The constants α,βα,β\alpha,\beta_{\alpha},\beta are always fix and will be given by our Assumptions 2.52.6.

For 1p1\leq p\leq\infty we denote by Lp()L^{p}(\mathbb{R}) the space of measurable functions whose pp-th power is Lebesgue integrable (with the usual modification for p=p=\infty) equipped with the norm Lp()\left\lVert\cdot\right\rVert_{L^{p}(\mathbb{R})}, by L1(,|x|2dx)L^{1}(\mathbb{R},|x|^{2}\,\mathrm{d}x) the space of all measurable functions ff such that |f(x)||x|2dx<\int_{\mathbb{R}}|f(x)||x|^{2}\,\mathrm{d}x<\infty, by Cc()C_{c}^{\infty}(\mathbb{R}) the space of all infinitely differentiable functions with compact support on \mathbb{R}, and by 𝒮()\mathcal{S}(\mathbb{R}) the space of all Schwartz functions, see [Yos80, Chapter 6] for more details.

Let (Z,Z)(Z,\left\lVert\cdot\right\rVert_{Z}) be a Banach space. We denote by Lp([0,T];Z)L^{p}([0,T];Z) the space of all strongly measurable functions u:[0,T]Zu\colon[0,T]\to Z such that

uLp([0,T];Z):={(0Tu(t)Zpdt)1p<,for 1p<,esssupt[0,T]u(t)Z<,for p=.\left\lVert u\right\rVert_{L^{p}([0,T];Z)}:=\left\{\begin{array}[]{ll}\Big{(}\displaystyle\int_{0}^{T}\left\lVert u(t)\right\rVert_{Z}^{p}\,\mathrm{d}t\Big{)}^{\frac{1}{p}}<\infty,&\text{for }1\leq p<\infty,\\[14.22636pt] \displaystyle\operatorname*{ess\,sup}_{t\in[0,T]}\left\lVert u(t)\right\rVert_{Z}<\infty,&\text{for }p=\infty.\end{array}\right.

The Banach space C([0,T];Z)C([0,T];Z) consists of all continuous functions u:[0,T]Zu\colon[0,T]\to Z, equipped with the norm

maxt[0,T]u(t)Z<.\max\limits_{t\in[0,T]}\left\lVert u(t)\right\rVert_{Z}<\infty.

For a smooth function u:[0,T]×du\colon[0,T]\times\mathbb{R}^{d}\mapsto\mathbb{R} and a multiindex κ\kappa with length |κ|:=iκi|\kappa|:=\sum_{i}\kappa_{i}, we denote the derivative with respect to xκ=x1κ1xdκdx^{\kappa}=x_{1}^{\kappa_{1}}\cdots x_{d}^{\kappa_{d}} by κu(t,x):=i(xi)κiu(t,x)\partial^{\kappa}u(t,x):=\prod_{i}\big{(}\frac{\partial}{\partial_{x_{i}}}\big{)}^{\kappa_{i}}u(t,x), where we write xiu\partial_{x_{i}}u or uxi(t,x)u_{x_{i}}(t,x) for xiu(t,x)\frac{\partial}{\partial x_{i}}u(t,x). The derivative with respect to time we denote by tu(t,x)\partial_{t}u(t,x). For u𝒮(d)u\in\mathcal{S}^{\prime}(\mathbb{R}^{d}) we define the Fourier transform [u]\mathcal{F}[u] and inverse Fourier transform 1[u]\mathcal{F}^{-1}[u] by

[u](ξ):=de2πiηxu(x)dxand1[u](ξ):=de2πiηxu(x)dx.\mathcal{F}[u](\xi):=\int_{\mathbb{R}^{d}}e^{-2\pi i\eta\cdot x}u(x)\,\mathrm{d}x\quad\text{and}\quad\mathcal{F}^{-1}[u](\xi):=\int_{\mathbb{R}^{d}}e^{2\pi i\eta\cdot x}u(x)\,\mathrm{d}x.

We denote the Bessel potential for each ss\in\mathbb{R} and u𝒮(d)u\in\mathcal{S}^{\prime}(\mathbb{R}^{d}) by

(1Δ)s/2u:=1[(1+4π2|ξ|2)s/2[u]](1-\Delta)^{s/2}u:=\mathcal{F}^{-1}[(1+4\pi^{2}|\xi|^{2})^{s/2}\mathcal{F}[u]]

and define the Bessel potential space Hps\mathnormal{H}_{p}^{s} for p(1,)p\in(1,\infty) and ss\in\mathbb{R} by

Hps:={u𝒮(d):(1Δ)s/2uLp(d)}, with norm uHps(d):=(1Δ)s/2uLp(d)\mathnormal{H}_{p}^{s}:=\{u\in\mathcal{S}^{\prime}(\mathbb{R}^{d})\;:\;(1-\Delta)^{s/2}u\in L^{p}(\mathbb{R}^{d})\},\mbox{ with norm }\left\lVert u\right\rVert_{\mathnormal{H}^{s}_{p}(\mathbb{R}^{d})}:=\left\lVert(1-\Delta)^{s/2}u\right\rVert_{L^{p}(\mathbb{R}^{d})}

Applying [Tri83, Theorem 2.5.6] we can characterize the above Bessel potential spaces Hpm\mathnormal{H}_{p}^{m} for 1<p<1<p<\infty and mm\in\mathbb{N} as Sobolev spaces

Wm,p(d):={uLp(d):uWm,p(d):=κ𝒜,|α|mκfLp(d)<},W^{m,p}(\mathbb{R}^{d}):=\bigg{\{}u\in L^{p}(\mathbb{R}^{d})\;:\;\left\lVert u\right\rVert_{W^{m,p}(\mathbb{R}^{d})}:=\sum\limits_{\kappa\in\mathcal{A},\ |\alpha|\leq m}\left\lVert\partial^{\kappa}f\right\rVert_{L^{p}(\mathbb{R}^{d})}<\infty\bigg{\}},

where κu\partial^{\kappa}u is to be understood as weak derivatives [AF03] and 𝒜\mathcal{A} is the set of all multi-indices. Moreover, we will use the following abbreviation Hs(d):=H2s(d)H^{s}(\mathbb{R}^{d}):=H^{s}_{2}(\mathbb{R}^{d}).

For the partial differential equations (2.6), (2.7), (2.8) and (2.9) we rely on the concept of weak solutions, which we recall in the next definition.

Definition 2.1 (Weak solutions).

Fix a time T>0T>0. A function ρN,εL2([0,T];H1(N))L([0,T];L2(N))\rho^{N,\varepsilon}\in L^{2}([0,T];H^{1}(\mathbb{R}^{N}))\cap L^{\infty}([0,T];L^{2}(\mathbb{R}^{N})) with tρN,εL2([0,T];H1(N))\partial_{t}\rho^{N,\varepsilon}\in L^{2}([0,T];H^{-1}(\mathbb{R}^{N})) is a weak solution of (2.8) if for every ηL2([0,T];H1(N))\eta\in L^{2}([0,T];H^{1}(\mathbb{R}^{N})),

(2.12) 0TtρtN,ε,ηtH1(N),H1(N)dt\displaystyle\;\int\limits_{0}^{T}\langle\partial_{t}\rho^{N,\varepsilon}_{t},\eta_{t}\rangle_{H^{-1}(\mathbb{R}^{N}),H^{1}(\mathbb{R}^{N})}\,\mathrm{d}t
=\displaystyle= i=1N0TN(σ22xiρtN,ε(XN)+ρtN,ε(XN)1Nj=1Nkε(xixj)ρtN,ε(XN))xiηt(XN)dXNdt\displaystyle\;-\sum\limits_{i=1}^{N}\int\limits_{0}^{T}\int_{\mathbb{R}^{N}}\left(\frac{\sigma^{2}}{2}\partial_{x_{i}}\rho_{t}^{N,\varepsilon}(\mathrm{X}^{N})+\rho^{N,\varepsilon}_{t}(\mathrm{X}^{N})\frac{1}{N}\sum\limits_{j=1}^{N}k^{\varepsilon}(x_{i}-x_{j})\rho^{N,\varepsilon}_{t}(\mathrm{X}^{N})\right)\partial_{x_{i}}\eta_{t}(\mathrm{X}^{N})\,\mathrm{d}\mathrm{X}^{N}\,\mathrm{d}t

and ρN,ε(0,XN)=i=1Nρ0(xi)\rho^{N,\varepsilon}(0,\mathrm{X}^{N})=\prod\limits_{i=1}^{N}\rho_{0}(x_{i}). We note that the regularity ρN,εL2([0,T];H1(N))\rho^{N,\varepsilon}\in L^{2}([0,T];H^{1}(\mathbb{R}^{N})) and tρN,εL2([0,T];H1(N))\partial_{t}\rho^{N,\varepsilon}\in L^{2}([0,T];H^{-1}(\mathbb{R}^{N})) imply ρN,εC([0,T];L2(N))\rho^{N,\varepsilon}\in C([0,T];L^{2}(\mathbb{R}^{N})) (see for example [Eva15, Chapter 5.9]). Similarly, we say that ρNL2([0,T];H1(N))L([0,T];L2(N))\rho^{N}\in L^{2}([0,T];H^{1}(\mathbb{R}^{N}))\cap L^{\infty}([0,T];L^{2}(\mathbb{R}^{N})) with tρNL2([0,T];H1(N))\partial_{t}\rho^{N}\in L^{2}([0,T];H^{-1}(\mathbb{R}^{N})) is a weak solution of (2.6) if (2.12) hold with the interaction force kernel kk instead of its approximation kεk^{\varepsilon}.

Definition 2.2 (Weak solutions).

Fix ε>0\varepsilon>0 and T>0T>0. We say ρεL2([0,T];H1())L([0,T];L2())\rho^{\varepsilon}\in L^{2}([0,T];H^{1}(\mathbb{R}))\cap L^{\infty}([0,T];L^{2}(\mathbb{R})) with tρεL2([0,T];H1())\partial_{t}\rho^{\varepsilon}\in L^{2}([0,T];H^{-1}(\mathbb{R})) is a weak solution of (2.9) if, for every ηL2([0,T];H1())\eta\in L^{2}([0,T];H^{1}(\mathbb{R})),

(2.13) 0Ttρtε,ηtH1(),H1()dt=0T(σ22xρtε+(kερtε)ρtε)xηtdxdt\int\limits_{0}^{T}\langle\partial_{t}\rho^{\varepsilon}_{t},\eta_{t}\rangle_{H^{-1}(\mathbb{R}),H^{1}(\mathbb{R})}\,\mathrm{d}t=-\int\limits_{0}^{T}\int_{\mathbb{R}}\left(\frac{\sigma^{2}}{2}\partial_{x}\rho_{t}^{\varepsilon}+(k^{\varepsilon}*\rho_{t}^{\varepsilon})\rho_{t}^{\varepsilon}\right)\partial_{x}\eta_{t}\,\mathrm{d}x\,\mathrm{d}t

and ρε(0,)=ρ0\rho^{\varepsilon}(0,\cdot)=\rho_{0}. Note that ρεL2([0,T];H1())\rho^{\varepsilon}\in L^{2}([0,T];H^{1}(\mathbb{R})) with tρεL2([0,T];H1())\partial_{t}\rho^{\varepsilon}\in L^{2}([0,T];H^{-1}(\mathbb{R})) implies ρεC([0,T];L2())\rho^{\varepsilon}\in C([0,T];L^{2}(\mathbb{R})), see [Eva15, Chapter 5.9]. Similarly, we say that ρL2([0,T];H1())L([0,T];L2())\rho\in L^{2}([0,T];H^{1}(\mathbb{R}))\cap L^{\infty}([0,T];L^{2}(\mathbb{R})) with tρL2([0,T];H1())\partial_{t}\rho\in L^{2}([0,T];H^{-1}(\mathbb{R})) is a weak solution of (2.7) if (2.13) holds with the interaction force kernel kk instead of its approximation kεk^{\varepsilon}.

By the regularity of the solution in Definition 2.2 we can actually weaken the assumption on η\eta in equations (2.12) and (2.13) to ηC([0,T];Cc())\eta\in C([0,T];C_{c}^{\infty}(\mathbb{R})).

Remark 2.3.

The divergence structure of the PDEs (2.7) and (2.9), respectively, implies mass conservation/the normalisation condition

1=ρt(x)dx=ρtε(x)dx1=\int_{\mathbb{R}}\rho_{t}(x)\,\mathrm{d}x=\int_{\mathbb{R}}\rho^{\varepsilon}_{t}(x)\,\mathrm{d}x

for all 0tT0\leq t\leq T under Assumption 2.4. This is an immediate consequence by plugging in a cut-off sequence, see [Bre11, Lemma 8.4], which converges to the constant function 11 as a test function in (2.13).

Throughout the entire paper we make the following assumptions on the initial condition ρ0\rho_{0} of the interacting particle system and the interaction force kernel kk.

Assumption 2.4.

The initial condition ρ0:\rho_{0}\colon\mathbb{R}\to\mathbb{R} fulfills

(2.14) ρ0L1()L()L1(,|x|2dx),ρ00,andρ0(x)dx=1.\rho_{0}\in L^{1}(\mathbb{R})\cap L^{\infty}(\mathbb{R})\cap L^{1}(\mathbb{R},|x|^{2}\,\mathrm{d}x),\quad\rho_{0}\geq 0,\quad\text{and}\quad\int_{\mathbb{R}}\rho_{0}(x)\,\mathrm{d}x=1.

We recall some general facts, which will be used throughout the article. First, we notice that we have a solution (ρtN,ε,t0)(\rho_{t}^{N,\varepsilon},t\geq 0) of the regularized PDE (2.8) in the sense of Definition 2.1, which follows from the regularity of kεk^{\varepsilon}. We also have a solution (ρtN,t0)(\rho_{t}^{N},t\geq 0) in the sense of Definition 2.1 in the case kL()k\in L^{\infty}(\mathbb{R}) and the equation is linear. By standard SDE theory we also obtain strong solutions (𝐗tN,ε,t0)(\mathbf{X}_{t}^{N,\varepsilon},t\geq 0), (𝐘tN,ε,t0)(\mathbf{Y}_{t}^{N,\varepsilon},t\geq 0) to the regularized SDEs (2.3), (2.4). For the well-posedness of the particle system (2.1) and McKean–Vlasov SDE (2.2) we refer to [HRZ22, Theorem 3.7] and [HRZ22, Theorem 4.10], respectively. Additionally, [CNP23, Section 3] guarantees the well-posedness of PDEs (2.9), (2.7), which are bounded in time and space uniformly in ε\varepsilon. Consequently, our framework is well-defined and, in particular, the empirical measure μtN,ε\mu_{t}^{N,\varepsilon} given by (2.5) is well-defined.

The analysis of the entropy relies on the convergence of the particle system (2.3) to the particle system (2.4) in probability. Hence, we introduce the following convergence in probability assumption.

Assumption 2.5.

Let (𝐗tN,ε,t0)(\mathbf{X}_{t}^{N,\varepsilon},t\geq 0), (𝐘tN,ε,t0)(\mathbf{Y}_{t}^{N,\varepsilon},t\geq 0) be given by (2.3), (2.4). Then for α(0,1/2)\alpha\in(0,1/2), βα(0,α)\beta_{\alpha}\in(0,\alpha), ββα\beta\leq\beta_{\alpha}, εNβ\varepsilon\sim N^{-\beta} there exists an N0N_{0}\in\mathbb{N} such that for all NN0N\geq N_{0}, γ>0\gamma>0 we have

(2.15) (sup0tTsup1iN|Xti,εYti,ε|Nα)C(γ)Nγ,\mathbb{P}\left(\sup\limits_{0\leq t\leq T}\sup\limits_{1\leq i\leq N}|X_{t}^{i,\varepsilon}-Y_{t}^{i,\varepsilon}|\geq N^{-\alpha}\right)\leq C(\gamma)N^{-\gamma},

where C(γ)C(\gamma) depends on the initial density ρ0\rho_{0}, the final time T>0T>0, α\alpha and γ\gamma.

This assumptions is satisfied by a variety of models [LP17, HLL19a, LY19, FHS19, HKPZ19, CCS19, HLP20, CLPY20]. In particular for bounded kk or even singular kernels this assumption is fulfilled, see [CNP23].

Furthermore, we need the following law of large numbers result.

Assumption 2.6.

Let (𝐘tN,ε,t0)(\mathbf{Y}_{t}^{N,\varepsilon},t\geq 0) and ρtε\rho^{\varepsilon}_{t} be given by  (2.4). Assume further that 0<α,δ0<\alpha,\delta, 0<α+δ<1/20<\alpha+\delta<1/2, εNβ\varepsilon\sim N^{-\beta} with βα(0,α)\beta_{\alpha}\in(0,\alpha), ββα\beta\leq\beta_{\alpha} and define for 0tT0\leq t\leq T the following sets

Btα:={max1iN|j=1Nkε(Yti,εYtj,ε)(kρtε)(Yti,ε)|N(δ+α)}.B_{t}^{\alpha}:=\bigg{\{}\max\limits_{1\leq i\leq N}\bigg{|}\sum\limits_{j=1}^{N}k^{\varepsilon}(Y_{t}^{i,\varepsilon}-Y_{t}^{j,\varepsilon})-(k*\rho_{t}^{\varepsilon})(Y_{t}^{i,\varepsilon})\bigg{|}\leq N^{-(\delta+\alpha)}\bigg{\}}.

Then, for each γ>0\gamma>0 there exists a C(γ)>0C(\gamma)>0 such that

(2.16) ((Btα)c)C(γ)Nγ\mathbb{P}((B_{t}^{\alpha})^{\mathrm{c}})\leq C(\gamma)N^{-\gamma}

for every 0tT0\leq t\leq T, where the constant C(γ)C(\gamma) is independent of t[0,T]t\in[0,T].

We refer again to [LP17, HLL19a, LY19, FHS19, HKPZ19, CCS19, HLP20, CLPY20]. In particular, the assumption is satisfied for bounded forces kk, which satisfy a local Lipschitz bound [CNP23].

2.4. Main results:

Let Jε(x):=1εJ(xε)J^{\varepsilon}(x):=\frac{1}{\varepsilon}J\big{(}\frac{x}{\varepsilon}\big{)} with J:J\colon\mathbb{R}\to\mathbb{R} a given mollification kernel and let ζ\zeta be a cut-off function, which satisfies |ζ|1|\zeta|\leq 1, ζ=1\zeta=1 on B(0,1)B(0,1) and ζ=0\zeta=0 on B(0,2)cB(0,2)^{\mathrm{c}}, ζε(x)=ζ(εx)\zeta^{\varepsilon}(x)=\zeta(\varepsilon x). We need the following assumptions on the mollified version of interaction force separately to state the main result of this paper.

Definition 2.7.

We say Wε,VεW^{\varepsilon},V^{\varepsilon} are admissible approximations, if WεL2()W^{\varepsilon}\in L^{2}(\mathbb{R}) and VεH2()V^{\varepsilon}\in H^{2}(\mathbb{R}) with

(2.17) WεL2()CεaW,VεH2()CεaV\left\lVert W^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}\leq C\varepsilon^{-a_{W}},\quad\left\lVert V^{\varepsilon}\right\rVert_{H^{2}(\mathbb{R})}\leq C\varepsilon^{-a_{V}}

for some C>0C>0 and aW,aV>0a_{W},a_{V}>0. We say admissible approximations Wε,VεW^{\varepsilon},V^{\varepsilon} are strongly admissible approximations, if the above inequality holds for WεH2()\left\lVert W^{\varepsilon}\right\rVert_{H^{2}(\mathbb{R})} instead of WεL2()\left\lVert W^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}.

In general we will consider two type of forces. First, kε=WεVεk^{\varepsilon}=W^{\varepsilon}*V^{\varepsilon} and second kε=(WεVε)xk^{\varepsilon}=(W^{\varepsilon}*V^{\varepsilon})_{x}. The potential field structure of the latter one will be required for the definition of the modulated energy (see Section 3). This assumption on kk include many different forces, where no potential field is needed.

Remark 2.8.

Some typical examples for the above structure are as follows:

  1. (1)

    The interaction force kernel kL2()k\in L^{2}(\mathbb{R}). Then Wε=kW^{\varepsilon}=k and Vε=JεV^{\varepsilon}=J^{\varepsilon} is just the standard mollified version of kk.

  2. (2)

    If kLp()k\in L^{p}(\mathbb{R}) for p<p<\infty, we can choose Wε=kJεW^{\varepsilon}=k*J^{\varepsilon} and Vε=JεV^{\varepsilon}=J^{\varepsilon}, which is also just a mollification of kk.

  3. (3)

    If kL()k\in L^{\infty}(\mathbb{R}) we may choose Wε=ζε(kJε)W^{\varepsilon}=\zeta^{\varepsilon}(k*J^{\varepsilon}) and Vε=JεV^{\varepsilon}=J^{\varepsilon}, where ζε\zeta^{\varepsilon} is defined as a cut-off function to guarantee integrability of the mollification kJεk*J^{\varepsilon}.

The first main result of this paper is the propagation of chaos on the mollified level with ε=Nβ\varepsilon=N^{-\beta}:

Theorem 2.9.

Let ρN,ε\rho^{N,\varepsilon} and ρε\rho^{\varepsilon} be the non-negative solutions of (2.8) and of (2.9) respectively. Assume that the convergence in probability, Assumption 2.5, and the law of large numbers, Assumption 2.6 hold for α(14,12)\alpha\in(\frac{1}{4},\frac{1}{2}). Let kε=WεVεk^{\varepsilon}=W^{\varepsilon}*V^{\varepsilon} and WεL2(),VεH2()W^{\varepsilon}\in L^{2}(\mathbb{R}),V^{\varepsilon}\in H^{2}(\mathbb{R}) be admissible in the sense of Definition 2.7 with rate aW,aVa_{W},a_{V}. Then there exists a β1(0,βα)\beta_{1}\in(0,\beta_{\alpha}) such that β(0,β1)\forall\beta\in(0,\beta_{1}), the following propagation of chaos result holds for ε=Nβ\varepsilon=N^{-\beta} between (2.8) and of (2.9).

(2.18) ρtN,2,ερtερtεL1(2)222(ρtN,1,ε|ρtε)4N(ρtN,ε|ρtN,ε)=o(1N).\left\lVert\rho_{t}^{N,2,\varepsilon}-\rho_{t}^{\varepsilon}\otimes\rho_{t}^{\varepsilon}\right\rVert_{L^{1}(\mathbb{R}^{2})}^{2}\leq 2{\mathcal{H}}_{2}(\rho_{t}^{N,1,\varepsilon}|\rho_{t}^{\varepsilon})\leq 4{\mathcal{H}}_{N}(\rho_{t}^{N,\varepsilon}|\rho_{t}^{\otimes N,\varepsilon})=o\bigg{(}\frac{1}{\sqrt{N}}\bigg{)}.

where ρN,2,ε\rho^{N,2,\varepsilon} is the 22-marginal density of ρN,ε\rho^{N,\varepsilon}.

Furthermore, if kε=(WεVε)xk^{\varepsilon}=(W^{\varepsilon}*V^{\varepsilon})_{x} with Wε,VεW^{\varepsilon},V^{\varepsilon} being admissible approximations with the same rate aW,aVa_{W},a_{V} as before, then the estimate (2.18) still holds with β(0,β1)\beta\in(0,\beta_{1}). Moreover, if Wε,VεW^{\varepsilon},V^{\varepsilon} are strongly admissible, then there exists β2(0,βα)\beta_{2}\in(0,\beta_{\alpha}) such that β(0,β2)\forall\beta\in(0,\beta_{2}), the following estimate for regularized modulated energy holds with ε=Nβ\varepsilon=N^{-\beta} between (2.8) and of (2.9).

𝒦N(ρtN,ε|ρtN,ε)=1σ2𝔼(2(WεVε)(xy)d(μtN,ερtε)(x)d(μtN,ερtε)(y))=o(1N).\displaystyle{\mathcal{K}}_{N}(\rho_{t}^{N,\varepsilon}|\rho_{t}^{\otimes N,\varepsilon})=\frac{1}{\sigma^{2}}\mathbb{E}\bigg{(}\int_{\mathbb{R}^{2}}(W^{\varepsilon}*V^{\varepsilon})(x-y)\ \,\mathrm{d}(\mu_{t}^{N,\varepsilon}-\rho^{\varepsilon}_{t})(x)\,\,\mathrm{d}(\mu_{t}^{N,\varepsilon}-\rho^{\varepsilon}_{t})(y)\bigg{)}=o\bigg{(}\frac{1}{\sqrt{N}}\bigg{)}.
Remark 2.10.

In obtaining the estimate for smoothed modulated energy, the proof has been done with the identity

𝒦N(ρtN,ε|ρtN,ε)=1σ2𝔼(W^ε(μtN,ερtε),Vε(μtN,ερtε)),\displaystyle{\mathcal{K}}_{N}(\rho_{t}^{N,\varepsilon}|\rho_{t}^{\otimes N,\varepsilon})=\;\frac{1}{\sigma^{2}}\mathbb{E}\bigg{(}\Big{\langle}\hat{W}^{\varepsilon}*(\mu_{t}^{N,\varepsilon}-\rho_{t}^{\varepsilon}),V^{\varepsilon}*(\mu_{t}^{N,\varepsilon}-\rho_{t}^{\varepsilon})\Big{\rangle}\bigg{)},

where W^(x)=W(x)\hat{W}(x)=W(-x) is the reflection. Again choosing for instance Wε=JεW^{\varepsilon}=J^{\varepsilon} we may borrow an additional factor from the mollification kernel JεJ^{\varepsilon}, which will weaken the convergence rate estimate, or in other words, one has to choose even smaller β\beta to achieve the order o(1N)o(\frac{1}{\sqrt{N}}). The restriction α(14,12)\alpha\in(\frac{1}{4},\frac{1}{2}) is in place to guarantee the order o(1N)o(\frac{1}{\sqrt{N}}). The convergence of the relative entropy holds also without this restriction.

Additionally, for bounded force, we know from [CNP23] that convergence in probability holds for approximations (kε,ε>0(k^{\varepsilon},\varepsilon>0, which satisfy a local Lipschitz bound. Therefore, we can obtain the propagation of chaos result without mollification.

Theorem 2.11.

Assume that kL()k\in L^{\infty}(\mathbb{R}), the condition for initial data 2.4 holds. Suppose the Assumptions 2.52.6 hold for the approximation kε=(ζε(kJε))Jεk^{\varepsilon}=(\zeta^{\varepsilon}(k*J^{\varepsilon}))*J^{\varepsilon}. Then, for any fix mm\in\mathbb{N}, we have the convergence of the mm-th marginal of the Lioville equation (2.6) to the aggregation-diffusion equation (2.7) in the L1(m)L^{1}(\mathbb{R}^{m})-norm, i.e.

limNρN,mρmL1([0,T];L1(m))=0.\lim\limits_{N\to\infty}\left\lVert\rho^{N,m}-\rho^{\otimes m}\right\rVert_{L^{1}([0,T];L^{1}(\mathbb{R}^{m}))}=0.
Remark 2.12.

The Theorem holds for more general approximation kεk^{\varepsilon} as long as the approximation kεL2()k^{\varepsilon}\in L^{2}(\mathbb{R}) and the convergence in probability holds. We refer to [CNP23] for an overview of the topic of convergence in probability in the bounded case kL()k\in L^{\infty}(\mathbb{R}).

Let us finish the section with an overview over the constants:

  • α(0,1/2)\alpha\in(0,1/2) provides the rate on the distance of the particles in the convergence in probability

    sup1iN|Xti,εYti,ε|Nα\sup\limits_{1\leq i\leq N}|X_{t}^{i,\varepsilon}-Y_{t}^{i,\varepsilon}|\geq N^{-\alpha}

    and in the law of large numbers

    {max1iN|j=1Nkε(Yti,εYtj,ε)(kρtε)(Yti,ε)|N(δ+α)}.\bigg{\{}\max\limits_{1\leq i\leq N}\bigg{|}\sum\limits_{j=1}^{N}k^{\varepsilon}(Y_{t}^{i,\varepsilon}-Y_{t}^{j,\varepsilon})-(k*\rho_{t}^{\varepsilon})(Y_{t}^{i,\varepsilon})\bigg{|}\geq N^{-(\delta+\alpha)}\bigg{\}}.
  • βα(0,α)\beta_{\alpha}\in(0,\alpha) provides the maximum interval (0,βα)(0,\beta_{\alpha}) for the cut-off parameter β\beta, for which the convergence in probability and law of large numbers hold.

  • β\beta is the convergence rate of the approximated particles Xti,ε,Yti,εX_{t}^{i,\varepsilon},Y_{t}^{i,\varepsilon} such that ε=Nβ\varepsilon=N^{-\beta}.

  • β1,β2\beta_{1},\beta_{2} provide the maximum intervals (0,β1),(0,β2)(0,\beta_{1}),(0,\beta_{2}) such that the relative entropy and modulated energy converges with rate greater than 1/21/2, (see (2.18)).

3. Relative entropy method

This section is devoted to present the relative entropy method for the moderate interacting problem and its connection to the L2L^{2} estimate proposed by Oelschläger [Oel87]. We derive the smoothed L2L^{2} estimate for given force kk (no requirement as a potential field), and the smoothed modulated energy for potential field with convolution structure. Both lead to the estimate of the relative entropy between ρN,ε\rho^{N,\varepsilon} and ρN,ε\rho^{\otimes N,\varepsilon}.

The main idea is to use the assumption of convergence in probability (Assumption 2.5), the structure of the PDEs (2.8), (2.9) and the law of large number (Assumption 2.6). Applying the Csiszár–Kullback–Pinsker inequality [Vil09, Chapter 22] we provide an estimate on the L1()L^{1}(\mathbb{R})-norm of the marginals ρN,m,ε\rho^{N,m,\varepsilon} and ρm,ε\rho^{\otimes m,\varepsilon} for fix mm\in\mathbb{N}.

We emphasize that the method developed in Theorem 3.3 can be applied in different settings. Indeed, since we are working on the approximation level, our assumptions are only needed in the regularized setting. Hence, in general the assumptions on kk, VV and WW itself can be chosen more irregular, extending even to singular models. We refer to Remark 3.4 and the applications Section 5 for more details.

3.1. Relative entropy and modulated energy

In this section we introduce our main quantities the relative entropy and the modulated free energy. We then show the connection between the L2L^{2}-norm

(3.1) 𝔼(Vε(μtN,ερtε)L2()2),\mathbb{E}\bigg{(}\left\lVert V^{\varepsilon}*(\mu_{t}^{N,\varepsilon}-\rho_{t}^{\varepsilon})\right\rVert_{L^{2}(\mathbb{R})}^{2}\bigg{)},

the relative entropy N(ρtN,ε|ρtN,ε){\mathcal{H}}_{N}(\rho_{t}^{N,\varepsilon}|\rho_{t}^{\otimes N,\varepsilon}) as well as the expectation of the modulated free energy 𝒦N(ρtN,ε|ρtN,ε){\mathcal{K}}_{N}(\rho_{t}^{N,\varepsilon}|\rho_{t}^{\otimes N,\varepsilon}). This can be viewed as a combination of Oelschläger’s results on moderated interaction and fluctuations [Oel87] and the relative entropy method developed among others by Serfaty, Jabin, Wang, Bresch and Lacker [JW16, JW18, BJW19, BJW20, Ser20, NRS22, RS23, BJS22, Lac23] for the mean-field setting. The aim is to demonstrate how both concepts connect under the convolution assumption. Finally, we derive an estimate on the relative entropy in terms of the above L2L^{2}-norm.

Following [BJW19], we introduce the modulated free energy

EN(ρN,ε|ρN,ε):=N(ρN,ε|ρN,ε)+𝒦N(ρN,ε|ρN,ε),E_{N}\bigg{(}\rho^{N,\varepsilon}\,|\;\rho^{\otimes N,\varepsilon}\bigg{)}:={\mathcal{H}}_{N}(\rho^{N,\varepsilon}|\rho^{\otimes N,\varepsilon})+{\mathcal{K}}_{N}(\rho^{N,\varepsilon}|\rho^{\otimes N,\varepsilon}),

where

N(ρtN,ε|ρtN,ε):=1NNρtN,ε(x1,,xN)log(ρtN,ε(x1,,xN)ρtN,ε(x1,,xN))dx1,,xN{\mathcal{H}}_{N}(\rho_{t}^{N,\varepsilon}|\rho_{t}^{\otimes N,\varepsilon}):=\frac{1}{N}\int_{\mathbb{R}^{N}}\rho_{t}^{N,\varepsilon}(x_{1},\ldots,x_{N})\log\Bigl{(}\frac{\rho_{t}^{N,\varepsilon}(x_{1},\ldots,x_{N})}{\rho_{t}^{\otimes N,\varepsilon}(x_{1},\ldots,x_{N})}\Bigr{)}\,\,\mathrm{d}x_{1},\ldots,x_{N}

is the relative entropy introduced in [JW16] and if kε=(WεVε)k^{\varepsilon}=(W^{\varepsilon}*V^{\varepsilon}) is a potential

𝒦N(ρtN,ε|ρtN,ε):=1σ2𝔼(2(WεVε)(xy)d(μtN,ερtε)(x)d(μtN,ερtε)(y)){\mathcal{K}}_{N}(\rho_{t}^{N,\varepsilon}|\rho_{t}^{\otimes N,\varepsilon}):=\frac{1}{\sigma^{2}}\mathbb{E}\bigg{(}\int_{\mathbb{R}^{2}}(W^{\varepsilon}*V^{\varepsilon})(x-y)\ \,\mathrm{d}(\mu_{t}^{N,\varepsilon}-\rho^{\varepsilon}_{t})(x)\,\,\mathrm{d}(\mu_{t}^{N,\varepsilon}-\rho^{\varepsilon}_{t})(y)\bigg{)}

is the expectation of the modulated energy. We refer to [BJW19] and the references therein for more details on the modulated free energy.

Let us now explore some connections between the relative entropy and the structure presented by Oelschläger [Oel87]. We start by rewriting the expectation of the free energy by using our convolution structure. A straightforward calculation shows

(3.2) 𝒦N(ρtN,ε|ρtN,ε)=1σ2𝔼(W^ε(μtN,ερtε),Vε(μtN,ερtε)),\displaystyle{\mathcal{K}}_{N}(\rho_{t}^{N,\varepsilon}|\rho_{t}^{\otimes N,\varepsilon})=\;\frac{1}{\sigma^{2}}\mathbb{E}\bigg{(}\Big{\langle}\hat{W}^{\varepsilon}*(\mu_{t}^{N,\varepsilon}-\rho_{t}^{\varepsilon}),V^{\varepsilon}*(\mu_{t}^{N,\varepsilon}-\rho_{t}^{\varepsilon})\Big{\rangle}\bigg{)},

where W^(x)=W(x)\hat{W}(x)=W(-x) is the reflection. Applying Young’s inequality we see that it is enough to control a term of the form

𝔼(Vε(μtN,ερtε)L2()2)\displaystyle\mathbb{E}\bigg{(}\left\lVert V^{\varepsilon}*(\mu_{t}^{N,\varepsilon}-\rho_{t}^{\varepsilon})\right\rVert_{L^{2}(\mathbb{R})}^{2}\bigg{)}

for some function VεV^{\varepsilon}, where we just write VεV^{\varepsilon} for simplicity and understand that we can chose Vε=W^εV^{\varepsilon}=\hat{W}^{\varepsilon} in all calculations below. Hence, in order to estimate 𝒦N(ρtN,ε|ρtN,ε){\mathcal{K}}_{N}(\rho_{t}^{N,\varepsilon}|\rho_{t}^{\otimes N,\varepsilon}) we can estimate the L2L^{2}-difference between the convoluted empirical measure and the solution the law of the mean-field limit (2.2). This will be accomplished in Theorem 3.3.

But let us recall that our initial goal is to estimate the relative entropy N(ρtN,ε|ρtN,ε){\mathcal{H}}_{N}(\rho_{t}^{N,\varepsilon}|\rho_{t}^{\otimes N,\varepsilon}) and not 𝒦N(ρtN,ε|ρtN,ε){\mathcal{K}}_{N}(\rho_{t}^{N,\varepsilon}|\rho_{t}^{\otimes N,\varepsilon}). Therefore, let us connect the relative free energy to the L2L^{2}-norm of the derivative Vxε(μtN,ερtε)V^{\varepsilon}_{x}*(\mu_{t}^{N,\varepsilon}-\rho_{t}^{\varepsilon}).

Lemma 3.1.

Let Wε,VεW^{\varepsilon},V^{\varepsilon} be admissible and kε=WεVεk^{\varepsilon}=W^{\varepsilon}*V^{\varepsilon}. Then for the non-negative solutions ρN,ε\rho^{N,\varepsilon} of (2.8) and ρε\rho^{\varepsilon} of (2.9), it holds t>0\forall t>0 that

N(ρtN,ε|ρtN,ε)+σ24N0ti=1NN|xilog(ρsN,ε(XN)ρsN,ε(XN))|2ρsN,ε(XN)dXNds\displaystyle{\mathcal{H}}_{N}(\rho_{t}^{N,\varepsilon}|\rho_{t}^{\otimes N,\varepsilon})+\frac{\sigma^{2}}{4N}\int^{t}_{0}\sum\limits_{i=1}^{N}\int_{\mathbb{R}^{N}}\Bigg{|}\partial_{x_{i}}\log\Bigl{(}\frac{\rho_{s}^{N,\varepsilon}(\mathrm{X}^{N})}{\rho_{s}^{\otimes N,\varepsilon}(\mathrm{X}^{N})}\Bigr{)}\Bigg{|}^{2}\rho_{s}^{N,\varepsilon}(\mathrm{X}^{N})\,\mathrm{d}\mathrm{X}^{N}\,\mathrm{d}s
(3.3) WεL2()2σ2𝔼(0t(Vε(μsN,ερsε)L2()2)ds),\displaystyle\quad\leq\;\frac{\left\lVert W^{\varepsilon}\right\rVert^{2}_{L^{2}(\mathbb{R})}}{\sigma^{2}}\mathbb{E}\bigg{(}\int^{t}_{0}\bigg{(}\left\lVert V^{\varepsilon}*(\mu_{s}^{N,\varepsilon}-\rho_{s}^{\varepsilon})\right\rVert_{L^{2}(\mathbb{R})}^{2}\bigg{)}\,\mathrm{d}s\bigg{)},
Proof.

Let us compute the time derivative of the relative entropy

ddtN(ρtN,ε|ρtN,ε)\displaystyle\;\frac{\,\mathrm{d}}{\,\mathrm{d}t}{\mathcal{H}}_{N}(\rho_{t}^{N,\varepsilon}|\rho_{t}^{\otimes N,\varepsilon})
=\displaystyle= 1NNtρtN,ε(XN)log(ρtN,ε(XN)ρtN,ε(XN))+tρtN,ε(XN)ρtN,ε(XN)ρtN,ε(XN)tρtN,ε(XN)dXN\displaystyle\;\frac{1}{N}\int_{\mathbb{R}^{N}}\partial_{t}\rho_{t}^{N,\varepsilon}(\mathrm{X}^{N})\log\Bigl{(}\frac{\rho_{t}^{N,\varepsilon}(\mathrm{X}^{N})}{\rho_{t}^{\otimes N,\varepsilon}(\mathrm{X}^{N})}\Bigr{)}+\partial_{t}\rho_{t}^{N,\varepsilon}(\mathrm{X}^{N})-\frac{\rho_{t}^{N,\varepsilon}(\mathrm{X}^{N})}{\rho_{t}^{\otimes N,\varepsilon}(\mathrm{X}^{N})}\partial_{t}\rho_{t}^{\otimes N,\varepsilon}(\mathrm{X}^{N})\,\mathrm{d}\mathrm{X}^{N}
=\displaystyle= 1NN(σ22i=1NxixiρtN,ε(XN)+i=1Nxi(ρtN,ε(XN)1Nj=1Nkε(xixj)))log(ρtN,ε(XN)ρtN,ε(XN))\displaystyle\;\frac{1}{N}\int_{\mathbb{R}^{N}}\bigg{(}\frac{\sigma^{2}}{2}\sum\limits_{i=1}^{N}\partial_{x_{i}x_{i}}\rho_{t}^{N,\varepsilon}(\mathrm{X}^{N})+\sum\limits_{i=1}^{N}\partial_{x_{i}}\bigg{(}\rho^{N,\varepsilon}_{t}(\mathrm{X}^{N})\frac{1}{N}\sum\limits_{j=1}^{N}k^{\varepsilon}(x_{i}-x_{j})\bigg{)}\bigg{)}\log\Bigl{(}\frac{\rho_{t}^{N,\varepsilon}(\mathrm{X}^{N})}{\rho_{t}^{\otimes N,\varepsilon}(\mathrm{X}^{N})}\Bigr{)}
ρtN,ε(XN)ρtN,ε(XN)i=1Nσ22xixiρtN,ε(XN)i=1Nxi((kερtε)(xi)ρtN,ε(XN))dXN\displaystyle\quad-\frac{\rho_{t}^{N,\varepsilon}(\mathrm{X}^{N})}{\rho_{t}^{\otimes N,\varepsilon}(\mathrm{X}^{N})}\sum\limits_{i=1}^{N}\frac{\sigma^{2}}{2}\partial_{x_{i}x_{i}}\rho_{t}^{\otimes N,\varepsilon}(\mathrm{X}^{N})-\sum\limits_{i=1}^{N}\partial_{x_{i}}((k^{\varepsilon}*\rho_{t}^{\varepsilon})(x_{i})\rho_{t}^{\otimes N,\varepsilon}(\mathrm{X}^{N}))\,\mathrm{d}\mathrm{X}^{N}
=\displaystyle= σ22Ni=1NN|xilog(ρtN,ε(XN)ρtN,ε(XN))|2ρtN,ε(XN)dXN\displaystyle\;-\frac{\sigma^{2}}{2N}\sum\limits_{i=1}^{N}\int_{\mathbb{R}^{N}}\Bigg{|}\partial_{x_{i}}\log\Bigl{(}\frac{\rho_{t}^{N,\varepsilon}(\mathrm{X}^{N})}{\rho_{t}^{\otimes N,\varepsilon}(\mathrm{X}^{N})}\Bigr{)}\Bigg{|}^{2}\rho_{t}^{N,\varepsilon}(\mathrm{X}^{N})\,\mathrm{d}\mathrm{X}^{N}
1N2i,j=1NN(kε(xixj)kερtε(xi))ρtN,ε(XN)xilog(ρtN,ε(XN)ρtN,ε(XN))dXN\displaystyle-\frac{1}{N^{2}}\sum\limits_{i,j=1}^{N}\int_{\mathbb{R}^{N}}\Big{(}k^{\varepsilon}(x_{i}-x_{j})-k^{\varepsilon}*\rho_{t}^{\varepsilon}(x_{i})\Big{)}\rho^{N,\varepsilon}_{t}(\mathrm{X}^{N})\partial_{x_{i}}\log\Bigl{(}\frac{\rho_{t}^{N,\varepsilon}(\mathrm{X}^{N})}{\rho_{t}^{\otimes N,\varepsilon}(\mathrm{X}^{N})}\Bigr{)}\,\mathrm{d}\mathrm{X}^{N}
\displaystyle\leq σ24Ni=1NN|xilog(ρtN,ε(XN)ρtN,ε(XN))|2ρtN,ε(XN)dXN\displaystyle\;-\frac{\sigma^{2}}{4N}\sum\limits_{i=1}^{N}\int_{\mathbb{R}^{N}}\Bigg{|}\partial_{x_{i}}\log\Bigl{(}\frac{\rho_{t}^{N,\varepsilon}(\mathrm{X}^{N})}{\rho_{t}^{\otimes N,\varepsilon}(\mathrm{X}^{N})}\Bigr{)}\Bigg{|}^{2}\rho_{t}^{N,\varepsilon}(\mathrm{X}^{N})\,\mathrm{d}\mathrm{X}^{N}
+1σ2Ni=1NN|1Nj=1Nkε(xixj)kερtε(xi)|2ρtN,ε(XN)dXN\displaystyle+\frac{1}{\sigma^{2}N}\sum\limits_{i=1}^{N}\int_{\mathbb{R}^{N}}\bigg{|}\frac{1}{N}\sum\limits_{j=1}^{N}k^{\varepsilon}(x_{i}-x_{j})-k^{\varepsilon}*\rho_{t}^{\varepsilon}(x_{i})\bigg{|}^{2}\rho^{N,\varepsilon}_{t}(\mathrm{X}^{N})\,\mathrm{d}\mathrm{X}^{N}
\displaystyle\leq σ24Ni=1NN|xilog(ρtN,ε(XN)ρtN,ε(XN))|2ρtN,ε(XN)dXN+1σ2𝔼(μtN,ε,|kε(μtN,ερtε)|2).\displaystyle\;-\frac{\sigma^{2}}{4N}\sum\limits_{i=1}^{N}\int_{\mathbb{R}^{N}}\Bigg{|}\partial_{x_{i}}\log\Bigl{(}\frac{\rho_{t}^{N,\varepsilon}(\mathrm{X}^{N})}{\rho_{t}^{\otimes N,\varepsilon}(\mathrm{X}^{N})}\Bigr{)}\Bigg{|}^{2}\rho_{t}^{N,\varepsilon}(\mathrm{X}^{N})\,\mathrm{d}\mathrm{X}^{N}+\frac{1}{\sigma^{2}}\mathbb{E}\bigg{(}\langle\mu_{t}^{N,\varepsilon},|k^{\varepsilon}*(\mu_{t}^{N,\varepsilon}-\rho_{t}^{\varepsilon})|^{2}\rangle\bigg{)}.

For kε=WεVεk^{\varepsilon}=W^{\varepsilon}*V^{\varepsilon} we have further estimates

1σ2𝔼(μtN,ε,|kε(μtN,ερtε)|2)=1σ2𝔼(μtN,ε,|WεVε(μtN,ερtε)|2)\displaystyle\;\frac{1}{\sigma^{2}}\mathbb{E}\bigg{(}\langle\mu_{t}^{N,\varepsilon},|k^{\varepsilon}*(\mu_{t}^{N,\varepsilon}-\rho_{t}^{\varepsilon})|^{2}\rangle\bigg{)}=\frac{1}{\sigma^{2}}\mathbb{E}\bigg{(}\langle\mu_{t}^{N,\varepsilon},|W^{\varepsilon}*V^{\varepsilon}*(\mu_{t}^{N,\varepsilon}-\rho_{t}^{\varepsilon})|^{2}\rangle\bigg{)}
WεL2()2σ2𝔼(Vε(μtN,ερtε)L2()2)\displaystyle\quad\leq\;\frac{\left\lVert W^{\varepsilon}\right\rVert^{2}_{L^{2}(\mathbb{R})}}{\sigma^{2}}\mathbb{E}\bigg{(}\left\lVert V^{\varepsilon}*(\mu_{t}^{N,\varepsilon}-\rho_{t}^{\varepsilon})\right\rVert_{L^{2}(\mathbb{R})}^{2}\bigg{)}

Substituting the above estimate into the first inequality, while recalling that N(ρ0N,ε|ρ0N,ε)=0{\mathcal{H}}_{N}(\rho_{0}^{N,\varepsilon}|\rho_{0}^{\otimes N,\varepsilon})=0, proves the lemma. ∎

Remark 3.2.

Depending on the regularity of VεV^{\varepsilon} and WεW^{\varepsilon} one may choose to interchange the roles in the estimate. Generally, one should choose the more regular function to be VεV^{\varepsilon}. Indeed, in the above estimate we need only the L2L^{2}-norm of WεW^{\varepsilon}, while later on in Theorem 3.3 we need the LL^{\infty}-norm as well as the L2L^{2}-norm of not only the function VεV^{\varepsilon} but also of its derivatives. Moreover, if the force kεk^{\varepsilon} is a potential field, the last term has the following structure

𝔼(Vxε(μtN,ερtε)L2()2),\mathbb{E}\bigg{(}\left\lVert V^{\varepsilon}_{x}*(\mu_{t}^{N,\varepsilon}-\rho_{t}^{\varepsilon})\right\rVert_{L^{2}(\mathbb{R})}^{2}\bigg{)},

which will also be estimated by Theorem 3.3. Hence, we do not lose convergence rates in the case kε=WεVxεk^{\varepsilon}=W^{\varepsilon}*V^{\varepsilon}_{x}, but as already mentioned, we obtained an additional estimate on the modulated energy 𝒦N(ρtN,ε|ρtN,ε){\mathcal{K}}_{N}(\rho_{t}^{N,\varepsilon}|\rho_{t}^{\otimes N,\varepsilon}).

Consequently, by the above discussion, in order to control the relative entropy and the modulated energy in the case kεk^{\varepsilon} is a potential field, we need to find an estimate for the L2L^{2}-norm (3.1), which was studied in the moderated regime by Oelschläger [Oel87] nearly forty years ago.

3.2. L2L^{2}-estimate

In this section we concentrate on estimating the rest term in the entropy estimate (3.3).

We present the main theorem of the article, which is formulated for a function VεV^{\varepsilon}, which depends on ε\varepsilon. This presentation is motivation by our case kε=WεVεk^{\varepsilon}=W^{\varepsilon}*V^{\varepsilon}. We emphasize that the function in the following Theorem can be chosen independent of ε\varepsilon, but than the estimate has no connection to the modulated energy or the relative entropy (see Lemma 3.1).

Theorem 3.3.

Suppose (2.14), the convergence in probability, Assumption 2.5, and the law of large numbers, Assumption 2.6 hold both with rates β,βα,α\beta,\beta_{\alpha},\alpha specified therein. Then for any VεH2()V^{\varepsilon}\in H^{2}(\mathbb{R}) the following L2L^{2}-estimate holds

𝔼(sup0tTVεμtN,εVερtεL2()2)+σ28𝔼(0TVxεμsN,εVxερsεL2()2ds)\displaystyle\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\left\lVert V^{\varepsilon}*\mu_{t}^{N,\varepsilon}-V^{\varepsilon}*\rho_{t}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}\bigg{)}+\frac{\sigma^{2}}{8}\mathbb{E}\bigg{(}\int\limits_{0}^{T}\left\lVert V_{x}^{\varepsilon}*\mu_{s}^{N,\varepsilon}-V_{x}^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}\,\mathrm{d}s\bigg{)}
CN(VεH1()2kεL2+VxxεL2()2)+CVεH1()2(1+kεL()2)Nγ\displaystyle\quad\leq\;\frac{C}{N}(\left\lVert V^{\varepsilon}\right\rVert_{H^{1}(\mathbb{R})}^{2}\left\lVert k^{\varepsilon}\right\rVert_{L^{\infty}}^{2}+\left\lVert V^{\varepsilon}_{xx}\right\rVert_{L^{2}(\mathbb{R})}^{2})+\frac{C\left\lVert V^{\varepsilon}\right\rVert_{H^{1}(\mathbb{R})}^{2}(1+\left\lVert k^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}^{2})}{N^{\gamma}}
+kxεL()VεL2()+kεL()2VxεL2()2+VεL2()2N2α\displaystyle\quad\quad+\frac{\left\lVert k^{\varepsilon}_{x}\right\rVert_{L^{\infty}(\mathbb{R})}\left\lVert V^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}+\left\lVert k^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}^{2}\left\lVert V_{x}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}+\left\lVert V^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}}{N^{2\alpha}}
+CVxεL2()2(1+kxεL())+VxεL2()VxxεL2()kεL(d)Nα+12,\displaystyle\quad\quad+C\frac{\left\lVert V_{x}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}(1+\left\lVert k_{x}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})})+\left\lVert V_{x}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}\left\lVert V^{\varepsilon}_{xx}\right\rVert_{L^{2}(\mathbb{R})}\left\lVert k^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R}^{d})}}{N^{\alpha+\frac{1}{2}}},

where CC depends on TT, σ\sigma, γ\gamma, CBDGC_{\mathrm{BDG}}.

Remark 3.4.

The only ingredients we need for completing the proof of theorem 3.3 are the convergence in probability of the particle system 𝐗N,ε\mathbf{X}^{N,\varepsilon} to the mean-field limit 𝐘N,ε\mathbf{Y}^{N,\varepsilon} (2.15) as well as the law of large numbers (2.16). But the convergence in probability and the law of large numbers are known for a variety of interaction force kernels, see for instance [LP17, FHS19, HLL19b, HLP20]. Hence, this result can be extended for a variety of interaction force kernels. Moreover, the kernels can also be dd-dimensional since the estimates we used are dimension-free. We refer to Section 5 for applicable models such as the case with Coulomb force. Actually, the estimates become dimension dependent by the choice of β\beta. Consequently, the rate of convergence becomes dependent on the dimension. Nevertheless, the steps of the proof work analogously in multi-dimensional setting by replacing the multiplication with the scalar product, the absolute value with the Euclidean norm and the Itô’s formula with its multidimensional counter part.

Remark 3.5.

The results in Theorem 3.3 state that μtN,ε\mu_{t}^{N,\varepsilon} is close to ρtε\rho_{t}^{\varepsilon} in the mollified L2L^{2}-norm. By the propagation of chaos we expect that this quantity should be small since μtN,ερtε\mu_{t}^{N,\varepsilon}-\rho_{t}^{\varepsilon} should ideally vanish in the limit. The majority of work, which lies ahead, is to estimate this L2L^{2}-norm with a good rate. In the process we will also obtain an estimate on the derivative Vxε(μtN,ερtε)V^{\varepsilon}_{x}*(\mu_{t}^{N,\varepsilon}-\rho_{t}^{\varepsilon}). This is no surprise, since the estimate follows the structure of the classic a priori L2L^{2}-estimate for the parabolic equation [WYW06, Chapter 3]. As a result, we obtain in the L2()L^{2}(\mathbb{P})-norm an L([0,T];L2())L^{\infty}([0,T];L^{2}(\mathbb{R}))-bound and as usual an L2([0,T];L2())L^{2}([0,T];L^{2}(\mathbb{R}))-bound for the derivative. In combination with Lemma 3.1 this will allow us to obtain a bound on the relative entropy (ρtN,ε|ρtN,ε){\mathcal{H}}(\rho_{t}^{N,\varepsilon}|\rho_{t}^{\otimes N,\varepsilon}). Additionally, if the interaction force is a potential field we obtain an estimate for 𝒦N(ρtN,ε|ρtN,ε){\mathcal{K}}_{N}(\rho_{t}^{N,\varepsilon}|\rho_{t}^{\otimes N,\varepsilon}) by equality (3.2).

Let us start by describing the dynamic of the empirical measure μtN,ε\mu_{t}^{N,\varepsilon}. Applying Itô’s formula to a sufficiently smooth function ff, we obtain

f,μtN,ε=\displaystyle\langle f,\mu_{t}^{N,\varepsilon}\rangle= 1Ni=1Nf(Xti,ε)\displaystyle\;\frac{1}{N}\sum\limits_{i=1}^{N}f(X_{t}^{i,\varepsilon})
=\displaystyle= f,μ0N1Ni=1N1Nj=1N0tfx(Xsi,ε)k(Xsi,εXsj,ε)ds+σNi=1N0tfx(Xsi,ε)dBs\displaystyle\;\langle f,\mu^{N}_{0}\rangle-\frac{1}{N}\sum\limits_{i=1}^{N}\frac{1}{N}\sum\limits_{j=1}^{N}\int\limits_{0}^{t}f_{x}(X_{s}^{i,\varepsilon})k(X_{s}^{i,\varepsilon}-X_{s}^{j,\varepsilon})\,\mathrm{d}s+\frac{\sigma}{N}\sum\limits_{i=1}^{N}\int\limits_{0}^{t}f_{x}(X_{s}^{i,\varepsilon})\,\mathrm{d}B_{s}
+σ22Ni=1N0tfxx(Xsi,ε)ds.\displaystyle+\frac{\sigma^{2}}{2N}\sum\limits_{i=1}^{N}\int\limits_{0}^{t}f_{xx}(X_{s}^{i,\varepsilon})\,\mathrm{d}s.

Taking the expectation and using the fact that we have a density of 𝐗sN,ε\mathbf{X}_{s}^{N,\varepsilon}, provides a weak formulation of the Lioville equation (2.8). If we want to compare it to the mean-field law, we need to make the crucial observation that the stochastic integral in the above equation should vanish after taking the expectation. In other words, we have no term in the regularized PDE (2.9), which corresponds to the stochastic integral. If the integrand is smooth enough then obviously the stochastic integral vanishes. However, we need to compute the following difference

𝔼(sup0tTVεμtN,εVερtεL2()2).\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\left\lVert V^{\varepsilon}*\mu_{t}^{N,\varepsilon}-V^{\varepsilon}*\rho_{t}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}\bigg{)}.

Therefore, we need somehow transfer the naive approach to the more complex expected value. Applying the above dynamic we prove the following lemma, which allows us to treat the convolution VεμtN,εV^{\varepsilon}*\mu_{t}^{N,\varepsilon} as if the stochastic integral vanishes.

Lemma 3.6.

Let μtN,ε\mu_{t}^{N,\varepsilon} defined by (2.5). Then, we have the following inequality

𝔼(sup0tTVεμtN,εVερtεL2()2)\displaystyle\;\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\left\lVert V^{\varepsilon}*\mu_{t}^{N,\varepsilon}-V^{\varepsilon}*\rho_{t}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}\bigg{)}
 2𝔼(sup0tT|1Ni=1N(Vε(yX0i)+1Nj=1N0tVxε(yXsi,ε)kε(Xsi,εXsj,ε)ds\displaystyle\quad\leq\;2\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int_{\mathbb{R}}\bigg{|}\frac{1}{N}\sum\limits_{i=1}^{N}\bigg{(}V^{\varepsilon}(y-X_{0}^{i})+\frac{1}{N}\sum\limits_{j=1}^{N}\int\limits_{0}^{t}V^{\varepsilon}_{x}(y-X_{s}^{i,\varepsilon})k^{\varepsilon}(X_{s}^{i,\varepsilon}-X_{s}^{j,\varepsilon})\,\mathrm{d}s
+σ220tVxxε(yXti,ε)ds)Vερtε(y)|2dy)+2Tσ2CBDGNVxεL2()2.\displaystyle\quad\quad+\frac{\sigma^{2}}{2}\int\limits_{0}^{t}V^{\varepsilon}_{xx}(y-X_{t}^{i,\varepsilon})\,\mathrm{d}s\bigg{)}-V^{\varepsilon}*\rho_{t}^{\varepsilon}(y)\bigg{|}^{2}\,\mathrm{d}y\bigg{)}+\frac{2T\sigma^{2}C_{\mathrm{BDG}}}{N}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}.
Proof.

We use Itô’s formula, the dynamics (2.3) and the Burkholder–Davis–Gundy inequality to find

𝔼(sup0tTVεμtN,εVερtεL2()2)\displaystyle\;\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\left\lVert V^{\varepsilon}*\mu_{t}^{N,\varepsilon}-V^{\varepsilon}*\rho_{t}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}\bigg{)}
=\displaystyle= 𝔼(sup0tT|1Ni=1NVε(yXsi,ε)Vερtε(y)|2dy)\displaystyle\;\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int_{\mathbb{R}}\bigg{|}\frac{1}{N}\sum\limits_{i=1}^{N}V^{\varepsilon}(y-X_{s}^{i,\varepsilon})-V^{\varepsilon}*\rho_{t}^{\varepsilon}(y)\bigg{|}^{2}\,\mathrm{d}y\bigg{)}
=\displaystyle= 𝔼(sup0tT|1Ni=1N(Vε(yX0i)+1Nj=1N0tVxε(yXsi,ε)kε(Xsi,εXsj,ε)ds\displaystyle\;\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int_{\mathbb{R}}\bigg{|}\frac{1}{N}\sum\limits_{i=1}^{N}\bigg{(}V^{\varepsilon}(y-X_{0}^{i})+\frac{1}{N}\sum\limits_{j=1}^{N}\int\limits_{0}^{t}V^{\varepsilon}_{x}(y-X_{s}^{i,\varepsilon})k^{\varepsilon}(X_{s}^{i,\varepsilon}-X_{s}^{j,\varepsilon})\,\mathrm{d}s
+σ220tVxxε(yXti,ε)dsσ0tVxε(yXsi,ε)dBsi)Vερtε(y)|2dy)\displaystyle\quad+\frac{\sigma^{2}}{2}\int\limits_{0}^{t}V^{\varepsilon}_{xx}(y-X_{t}^{i,\varepsilon})\,\mathrm{d}s-\sigma\int\limits_{0}^{t}V^{\varepsilon}_{x}(y-X_{s}^{i,\varepsilon})\,\mathrm{d}B_{s}^{i}\bigg{)}-V^{\varepsilon}*\rho_{t}^{\varepsilon}(y)\bigg{|}^{2}\,\mathrm{d}y\bigg{)}
\displaystyle\leq  2𝔼(sup0tT|1Ni=1N(Vε(yX0i)+1Nj=1N0tVxε(yXsi,ε)kε(Xsi,εXsj,ε)ds\displaystyle\;2\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int_{\mathbb{R}}\bigg{|}\frac{1}{N}\sum\limits_{i=1}^{N}\bigg{(}V^{\varepsilon}(y-X_{0}^{i})+\frac{1}{N}\sum\limits_{j=1}^{N}\int\limits_{0}^{t}V^{\varepsilon}_{x}(y-X_{s}^{i,\varepsilon})k^{\varepsilon}(X_{s}^{i,\varepsilon}-X_{s}^{j,\varepsilon})\,\mathrm{d}s
+σ220tVxxε(yXti,ε)ds)Vερtε(y)|2dy)\displaystyle\quad+\frac{\sigma^{2}}{2}\int\limits_{0}^{t}V^{\varepsilon}_{xx}(y-X_{t}^{i,\varepsilon})\,\mathrm{d}s\bigg{)}-V^{\varepsilon}*\rho_{t}^{\varepsilon}(y)\bigg{|}^{2}\,\mathrm{d}y\bigg{)}
+2σ2𝔼(sup0tT|1Ni=1N0tVxε(yXsi,ε)dBsi|2dy).\displaystyle+2\sigma^{2}\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int_{\mathbb{R}}\bigg{|}\frac{1}{N}\sum\limits_{i=1}^{N}\int\limits_{0}^{t}V^{\varepsilon}_{x}(y-X_{s}^{i,\varepsilon})\,\mathrm{d}B_{s}^{i}\bigg{|}^{2}\,\mathrm{d}y\bigg{)}.

It remains to estimate the last term by the Burkholder–Davis–Gundy (BDG) inequality,

2σ2𝔼(sup0tT|1Ni=1N0tVxε(yXsi,ε)dBsi|2dy)\displaystyle 2\sigma^{2}\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int_{\mathbb{R}}\bigg{|}\frac{1}{N}\sum\limits_{i=1}^{N}\int\limits_{0}^{t}V^{\varepsilon}_{x}(y-X_{s}^{i,\varepsilon})\,\mathrm{d}B_{s}^{i}\bigg{|}^{2}\,\mathrm{d}y\bigg{)}
\displaystyle\leq  2σ2𝔼(sup0tT|1Ni=1N0tVxε(yXsi,ε)dBsi|2)dy\displaystyle\;2\sigma^{2}\int_{\mathbb{R}}\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\bigg{|}\frac{1}{N}\sum\limits_{i=1}^{N}\int\limits_{0}^{t}V^{\varepsilon}_{x}(y-X_{s}^{i,\varepsilon})\,\mathrm{d}B_{s}^{i}\bigg{|}^{2}\bigg{)}\,\mathrm{d}y
\displaystyle\leq  2σ2CBDG𝔼(1Ni=1N0Vxε(yXsi,ε)dBsiT)dy\displaystyle\;2\sigma^{2}C_{\mathrm{BDG}}\int_{\mathbb{R}}\mathbb{E}\bigg{(}\bigg{\langle}\frac{1}{N}\sum\limits_{i=1}^{N}\int\limits_{0}^{\cdot}V^{\varepsilon}_{x}(y-X_{s}^{i,\varepsilon})\,\mathrm{d}B_{s}^{i}\bigg{\rangle}_{T}\bigg{)}\,\mathrm{d}y
\displaystyle\leq 2σ2CBDGN2𝔼(i=1N0T|Vxε(yXsi,ε)|2ds)dy2Tσ2CBDGNVxεL2()2.\displaystyle\;\frac{2\sigma^{2}C_{\mathrm{BDG}}}{N^{2}}\int_{\mathbb{R}}\mathbb{E}\bigg{(}\sum\limits_{i=1}^{N}\int\limits_{0}^{T}|V^{\varepsilon}_{x}(y-X_{s}^{i,\varepsilon})|^{2}\,\mathrm{d}s\bigg{)}\,\mathrm{d}y\leq\;\frac{2T\sigma^{2}C_{\mathrm{BDG}}}{N}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}.

Inserting this calculation into the previous inequality proves the lemma. ∎

Proof of Theorem 3.3.

By Lemma 3.6 we can ignore the stochastic integral in the processes (Xi,ε,t0)(X^{i},\varepsilon,t\geq 0), which determine the empirical measure μtN,ε\mu_{t}^{N,\varepsilon}. Hence, let us write

Vε~μtN,ε(y)\displaystyle\;V^{\varepsilon}\tilde{*}\mu_{t}^{N,\varepsilon}(y)
:=\displaystyle:= 1Ni=1N(Vε(yX0i)+1Nj=1N0tVxε(yXsi,ε)kε(Xsi,εXsj,ε)ds+σ220tVxxε(yXti,ε)ds)\displaystyle\;\frac{1}{N}\sum\limits_{i=1}^{N}\bigg{(}V^{\varepsilon}(y-X_{0}^{i})+\frac{1}{N}\sum\limits_{j=1}^{N}\int\limits_{0}^{t}V^{\varepsilon}_{x}(y-X_{s}^{i,\varepsilon})k^{\varepsilon}(X_{s}^{i,\varepsilon}-X_{s}^{j,\varepsilon})\,\mathrm{d}s+\frac{\sigma^{2}}{2}\int\limits_{0}^{t}V^{\varepsilon}_{xx}(y-X_{t}^{i,\varepsilon})\,\mathrm{d}s\bigg{)}

for the convolution VεμtN,εV^{\varepsilon}*\mu_{t}^{N,\varepsilon} after applying Itô’s formula but without the stochastic integral. Then, we have

Vε~μtN,εVερtεL2()2\displaystyle\;\left\lVert V^{\varepsilon}\tilde{*}\mu_{t}^{N,\varepsilon}-V^{\varepsilon}*\rho_{t}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}
=\displaystyle= Vεμ0NVερ0L2()2+20ts(Vε~μsN,εVερsε),Vε~μsN,εVερsεL2()ds,\displaystyle\;\left\lVert V^{\varepsilon}*\mu^{N}_{0}-V^{\varepsilon}*\rho_{0}\right\rVert_{L^{2}(\mathbb{R})}^{2}+2\int\limits_{0}^{t}\langle\partial_{s}(V^{\varepsilon}\tilde{*}\mu_{s}^{N,\varepsilon}-V^{\varepsilon}*\rho_{s}^{\varepsilon}),V^{\varepsilon}\tilde{*}\mu_{s}^{N,\varepsilon}-V^{\varepsilon}*\rho_{s}^{\varepsilon}\rangle_{L^{2}(\mathbb{R})}\,\mathrm{d}s,

where we notice that for the initial time t=0t=0, we have Vε~μ0N=Vεμ0NV^{\varepsilon}\tilde{*}\mu_{0}^{N}=V^{\varepsilon}*\mu_{0}^{N} by definition. Let us remark that since all integrands are smooth enough we have (Vε~μtN,ε)x=Vxε~μtN,ε(V^{\varepsilon}\tilde{*}\mu_{t}^{N,\varepsilon})_{x}=V^{\varepsilon}_{x}\tilde{*}\mu_{t}^{N,\varepsilon}. Next, plugging in Vε~μsN,εV^{\varepsilon}\tilde{*}\mu_{s}^{N,\varepsilon} and differentiate we obtain

sVε~μsN,ε,Vε~μsN,εVερsεL2()\displaystyle\;\langle\partial_{s}V^{\varepsilon}\tilde{*}\mu_{s}^{N,\varepsilon},V^{\varepsilon}\tilde{*}\mu_{s}^{N,\varepsilon}-V^{\varepsilon}*\rho_{s}^{\varepsilon}\rangle_{L^{2}(\mathbb{R})}
=\displaystyle= 1Ni=1N1Nj=1NVxε(Xsi,ε)kε(Xsi,εXsj,ε)\displaystyle\;\bigg{\langle}\frac{1}{N}\sum\limits_{i=1}^{N}\frac{1}{N}\sum\limits_{j=1}^{N}V^{\varepsilon}_{x}(\cdot-X_{s}^{i,\varepsilon})k^{\varepsilon}(X_{s}^{i,\varepsilon}-X_{s}^{j,\varepsilon})
+σ22Ni=1NVxxε(Xsi,ε),Vε~μsN,εVερsεL2()\displaystyle\quad+\frac{\sigma^{2}}{2N}\sum\limits_{i=1}^{N}V^{\varepsilon}_{xx}(\cdot-X_{s}^{i,\varepsilon}),V^{\varepsilon}\tilde{*}\mu_{s}^{N,\varepsilon}-V^{\varepsilon}*\rho_{s}^{\varepsilon}\bigg{\rangle}_{L^{2}(\mathbb{R})}
=\displaystyle= 1Ni=1N1Nj=1NVxε(Xsi,ε)kε(Xsi,εXsj,ε),Vε~μsN,εVερsεL2()\displaystyle\;\bigg{\langle}\frac{1}{N}\sum\limits_{i=1}^{N}\frac{1}{N}\sum\limits_{j=1}^{N}V^{\varepsilon}_{x}(\cdot-X_{s}^{i,\varepsilon})k^{\varepsilon}(X_{s}^{i,\varepsilon}-X_{s}^{j,\varepsilon}),V^{\varepsilon}\tilde{*}\mu_{s}^{N,\varepsilon}-V^{\varepsilon}*\rho_{s}^{\varepsilon}\bigg{\rangle}_{L^{2}(\mathbb{R})}
σ22VxεμsN,ε,Vxε~μsN,εVxερsεL2().\displaystyle-\frac{\sigma^{2}}{2}\langle V^{\varepsilon}_{x}*\mu_{s}^{N,\varepsilon},V^{\varepsilon}_{x}\tilde{*}\mu_{s}^{N,\varepsilon}-V^{\varepsilon}_{x}*\rho_{s}^{\varepsilon}\rangle_{L^{2}(\mathbb{R})}.

Similar, ρsε\rho_{s}^{\varepsilon} is a weak solution to our PDE (2.9), which implies

sVερsε),Vε~μsN,εVερsεL2()\displaystyle\;\langle\partial_{s}V^{\varepsilon}*\rho_{s}^{\varepsilon}),V^{\varepsilon}\tilde{*}\mu_{s}^{N,\varepsilon}-V^{\varepsilon}*\rho_{s}^{\varepsilon}\rangle_{L^{2}(\mathbb{R})}
=\displaystyle= Vε(σ22(ρsε)xx+((kερsε)ρsε)x),Vε~μsN,εVερsεL2()\displaystyle\;\bigg{\langle}V^{\varepsilon}*\bigg{(}\frac{\sigma^{2}}{2}(\rho_{s}^{\varepsilon})_{xx}+((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon})_{x}\bigg{)},V^{\varepsilon}\tilde{*}\mu_{s}^{N,\varepsilon}-V^{\varepsilon}*\rho_{s}^{\varepsilon}\bigg{\rangle}_{L^{2}(\mathbb{R})}
=\displaystyle= σ22Vε(ρsε)x,Vxε~μsN,εVxερsεL2()\displaystyle\;-\frac{\sigma^{2}}{2}\langle V^{\varepsilon}*(\rho_{s}^{\varepsilon})_{x},V^{\varepsilon}_{x}\tilde{*}\mu_{s}^{N,\varepsilon}-V^{\varepsilon}_{x}*\rho_{s}^{\varepsilon}\rangle_{L^{2}(\mathbb{R})}
+Vε((kερsε)ρsε)x,Vε~μsN,εVερsεL2().\displaystyle+\langle V^{\varepsilon}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon})_{x},V^{\varepsilon}\tilde{*}\mu_{s}^{N,\varepsilon}-V^{\varepsilon}*\rho_{s}^{\varepsilon}\rangle_{L^{2}(\mathbb{R})}.

Combing the last two calculations, we find

Vε~μtN,εVερtεL2()2\displaystyle\;\left\lVert V^{\varepsilon}\tilde{*}\mu_{t}^{N,\varepsilon}-V^{\varepsilon}*\rho_{t}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}
=\displaystyle= Vεμ0NVερ0L2()220tσ22VxεμsN,εVxερsε,Vxε~μsN,εVxερsεL2()ds\displaystyle\;\left\lVert V^{\varepsilon}*\mu^{N}_{0}-V^{\varepsilon}*\rho_{0}\right\rVert_{L^{2}(\mathbb{R})}^{2}-2\int\limits_{0}^{t}\frac{\sigma^{2}}{2}\langle V^{\varepsilon}_{x}*\mu_{s}^{N,\varepsilon}-V^{\varepsilon}_{x}*\rho_{s}^{\varepsilon},V^{\varepsilon}_{x}\tilde{*}\mu_{s}^{N,\varepsilon}-V^{\varepsilon}_{x}*\rho_{s}^{\varepsilon}\rangle_{L^{2}(\mathbb{R})}\,\mathrm{d}s
+0t2N2i,j=1NVxε(Xsi,ε)kε(Xsi,εXsj,ε)Vxε((kερsε)ρsε),Vε~μsN,εVερsεL2()ds.\displaystyle+\int\limits_{0}^{t}\bigg{\langle}\frac{2}{N^{2}}\sum\limits_{i,j=1}^{N}V^{\varepsilon}_{x}(\cdot-X_{s}^{i,\varepsilon})k^{\varepsilon}(X_{s}^{i,\varepsilon}-X_{s}^{j,\varepsilon})-V^{\varepsilon}_{x}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon}),V^{\varepsilon}\tilde{*}\mu_{s}^{N,\varepsilon}-V^{\varepsilon}*\rho_{s}^{\varepsilon}\bigg{\rangle}_{L^{2}(\mathbb{R})}\,\mathrm{d}s.

The goal is now to insert VεμsN,εV^{\varepsilon}*\mu_{s}^{N,\varepsilon} back into the equation. Hence, for the absorption term we have

0tσ22VxεμsN,εVxερsε,Vxε~μsN,εVxερsεL2()ds\displaystyle\;-\int\limits_{0}^{t}\frac{\sigma^{2}}{2}\langle V^{\varepsilon}_{x}*\mu_{s}^{N,\varepsilon}-V^{\varepsilon}_{x}*\rho_{s}^{\varepsilon},V^{\varepsilon}_{x}\tilde{*}\mu_{s}^{N,\varepsilon}-V^{\varepsilon}_{x}*\rho_{s}^{\varepsilon}\rangle_{L^{2}(\mathbb{R})}\,\mathrm{d}s
=\displaystyle= 0tσ22VxεμsN,εVxερsεL2()2ds\displaystyle\;-\int\limits_{0}^{t}\frac{\sigma^{2}}{2}\left\lVert V^{\varepsilon}_{x}*\mu_{s}^{N,\varepsilon}-V^{\varepsilon}_{x}*\rho_{s}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}\,\mathrm{d}s
+0tσ22VxεμsN,εVxερsε,σNi=1N0sVxxε(Xui)dBuL2()ds\displaystyle+\int\limits_{0}^{t}\frac{\sigma^{2}}{2}\bigg{\langle}V^{\varepsilon}_{x}*\mu_{s}^{N,\varepsilon}-V^{\varepsilon}_{x}*\rho_{s}^{\varepsilon},\frac{\sigma}{N}\sum\limits_{i=1}^{N}\int\limits_{0}^{s}V^{\varepsilon}_{xx}(\cdot-X_{u}^{i})\,\mathrm{d}B_{u}\bigg{\rangle}_{L^{2}(\mathbb{R})}\,\mathrm{d}s
\displaystyle\leq 0tσ22VxεμsN,εVxερsεL2()2ds+0tσ216VxεμsN,εVxερsεL2()2\displaystyle\;-\int\limits_{0}^{t}\frac{\sigma^{2}}{2}\left\lVert V^{\varepsilon}_{x}*\mu_{s}^{N,\varepsilon}-V^{\varepsilon}_{x}*\rho_{s}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}\,\mathrm{d}s+\int\limits_{0}^{t}\frac{\sigma^{2}}{16}\left\lVert V^{\varepsilon}_{x}*\mu_{s}^{N,\varepsilon}-V^{\varepsilon}_{x}*\rho_{s}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}
+2σ2σNi=1N0sVxxε(Xui)dBuL2()2ds\displaystyle+2\sigma^{2}\left\lVert\frac{\sigma}{N}\sum\limits_{i=1}^{N}\int\limits_{0}^{s}V^{\varepsilon}_{xx}(\cdot-X_{u}^{i})\,\mathrm{d}B_{u}\right\rVert_{L^{2}(\mathbb{R})}^{2}\,\mathrm{d}s
=\displaystyle= 0t7σ216VxεμsN,εVxερsεL2()2ds+2σ20tσNi=1N0sVxxε(Xui)dBuL2()2ds\displaystyle\;-\int\limits_{0}^{t}\frac{7\sigma^{2}}{16}\left\lVert V^{\varepsilon}_{x}*\mu_{s}^{N,\varepsilon}-V^{\varepsilon}_{x}*\rho_{s}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}\,\mathrm{d}s+2\sigma^{2}\int\limits_{0}^{t}\left\lVert\frac{\sigma}{N}\sum\limits_{i=1}^{N}\int\limits_{0}^{s}V^{\varepsilon}_{xx}(\cdot-X_{u}^{i})\,\mathrm{d}B_{u}\right\rVert_{L^{2}(\mathbb{R})}^{2}\,\mathrm{d}s

and for the last term

𝔼(sup0tT0t1N2i,j=1NVxε(Xsi,ε)kε(Xsi,εXsj,ε)Vxε((kερsε)ρsε),\displaystyle\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int\limits_{0}^{t}\bigg{\langle}\frac{1}{N^{2}}\sum\limits_{i,j=1}^{N}V^{\varepsilon}_{x}(\cdot-X_{s}^{i,\varepsilon})k^{\varepsilon}(X_{s}^{i,\varepsilon}-X_{s}^{j,\varepsilon})-V^{\varepsilon}_{x}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon}),
Vε~μsN,εVερsεL2()ds)\displaystyle\quad V^{\varepsilon}\tilde{*}\mu_{s}^{N,\varepsilon}-V^{\varepsilon}*\rho_{s}^{\varepsilon}\bigg{\rangle}_{L^{2}(\mathbb{R})}\,\mathrm{d}s\bigg{)}
\displaystyle\leq 𝔼(sup0tT0t|1N2i,j=1NVxε(Xsi,ε)kε(Xsi,εXsj,ε)Vxε((kερsε)ρsε),\displaystyle\,\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int\limits_{0}^{t}\bigg{|}\bigg{\langle}\frac{1}{N^{2}}\sum\limits_{i,j=1}^{N}V^{\varepsilon}_{x}(\cdot-X_{s}^{i,\varepsilon})k^{\varepsilon}(X_{s}^{i,\varepsilon}-X_{s}^{j,\varepsilon})-V^{\varepsilon}_{x}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon}),
VεμsN,εVερsεL2()|ds)+𝔼(sup0tT0t|1N2i,j=1NVxε(Xsi,ε)kε(Xsi,εXsj,ε)\displaystyle\quad V^{\varepsilon}*\mu_{s}^{N,\varepsilon}-V^{\varepsilon}*\rho_{s}^{\varepsilon}\bigg{\rangle}_{L^{2}(\mathbb{R})}\bigg{|}\,\mathrm{d}s\bigg{)}+\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int\limits_{0}^{t}\bigg{|}\bigg{\langle}\frac{1}{N^{2}}\sum\limits_{i,j=1}^{N}V^{\varepsilon}_{x}(\cdot-X_{s}^{i,\varepsilon})k^{\varepsilon}(X_{s}^{i,\varepsilon}-X_{s}^{j,\varepsilon})
Vxε((kερsε)ρsε),σNl=1N0sVxε(Xul)dBulL2()|ds).\displaystyle\quad-V^{\varepsilon}_{x}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon}),\frac{\sigma}{N}\sum\limits_{l=1}^{N}-\int\limits_{0}^{s}V^{\varepsilon}_{x}(\cdot-X_{u}^{l})\,\mathrm{d}B_{u}^{l}\bigg{\rangle}_{L^{2}(\mathbb{R})}\bigg{|}\,\mathrm{d}s\bigg{)}.

Applying Lemma 3.6 and put together the above estimates we have shown

(3.4) 𝔼(sup0tTVεμtN,εVερtεL2()2)\displaystyle\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\left\lVert V^{\varepsilon}*\mu_{t}^{N,\varepsilon}-V^{\varepsilon}*\rho_{t}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}\bigg{)}
\displaystyle\leq 𝔼(sup0tTVε~μtN,εVερtεL2()2)+2Tσ2CBDGNVxεL2()2\displaystyle\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\left\lVert V^{\varepsilon}\tilde{*}\mu_{t}^{N,\varepsilon}-V^{\varepsilon}*\rho_{t}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}\bigg{)}+\frac{2T\sigma^{2}C_{\mathrm{BDG}}}{N}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}
\displaystyle\leq  2𝔼(sup0tT(Vεμ0NVερ0L2()20t7σ216VxεμsN,εVxερsεL2()2ds\displaystyle\;2\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\bigg{(}\left\lVert V^{\varepsilon}*\mu^{N}_{0}-V^{\varepsilon}*\rho_{0}\right\rVert_{L^{2}(\mathbb{R})}^{2}-\int\limits_{0}^{t}\frac{7\sigma^{2}}{16}\left\lVert V^{\varepsilon}_{x}*\mu_{s}^{N,\varepsilon}-V^{\varepsilon}_{x}*\rho_{s}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}\,\mathrm{d}s
+0t|1N2i,j=1NVxε(Xsi,ε)kε(Xsi,εXsj,ε)Vxε((kερsε)ρsε),\displaystyle+\int\limits_{0}^{t}\bigg{|}\bigg{\langle}\frac{1}{N^{2}}\sum\limits_{i,j=1}^{N}V^{\varepsilon}_{x}(\cdot-X_{s}^{i,\varepsilon})k^{\varepsilon}(X_{s}^{i,\varepsilon}-X_{s}^{j,\varepsilon})-V^{\varepsilon}_{x}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon}),
VεμsN,εVερsεL2()|ds))\displaystyle\quad V^{\varepsilon}*\mu_{s}^{N,\varepsilon}-V^{\varepsilon}*\rho_{s}^{\varepsilon}\bigg{\rangle}_{L^{2}(\mathbb{R})}\bigg{|}\,\mathrm{d}s\bigg{)}\bigg{)}
+2𝔼(sup0tT0t|1N2i,j=1NVxε(Xsi,ε)kε(Xsi,εXsj,ε)Vxε((kερsε)ρsε),\displaystyle+2\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int\limits_{0}^{t}\bigg{|}\bigg{\langle}\frac{1}{N^{2}}\sum\limits_{i,j=1}^{N}V^{\varepsilon}_{x}(\cdot-X_{s}^{i,\varepsilon})k^{\varepsilon}(X_{s}^{i,\varepsilon}-X_{s}^{j,\varepsilon})-V^{\varepsilon}_{x}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon}),
σNl=1N0sVxε(Xul)dBulL2()|ds)\displaystyle\quad\frac{\sigma}{N}\sum\limits_{l=1}^{N}-\int\limits_{0}^{s}V^{\varepsilon}_{x}(\cdot-X_{u}^{l})\,\mathrm{d}B_{u}^{l}\bigg{\rangle}_{L^{2}(\mathbb{R})}\bigg{|}\,\mathrm{d}s\bigg{)}
+4σ2𝔼(sup0tT0tσNi=1N0sVxxε(Xui)dBuL2()2ds)+2Tσ2CBDGNVxεL2()2.\displaystyle+4\sigma^{2}\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int\limits_{0}^{t}\left\lVert\frac{\sigma}{N}\sum\limits_{i=1}^{N}\int\limits_{0}^{s}V^{\varepsilon}_{xx}(\cdot-X_{u}^{i})\,\mathrm{d}B_{u}\right\rVert_{L^{2}(\mathbb{R})}^{2}\,\mathrm{d}s\bigg{)}+\frac{2T\sigma^{2}C_{\mathrm{BDG}}}{N}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}.

Now we want to estimate each term on its own. We will split the fourth terms into fourth separate lemmata to keep a readable structure. The theorem follows immediately by combining Lemma 3.8 and the inequalities (3.5), (3.9), (3.20) in the lemmata below. We will summarize the estimate after we prove the following lemmata.

Lemma 3.7 (Initial Value Inequality).

Let the assumptions of Theorem 3.3 hold true. Then

(3.5) 𝔼(Vεμ0NVερ0L2()2)2NVεL2()2.\mathbb{E}\bigg{(}\left\lVert V^{\varepsilon}*\mu^{N}_{0}-V^{\varepsilon}*\rho_{0}\right\rVert_{L^{2}(\mathbb{R})}^{2}\bigg{)}\leq\frac{2}{N}\left\lVert V^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}.
Proof.

We compute

𝔼(Vεμ0NVερ0L2()2)\displaystyle\;\mathbb{E}\bigg{(}\left\lVert V^{\varepsilon}*\mu^{N}_{0}-V^{\varepsilon}*\rho_{0}\right\rVert_{L^{2}(\mathbb{R})}^{2}\bigg{)}
=\displaystyle= 𝔼((Vεμ0N(y))22Vεμ0N(y)Vερ0(y)+(Vερ0(y))2)dy\displaystyle\;\int_{\mathbb{R}}\mathbb{E}\bigg{(}(V^{\varepsilon}*\mu^{N}_{0}(y))^{2}-2V^{\varepsilon}*\mu^{N}_{0}(y)V^{\varepsilon}*\rho_{0}(y)+(V^{\varepsilon}*\rho_{0}(y))^{2}\bigg{)}\,\mathrm{d}y
=\displaystyle= 1N2i,j=1N𝔼(Vε(yX0i)Vε(yX0j))2Ni=1N𝔼(Vε(yX0i))Vερ0(y)\displaystyle\;\int_{\mathbb{R}}\frac{1}{N^{2}}\sum\limits_{i,j=1}^{N}\mathbb{E}\bigg{(}V^{\varepsilon}(y-X_{0}^{i})V^{\varepsilon}(y-X_{0}^{j})\bigg{)}-\frac{2}{N}\sum\limits_{i=1}^{N}\mathbb{E}\bigg{(}V^{\varepsilon}(y-X_{0}^{i})\bigg{)}V^{\varepsilon}*\rho_{0}(y)
+(Vερ0(y))2dy\displaystyle\quad+(V^{\varepsilon}*\rho_{0}(y))^{2}\,\mathrm{d}y
=\displaystyle= N2NN2(Vερ0(y))2+1N(Vε)2ρ0(y)(Vερ0(y))2dy\displaystyle\int_{\mathbb{R}}\frac{N^{2}-N}{N^{2}}(V^{\varepsilon}*\rho_{0}(y))^{2}+\frac{1}{N}(V^{\varepsilon})^{2}*\rho_{0}(y)-(V^{\varepsilon}*\rho_{0}(y))^{2}\,\mathrm{d}y
=\displaystyle= 1N(Vε)2ρ0(y)(Vερ0(y))2dy\displaystyle\;\frac{1}{N}\int_{\mathbb{R}}(V^{\varepsilon})^{2}*\rho_{0}(y)-(V^{\varepsilon}*\rho_{0}(y))^{2}\,\mathrm{d}y
\displaystyle\leq 1N((Vε)2ρ0L1()+Vερ0L2()2)2NVεL2()2ρ0L1(),\displaystyle\;\frac{1}{N}\big{(}\left\lVert(V^{\varepsilon})^{2}*\rho_{0}\right\rVert_{L^{1}(\mathbb{R})}+\left\lVert V^{\varepsilon}*\rho_{0}\right\rVert_{L^{2}(\mathbb{R})}^{2}\big{)}\leq\;\frac{2}{N}\left\lVert V^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}\left\lVert\rho_{0}\right\rVert_{L^{1}(\mathbb{R})},

where we used the fact that the initial particles are i.i.d. and Young’s inequality for convolutions in the last step. ∎

Lemma 3.8 (Absorbation Inequality).

Let the assumptions of Theorem 3.3 hold true. Then

𝔼(sup0tT0t7σ216VxεμsN,εVxερsεL2()2+|1N2i,j=1NVxε(Xsi,ε)kε(Xsi,εXsj,ε)Vxε((kερsε)ρsε),Vε(μsN,ερsε)L2()|ds)16TkxεL()2VεL2()2σ2N2α+4TVεL2()2σ2N2(α+δ)+(4VxεL2()2N2ασ2+16VεL2()2Nσ2)0TkερsεL()2ds+C(γ)TNγ(VεL2()2kεL()2+VxεL2()2)σ28𝔼(sup0tT0tVxεμsN,εVxερsεL2()2ds).\displaystyle\begin{split}&\;\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int\limits_{0}^{t}-\frac{7\sigma^{2}}{16}\left\lVert V^{\varepsilon}_{x}*\mu_{s}^{N,\varepsilon}-V^{\varepsilon}_{x}*\rho_{s}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}\\ &\quad+\bigg{|}\bigg{\langle}\frac{1}{N^{2}}\sum\limits_{i,j=1}^{N}V^{\varepsilon}_{x}(\cdot-X_{s}^{i,\varepsilon})k^{\varepsilon}(X_{s}^{i,\varepsilon}-X_{s}^{j,\varepsilon})-V^{\varepsilon}_{x}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon}),V^{\varepsilon}*(\mu_{s}^{N,\varepsilon}-\rho_{s}^{\varepsilon})\bigg{\rangle}_{L^{2}(\mathbb{R})}\bigg{|}\,\mathrm{d}s\bigg{)}\\ &\leq\;\frac{16T\left\lVert k_{x}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}^{2}\left\lVert V^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}}{\sigma^{2}N^{2\alpha}}+\frac{4T\left\lVert V^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}}{\sigma^{2}N^{2(\alpha+\delta)}}\\ &\quad+\Big{(}\frac{4\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}}{N^{2\alpha}\sigma^{2}}+\frac{16\left\lVert V^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}}{N\sigma^{2}}\Big{)}\int\limits_{0}^{T}\left\lVert k^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}^{2}\,\mathrm{d}s\\ &\quad+\frac{C(\gamma)T}{N^{\gamma}}\Big{(}\left\lVert V^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}\left\lVert k^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}^{2}+\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}\Big{)}\\ &\quad-\frac{\sigma^{2}}{8}\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int\limits_{0}^{t}\left\lVert V_{x}^{\varepsilon}*\mu_{s}^{N,\varepsilon}-V_{x}^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}\,\mathrm{d}s\bigg{)}.\end{split}
Proof.

Before we begin the proof of this lemma, we will provide an overview of our approach. Our main strategy is to utilize the convergence in probability of the particle Xti,εX_{t}^{i,\varepsilon} to their mean-field limit Yti,εY_{t}^{i,\varepsilon} (Assumption 2.5) in combination with the law of large numbers (Assumption 2.6). This implies that the ”bad set”, where the particles are apart is small in probability with arbitrary algebraic convergence rate. Therefore, we may assume that Xti,εX_{t}^{i,\varepsilon} is close to Yti,εY_{t}^{i,\varepsilon}, and we formally replace the empirical measure of (Xti,ε,i=1,,N)(X_{t}^{i,\varepsilon},i=1,\ldots,N) with the empirical measure associated with (Yti,ε,i=1,,N)(Y_{t}^{i,\varepsilon},i=1,\ldots,N). However, (Yti,ε,i=1,,N)(Y_{t}^{i,\varepsilon},i=1,\ldots,N) has more desirable properties. For instance, the particles are independent and have density ρtεL1()\rho_{t}^{\varepsilon}\in L^{1}(\mathbb{R}) and often even ρtεL()\rho_{t}^{\varepsilon}\in L^{\infty}(\mathbb{R}). This allows us to apply the law of large numbers (2.16), which ultimately proves the claim.

Let us start by splitting our probability space Ω\Omega into two sets. On one set BsαB_{s}^{\alpha} the particles are close to the mean-field particles in probability and “satisfy” the law of large numbers. The other set we take as the complement (Bsα)c(B_{s}^{\alpha})^{\mathrm{c}}, which has small probability by inequalities (2.15) and (2.16).

More precisely, we have

Bsα\displaystyle B_{s}^{\alpha} ={ωΩ:maxi=1,,N|Xsi,ε(ω)Ysi,ε(ω)|Nα}\displaystyle=\bigg{\{}\omega\in\Omega\colon\max\limits_{i=1,\ldots,N}|X^{i,\varepsilon}_{s}(\omega)-Y^{i,\varepsilon}_{s}(\omega)|\leq N^{-\alpha}\bigg{\}}
(3.6) {ωΩ:maxi=1,,N|1Nj=1Nkε(Ysi,ε(ω)Ysj,ε(ω))(kερsε)(Ysi,ε(ω))|N(α+δ)}\displaystyle\quad\cap\bigg{\{}\omega\in\Omega\colon\max\limits_{i=1,\ldots,N}\bigg{|}\frac{1}{N}\sum\limits_{j=1}^{N}k^{\varepsilon}(Y_{s}^{i,\varepsilon}(\omega)-Y_{s}^{j,\varepsilon}(\omega))-(k^{\varepsilon}*\rho_{s}^{\varepsilon})(Y_{s}^{i,\varepsilon}(\omega))\bigg{|}\leq N^{-(\alpha+\delta)}\bigg{\}}

for some δ>0\delta>0 such that 0<α+δ<1/20<\alpha+\delta<1/2 and we have the estimate ((Bsα)c)C(γ)Nγ\mathbb{P}((B_{s}^{\alpha})^{\mathrm{c}})\leq C(\gamma)N^{-\gamma} for all γ>0\gamma>0 by  (2.15) and (2.16). Let us rewrite the last Lebesgue integral on the left-hand side of our claim as follows

𝔼(sup0tT0t|1N2i,j=1NVxε(Xsi,ε)kε(Xsi,εXsj,ε)Vxε((kερsε)ρsε),\displaystyle\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int\limits_{0}^{t}\bigg{|}\bigg{\langle}\frac{1}{N^{2}}\sum\limits_{i,j=1}^{N}V^{\varepsilon}_{x}(\cdot-X_{s}^{i,\varepsilon})k^{\varepsilon}(X_{s}^{i,\varepsilon}-X_{s}^{j,\varepsilon})-V^{\varepsilon}_{x}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon}),
VεμsN,εVερsεL2()|ds)\displaystyle\qquad V^{\varepsilon}*\mu_{s}^{N,\varepsilon}-V^{\varepsilon}*\rho_{s}^{\varepsilon}\bigg{\rangle}_{L^{2}(\mathbb{R})}\bigg{|}\,\mathrm{d}s\bigg{)}
𝔼(sup0tT0t|1N2i,j=1NVxε(Xsi,ε)kε(Xsi,εXsj,ε)Vxε((kερsε)ρsε),\displaystyle\quad\leq\;\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int\limits_{0}^{t}\bigg{|}\bigg{\langle}\frac{1}{N^{2}}\sum\limits_{i,j=1}^{N}V^{\varepsilon}_{x}(\cdot-X_{s}^{i,\varepsilon})k^{\varepsilon}(X_{s}^{i,\varepsilon}-X_{s}^{j,\varepsilon})-V^{\varepsilon}_{x}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon}),
VεμsN,εVερsεL2()|(𝟙(Bsα)+𝟙(Bsα)c)ds).\displaystyle\qquad V^{\varepsilon}*\mu_{s}^{N,\varepsilon}-V^{\varepsilon}*\rho_{s}^{\varepsilon}\bigg{\rangle}_{L^{2}(\mathbb{R})}\bigg{|}\Big{(}\mathbbm{1}_{(B_{s}^{\alpha})}+\mathbbm{1}_{(B_{s}^{\alpha})^{\mathrm{c}}}\Big{)}\,\mathrm{d}s\bigg{)}.

We are going to estimate each term by itself.

On the set BsαB_{s}^{\alpha}: In order to estimate the first term above we let ωBsα\omega\in B_{s}^{\alpha} and will not write the indicator function. Then we have

1N2i,j=1NVxε(Xsi,ε(ω))kε(Xsi,ε(ω)Xsj,ε(ω))Vxε((kερsε)ρsε)),Vε(μsN,ε(ω)ρsε)L2\displaystyle\frac{1}{N^{2}}\sum\limits_{i,j=1}^{N}\Big{\langle}V^{\varepsilon}_{x}(\cdot-X_{s}^{i,\varepsilon}(\omega))k^{\varepsilon}(X_{s}^{i,\varepsilon}(\omega)-X_{s}^{j,\varepsilon}(\omega))-V^{\varepsilon}_{x}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon})),V^{\varepsilon}*(\mu_{s}^{N,\varepsilon}(\omega)-\rho_{s}^{\varepsilon})\Big{\rangle}_{L^{2}}
=\displaystyle= 1N2i,j=1NVxε(Xsi,ε(ω))(kε(Xsi,ε(ω)Xsj,ε(ω))kε(Ysi,ε(ω)Ysj,ε(ω))),\displaystyle\;\frac{1}{N^{2}}\sum\limits_{i,j=1}^{N}\Big{\langle}V^{\varepsilon}_{x}(\cdot-X_{s}^{i,\varepsilon}(\omega))(k^{\varepsilon}(X_{s}^{i,\varepsilon}(\omega)-X_{s}^{j,\varepsilon}(\omega))-k^{\varepsilon}(Y_{s}^{i,\varepsilon}(\omega)-Y_{s}^{j,\varepsilon}(\omega))),
Vε(μsN,ε(ω)ρsε)L2()\displaystyle\quad\quad V^{\varepsilon}*(\mu_{s}^{N,\varepsilon}(\omega)-\rho_{s}^{\varepsilon})\Big{\rangle}_{L^{2}(\mathbb{R})}
+1N2i,j=1NVxε(Xsi,ε(ω))kε(Ysi,ε(ω)Ysj,ε(ω))Vxε((kερsε)ρsε)),\displaystyle+\frac{1}{N^{2}}\sum\limits_{i,j=1}^{N}\Big{\langle}V^{\varepsilon}_{x}(\cdot-X_{s}^{i,\varepsilon}(\omega))k^{\varepsilon}(Y_{s}^{i,\varepsilon}(\omega)-Y_{s}^{j,\varepsilon}(\omega))-V^{\varepsilon}_{x}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon})),
Vε(μsN,ε(ω)ρsε)L2()\displaystyle\quad\quad V^{\varepsilon}*(\mu_{s}^{N,\varepsilon}(\omega)-\rho_{s}^{\varepsilon})\Big{\rangle}_{L^{2}(\mathbb{R})}
=\displaystyle= Is1(ω)+Is2(ω).\displaystyle\;I_{s}^{1}(\omega)+I_{s}^{2}(\omega).

For the first term we obtain

|Is1(ω)|=\displaystyle|I_{s}^{1}(\omega)|= |1Nj=1N1Ni=1NVε(Xsi,ε(ω))(kε(Xsi,ε(ω)Xsj,ε(ω))kε(Ysi,ε(ω)Ysj,ε(ω))),\displaystyle\;\bigg{|}\frac{1}{N}\sum\limits_{j=1}^{N}\bigg{\langle}\frac{1}{N}\sum\limits_{i=1}^{N}V^{\varepsilon}(\cdot-X_{s}^{i,\varepsilon}(\omega))(k^{\varepsilon}(X_{s}^{i,\varepsilon}(\omega)-X_{s}^{j,\varepsilon}(\omega))-k^{\varepsilon}(Y_{s}^{i,\varepsilon}(\omega)-Y_{s}^{j,\varepsilon}(\omega))),
Vxε(μsN,ε(ω)ρsε)L2()|\displaystyle\qquad V^{\varepsilon}_{x}*(\mu_{s}^{N,\varepsilon}(\omega)-\rho_{s}^{\varepsilon})\bigg{\rangle}_{L^{2}(\mathbb{R})}\bigg{|}
\displaystyle\leq 1N2i,j=1N|(Vε(Xsi,ε(ω))|max1iN|kε(Xsi,ε(ω)Xsj,ε(ω))kε(Ysi,ε(ω)Ysj,ε(ω))|,\displaystyle\;\frac{1}{N^{2}}\sum\limits_{i,j=1}^{N}\bigg{\langle}|(V^{\varepsilon}(\cdot-X_{s}^{i,\varepsilon}(\omega))|\max\limits_{1\leq i\leq N}|k^{\varepsilon}(X_{s}^{i,\varepsilon}(\omega)-X_{s}^{j,\varepsilon}(\omega))-k^{\varepsilon}(Y_{s}^{i,\varepsilon}(\omega)-Y_{s}^{j,\varepsilon}(\omega))|,
|Vxε(μsN,ε(ω)ρsε|L2()\displaystyle\;|V^{\varepsilon}_{x}*(\mu_{s}^{N,\varepsilon}(\omega)-\rho_{s}^{\varepsilon}|\bigg{\rangle}_{L^{2}(\mathbb{R})}
\displaystyle\leq 2Ni=1NkxεL()|Vε(Xsi,ε(ω))|max1iN|Xsi,NYsi,N|,|Vxε(μsN,ε(ω)ρsε)|L2\displaystyle\frac{2}{N}\sum\limits_{i=1}^{N}\bigg{\langle}\left\lVert k_{x}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}|V^{\varepsilon}(\cdot-X_{s}^{i,\varepsilon}(\omega))|\max\limits_{1\leq i\leq N}|X^{i,N}_{s}-Y_{s}^{i,N}|,|V^{\varepsilon}_{x}*(\mu_{s}^{N,\varepsilon}(\omega)-\rho_{s}^{\varepsilon})|\bigg{\rangle}_{L^{2}}
\displaystyle\leq 2Ni=1NNαkxεL()|Vε(Xsi,ε(ω))|,|Vxε(μsN,ε(ω)ρsε|L2()\displaystyle\;\frac{2}{N}\sum\limits_{i=1}^{N}\langle N^{-\alpha}\left\lVert k_{x}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}|V^{\varepsilon}(\cdot-X_{s}^{i,\varepsilon}(\omega))|,|V^{\varepsilon}_{x}*(\mu_{s}^{N,\varepsilon}(\omega)-\rho_{s}^{\varepsilon}|\rangle_{L^{2}(\mathbb{R})}
\displaystyle\leq 2Ni=1N8kxεL()2σ2N2α|Vε(yXsi,ε(ω))|2dy+σ232|Vxε(μsN,ε(ω)ρsε)(y)|2dy\displaystyle\;\frac{2}{N}\sum\limits_{i=1}^{N}\int_{\mathbb{R}}\frac{8\left\lVert k_{x}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}^{2}}{\sigma^{2}N^{2\alpha}}|V^{\varepsilon}(y-X_{s}^{i,\varepsilon}(\omega))|^{2}\,\mathrm{d}y+\frac{\sigma^{2}}{32}\int_{\mathbb{R}}|V^{\varepsilon}_{x}*(\mu_{s}^{N,\varepsilon}(\omega)-\rho_{s}^{\varepsilon})(y)|^{2}\,\mathrm{d}y
(3.7) \displaystyle\leq 16σ2N2αkxεL()2VεL2()2+σ216Vxε(μsN,ε(ω)ρsε)L2()2.\displaystyle\;\frac{16}{\sigma^{2}N^{2\alpha}}\left\lVert k_{x}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}^{2}\left\lVert V^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}+\frac{\sigma^{2}}{16}\left\lVert V^{\varepsilon}_{x}*(\mu_{s}^{N,\varepsilon}(\omega)-\rho_{s}^{\varepsilon})\right\rVert_{L^{2}(\mathbb{R})}^{2}.

Here we used integration by parts in the first step, the property of the set BsαB_{s}^{\alpha} in the fourth step. As always, we neglect the last term by absorbing it into the diffusion in our statement.

We treat the term Is2(ω)I_{s}^{2}(\omega) using the law of large numbers property of the second term in BsαB_{s}^{\alpha}. For ωBsα\omega\in B_{s}^{\alpha} we rewrite

|Is2(ω)|\displaystyle|I_{s}^{2}(\omega)| =|1N2i,j=1NVxε(Xsi,ε(ω))kε(Ysi,ε(ω)Ysj,ε(ω))Vxε((kερsε)ρsε)),\displaystyle=\Big{|}\frac{1}{N^{2}}\sum\limits_{i,j=1}^{N}\Big{\langle}V^{\varepsilon}_{x}(\cdot-X_{s}^{i,\varepsilon}(\omega))k^{\varepsilon}(Y_{s}^{i,\varepsilon}(\omega)-Y_{s}^{j,\varepsilon}(\omega))-V^{\varepsilon}_{x}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon})),
Vε(μsN,ε(ω)ρsε)L2()|\displaystyle\qquad V^{\varepsilon}*(\mu_{s}^{N,\varepsilon}(\omega)-\rho_{s}^{\varepsilon})\Big{\rangle}_{L^{2}(\mathbb{R})}\Big{|}
=|1N2i,j=1NVε(Xsi,ε(ω))kε(Ysi,ε(ω)Ysj,ε(ω))Vε((kερsε)ρsε)),\displaystyle=\Big{|}\frac{1}{N^{2}}\sum\limits_{i,j=1}^{N}\Big{\langle}V^{\varepsilon}(\cdot-X_{s}^{i,\varepsilon}(\omega))k^{\varepsilon}(Y_{s}^{i,\varepsilon}(\omega)-Y_{s}^{j,\varepsilon}(\omega))-V^{\varepsilon}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon})),
Vxε(μsN,ε(ω)ρsε)L2()|\displaystyle\qquad V^{\varepsilon}_{x}*(\mu_{s}^{N,\varepsilon}(\omega)-\rho_{s}^{\varepsilon})\Big{\rangle}_{L^{2}(\mathbb{R})}\Big{|}
=|1N2i,j=1NVε(Xsi,ε(ω))(kε(Ysi,ε(ω)Ysj,ε(ω))(kερsε)(Ysi,ε(ω)))\displaystyle=\Big{|}\frac{1}{N^{2}}\sum\limits_{i,j=1}^{N}\Big{\langle}V^{\varepsilon}(\cdot-X_{s}^{i,\varepsilon}(\omega))(k^{\varepsilon}(Y_{s}^{i,\varepsilon}(\omega)-Y_{s}^{j,\varepsilon}(\omega))-(k^{\varepsilon}*\rho_{s}^{\varepsilon})(Y_{s}^{i,\varepsilon}(\omega)))
+(Vε(Xsi,ε(ω))Vε(Ysi,ε(ω)))(kερsε)(Ysi,ε(ω))\displaystyle+(V^{\varepsilon}(\cdot-X_{s}^{i,\varepsilon}(\omega))-V^{\varepsilon}(\cdot-Y_{s}^{i,\varepsilon}(\omega)))(k^{\varepsilon}*\rho_{s}^{\varepsilon})(Y_{s}^{i,\varepsilon}(\omega))
+Vε(Ysi,ε(ω))(kερsε)(Ysi,ε(ω))Vε((kερsε)ρsε)),Vxε(μsN,ε(ω)ρsε)L2()|\displaystyle+V^{\varepsilon}(\cdot-Y_{s}^{i,\varepsilon}(\omega))(k^{\varepsilon}*\rho_{s}^{\varepsilon})(Y_{s}^{i,\varepsilon}(\omega))-V^{\varepsilon}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon})),V^{\varepsilon}_{x}*(\mu_{s}^{N,\varepsilon}(\omega)-\rho_{s}^{\varepsilon})\Big{\rangle}_{L^{2}(\mathbb{R})}\Big{|}
(3.8) =|Is21(ω)|+|Is22(ω)|+|Is23(ω)|.\displaystyle=|I_{s}^{21}(\omega)|+|I_{s}^{22}(\omega)|+|I_{s}^{23}(\omega)|.

For the first term Is21(ω)I_{s}^{21}(\omega) we obtain

|Is21(ω)|\displaystyle|I_{s}^{21}(\omega)| 1Ni=1N|Vε(Xsi,ε(ω))||1Nj=1Nkε(Ysi,ε(ω)Ysj,ε(ω))(kερsε)(Ysi,ε(ω))|,\displaystyle\leq\frac{1}{N}\sum\limits_{i=1}^{N}\bigg{\langle}|V^{\varepsilon}(\cdot-X_{s}^{i,\varepsilon}(\omega))|\bigg{|}\frac{1}{N}\sum\limits_{j=1}^{N}k^{\varepsilon}(Y_{s}^{i,\varepsilon}(\omega)-Y_{s}^{j,\varepsilon}(\omega))-(k^{\varepsilon}*\rho_{s}^{\varepsilon})(Y_{s}^{i,\varepsilon}(\omega))\bigg{|},
|Vxε(μsN,ε(ω)ρsε)|L2()\displaystyle\quad|V^{\varepsilon}_{x}*(\mu_{s}^{N,\varepsilon}(\omega)-\rho_{s}^{\varepsilon})|\bigg{\rangle}_{L^{2}(\mathbb{R})}
1Ni=1NN(α+δ)|Vε(Xsi,ε(ω))|,|Vxε(μsN,ε(ω)ρsε)|L2()\displaystyle\leq\frac{1}{N}\sum\limits_{i=1}^{N}\langle N^{-(\alpha+\delta)}|V^{\varepsilon}(\cdot-X_{s}^{i,\varepsilon}(\omega))|,|V^{\varepsilon}_{x}*(\mu_{s}^{N,\varepsilon}(\omega)-\rho_{s}^{\varepsilon})|\rangle_{L^{2}(\mathbb{R})}
(3.9) 4N2(α+δ)σ2VεL2()2+σ216Vxε(μsN,ε(ω)ρsε)L2()2,\displaystyle\leq\frac{4N^{-2(\alpha+\delta)}}{\sigma^{2}}\left\lVert V^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}+\frac{\sigma^{2}}{16}\left\lVert V^{\varepsilon}_{x}*(\mu_{s}^{N,\varepsilon}(\omega)-\rho_{s}^{\varepsilon})\right\rVert_{L^{2}(\mathbb{R})}^{2},

where we used the property of the set BsαB_{s}^{\alpha} in the second step and Young’s inequality.

Using the fact that we are still on the set BsαB_{s}^{\alpha} we obtain for the second term Is22(ω)I_{s}^{22}(\omega) the following estimate

|Is22(ω)|\displaystyle|I_{s}^{22}(\omega)|\leq 4kερsεL()2Nσ2i=1N|Vε(yXsi,ε(ω))Vε(yYsi,ε(ω)))|2dy\displaystyle\;\frac{4\left\lVert k^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}^{2}}{N\sigma^{2}}\sum\limits_{i=1}^{N}\int_{\mathbb{R}}|V^{\varepsilon}(y-X_{s}^{i,\varepsilon}(\omega))-V^{\varepsilon}(y-Y_{s}^{i,\varepsilon}(\omega)))|^{2}\,\mathrm{d}y
+σ216|Vxε(μsN,ε(ω)ρsε)(y)|2dy\displaystyle\qquad+\frac{\sigma^{2}}{16}\int_{\mathbb{R}}|V^{\varepsilon}_{x}*(\mu_{s}^{N,\varepsilon}(\omega)-\rho_{s}^{\varepsilon})(y)|^{2}\,\mathrm{d}y
=\displaystyle= 4kερsεL()2Nσ2i=1N|01ddrVε(yYsi,ε(ω)+r(Ysi,ε(ω)Xsi,ε(ω)))dr|2dy\displaystyle\;\frac{4\left\lVert k^{\varepsilon}*\rho^{\varepsilon}_{s}\right\rVert_{L^{\infty}(\mathbb{R})}^{2}}{N\sigma^{2}}\sum\limits_{i=1}^{N}\int_{\mathbb{R}}\bigg{|}\int\limits_{0}^{1}\frac{\,\mathrm{d}}{\,\mathrm{d}r}V^{\varepsilon}(y-Y_{s}^{i,\varepsilon}(\omega)+r(Y_{s}^{i,\varepsilon}(\omega)-X_{s}^{i,\varepsilon}(\omega)))\,\mathrm{d}r\bigg{|}^{2}\,\mathrm{d}y
+σ216|Vxε(μsN,ε(ω)ρsε)(y)|2dy\displaystyle\qquad+\frac{\sigma^{2}}{16}\int_{\mathbb{R}}|V^{\varepsilon}_{x}*(\mu_{s}^{N,\varepsilon}(\omega)-\rho_{s}^{\varepsilon})(y)|^{2}\,\mathrm{d}y
4kερsεL()2σ2max1iN|Ysi,ε(ω)Xsi,ε(ω)|2\displaystyle\leq\;\frac{4\left\lVert k^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}^{2}}{\sigma^{2}}\max_{1\leq i\leq N}|Y_{s}^{i,\varepsilon}(\omega)-X_{s}^{i,\varepsilon}(\omega)|^{2}
1Ni=1N01|ddxVε(yYsi,ε(ω)+r(Ysi,ε(ω)Xsi,ε(ω)))|2dydr\displaystyle\qquad\qquad\cdot\frac{1}{N}\sum\limits_{i=1}^{N}\int\limits_{0}^{1}\int_{\mathbb{R}}\Big{|}\frac{\,\mathrm{d}}{\,\mathrm{d}x}V^{\varepsilon}(y-Y_{s}^{i,\varepsilon}(\omega)+r(Y_{s}^{i,\varepsilon}(\omega)-X_{s}^{i,\varepsilon}(\omega)))\Big{|}^{2}\,\mathrm{d}y\,\mathrm{d}r
+σ216|Vxε(μsN,ε(ω)ρsε)(y)|2dy\displaystyle\qquad+\frac{\sigma^{2}}{16}\int_{\mathbb{R}}|V^{\varepsilon}_{x}*(\mu_{s}^{N,\varepsilon}(\omega)-\rho_{s}^{\varepsilon})(y)|^{2}\,\mathrm{d}y
=4kερsεL()2σ2max1iN|Ysi,ε(ω)Xsi,ε(ω)|201|Vzε(z)|2dzdr\displaystyle=\;\frac{4\left\lVert k^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}^{2}}{\sigma^{2}}\max_{1\leq i\leq N}|Y_{s}^{i,\varepsilon}(\omega)-X_{s}^{i,\varepsilon}(\omega)|^{2}\int\limits_{0}^{1}\int_{\mathbb{R}}|V^{\varepsilon}_{z}(z)|^{2}\,\mathrm{d}z\,\mathrm{d}r
+σ216|Vxε(μsN,ε(ω)ρsε)(y)|2dy\displaystyle\quad+\frac{\sigma^{2}}{16}\int_{\mathbb{R}}|V^{\varepsilon}_{x}*(\mu_{s}^{N,\varepsilon}(\omega)-\rho_{s}^{\varepsilon})(y)|^{2}\,\mathrm{d}y
(3.10) 4kερsεL()2N2ασ2VxεL2()2+σ216Vxε(μsN,ε(ω)ρsε)L2()2.\displaystyle\leq\;\frac{4\left\lVert k^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}^{2}}{N^{2\alpha}\sigma^{2}}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}+\frac{\sigma^{2}}{16}\left\lVert V^{\varepsilon}_{x}*(\mu_{s}^{N,\varepsilon}(\omega)-\rho_{s}^{\varepsilon})\right\rVert_{L^{2}(\mathbb{R})}^{2}.

In the above calculations we used Young’s inequality in the first step, Jensen inequality in the second estimate, the property of the set BsαB_{s}^{\alpha} in the third estimate.

In order to estimate the last term Is23(ω)I_{s}^{23}(\omega) in (3.2) we use the independence of our mean-field particles (Yti,ε,i=1,,N)(Y_{t}^{i,\varepsilon},i=1,\ldots,N). Hence, we can no longer do the estimates pathwise and need to take advantage of the expectation. First, applying Young’s inequality we find

|Is23(ω)|\displaystyle|I_{s}^{23}(\omega)| 4σ21N2|i=1NVε(yYsi,ε(ω))(kερsε)(Ysi,ε(ω))Vε((kερsε)ρsε))(y)|2dy\displaystyle\leq\;\frac{4}{\sigma^{2}}\int_{\mathbb{R}}\frac{1}{N^{2}}\bigg{|}\sum\limits_{i=1}^{N}V^{\varepsilon}(y-Y_{s}^{i,\varepsilon}(\omega))(k^{\varepsilon}*\rho_{s}^{\varepsilon})(Y_{s}^{i,\varepsilon}(\omega))-V^{\varepsilon}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon}))(y)\bigg{|}^{2}\,\mathrm{d}y
+σ216Vxε(μsN,ε(ω)ρsε)L2()2.\displaystyle\quad\quad+\frac{\sigma^{2}}{16}\left\lVert V^{\varepsilon}_{x}*(\mu_{s}^{N,\varepsilon}(\omega)-\rho_{s}^{\varepsilon})\right\rVert_{L^{2}(\mathbb{R})}^{2}.

As always, the last term is going to be absorbed. For the first term, we recall that our statement has an supremum over all 0tT0\leq t\leq T and an expectation. Hence, it is enough to estimate

𝔼(sup0tT0t4σ21N2|i=1NVε(yYsi,ε(ω))(kερsε)(Ysi,ε(ω))Vε((kερsε)ρsε))(y)|2dy)\displaystyle\;\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int\limits_{0}^{t}\frac{4}{\sigma^{2}}\int_{\mathbb{R}}\frac{1}{N^{2}}\bigg{|}\sum\limits_{i=1}^{N}V^{\varepsilon}(y-Y_{s}^{i,\varepsilon}(\omega))(k^{\varepsilon}*\rho_{s}^{\varepsilon})(Y_{s}^{i,\varepsilon}(\omega))-V^{\varepsilon}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon}))(y)\bigg{|}^{2}\,\mathrm{d}y\bigg{)}
=\displaystyle= 0T4N2σ2𝔼(|i=1NVε(yYsi,ε(ω))(kερsε)(Ysi,ε(ω))Vε((kερsε)ρsε))(y)|2)dy.\displaystyle\;\int\limits_{0}^{T}\frac{4}{N^{2}\sigma^{2}}\int_{\mathbb{R}}\mathbb{E}\bigg{(}\bigg{|}\sum\limits_{i=1}^{N}V^{\varepsilon}(y-Y_{s}^{i,\varepsilon}(\omega))(k^{\varepsilon}*\rho_{s}^{\varepsilon})(Y_{s}^{i,\varepsilon}(\omega))-V^{\varepsilon}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon}))(y)\bigg{|}^{2}\bigg{)}\,\mathrm{d}y.

Let us denote for fix yy\in\mathbb{R}

Zsi(ω):=Vε(yYsi,ε(ω))(kερsε)(Ysi,ε(ω))Vε((kερsε)ρsε))(y).Z^{i}_{s}(\omega):=V^{\varepsilon}(y-Y_{s}^{i,\varepsilon}(\omega))(k^{\varepsilon}*\rho_{s}^{\varepsilon})(Y_{s}^{i,\varepsilon}(\omega))-V^{\varepsilon}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon}))(y).

Then we notice that

𝔼(Zsi)\displaystyle\mathbb{E}(Z^{i}_{s}) =𝔼(Vε(yYsi,ε)(kερsε)(Ysi,ε))Vε((kερsε)ρsε))(y)\displaystyle=\mathbb{E}(V^{\varepsilon}(y-Y_{s}^{i,\varepsilon})(k^{\varepsilon}*\rho_{s}^{\varepsilon})(Y_{s}^{i,\varepsilon}))-V^{\varepsilon}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon}))(y)
=Vε(yz)(kερsε)(z)ρsε(z)dzVε((kερsε)ρsε))(y)=0.\displaystyle=\int_{\mathbb{R}}V^{\varepsilon}(y-z)(k^{\varepsilon}*\rho_{s}^{\varepsilon})(z)\rho_{s}^{\varepsilon}(z)\,\mathrm{d}z-V^{\varepsilon}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon}))(y)=0.

Furthermore, we have the random variables (Zsi,i=1,,N)(Z_{s}^{i},i=1,\ldots,N) are pairwise independent. Hence, if iji\neq j we find

𝔼(ZsiZsj)=𝔼(Zsi)𝔼(Zsj)=0.\mathbb{E}(Z^{i}_{s}Z^{j}_{s})=\mathbb{E}(Z^{i}_{s})\mathbb{E}(Z^{j}_{s})=0.

We notice that we have

𝔼(|i=1NVε(yYsi,ε(ω))(kερsε)(Ysi,ε(ω))Vε((kερsε)ρsε))(y)|2)\displaystyle\;\mathbb{E}\bigg{(}\bigg{|}\sum\limits_{i=1}^{N}V^{\varepsilon}(y-Y_{s}^{i,\varepsilon}(\omega))(k^{\varepsilon}*\rho_{s}^{\varepsilon})(Y_{s}^{i,\varepsilon}(\omega))-V^{\varepsilon}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon}))(y)\bigg{|}^{2}\bigg{)}
=𝔼(|i=1NZsi|2)=i=1N𝔼(|Zsi|2).\displaystyle\quad=\;\mathbb{E}\bigg{(}\bigg{|}\sum\limits_{i=1}^{N}Z_{s}^{i}\bigg{|}^{2}\bigg{)}=\sum\limits_{i=1}^{N}\mathbb{E}(|Z^{i}_{s}|^{2}).

On the other hand by using the trivial inequality (a+b)22(a2+b2)(a+b)^{2}\leq 2(a^{2}+b^{2}) and Young’s inequality for convolution we obtain

𝔼(|Zsi|2)\displaystyle\int_{\mathbb{R}}\mathbb{E}(|Z^{i}_{s}|^{2}) 2𝔼(|Vε(yYsi,ε(ω))(kερsε)(Ysi,ε(ω))|2+|Vε((kερsε)ρsε))(y)|2dy)\displaystyle\leq 2\mathbb{E}\bigg{(}\int_{\mathbb{R}}|V^{\varepsilon}(y-Y_{s}^{i,\varepsilon}(\omega))(k^{\varepsilon}*\rho_{s}^{\varepsilon})(Y_{s}^{i,\varepsilon}(\omega))|^{2}+|V^{\varepsilon}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon}))(y)|^{2}\,\mathrm{d}y\bigg{)}
2kερsεL()2VεL2()2+2VεL2()2(kερsε)ρsεL1()2\displaystyle\leq 2\left\lVert k^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}^{2}\left\lVert V^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}+2\left\lVert V^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}\left\lVert(k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon}\right\rVert_{L^{1}(\mathbb{R})}^{2}
=4kερsεL()2VεL2()2.\displaystyle=4\left\lVert k^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}^{2}\left\lVert V^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}.

Hence, the estimate for I23I^{23} follows by the previous law of large numbers argument and is obtained in the following

(3.11) 𝔼(sup0tT0t|Is23(ω)|ds)\displaystyle\;\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int\limits_{0}^{t}|I^{23}_{s}(\omega)|\,\mathrm{d}s\bigg{)}
σ2160tVxε(μsN,ε(ω)ρsε)L2()2+16VεL2()2Nσ20TkερsεL()2ds.\displaystyle\quad\leq\;\frac{\sigma^{2}}{16}\int\limits_{0}^{t}\left\lVert V^{\varepsilon}_{x}*(\mu_{s}^{N,\varepsilon}(\omega)-\rho_{s}^{\varepsilon})\right\rVert_{L^{2}(\mathbb{R})}^{2}+\frac{16\left\lVert V^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}}{N\sigma^{2}}\int\limits_{0}^{T}\left\lVert k^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}^{2}\,\mathrm{d}s.

By combining the estimates (3.2)(3.2)(3.11) with (3.2) and (3.7) we obtain the estimate on the set BsαB^{\alpha}_{s}

(3.12) 𝔼(sup0tT0t(|Is1(ω)|+|Is2(ω)|)𝟙(Bsα)ds)\displaystyle\;\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int\limits_{0}^{t}(|I^{1}_{s}(\omega)|+|I^{2}_{s}(\omega)|)\mathbbm{1}_{(B_{s}^{\alpha})}\,\mathrm{d}s\bigg{)}
\displaystyle\leq 3σ2160TVxε(μsN,ε(ω)ρsε)L2()2ds+16TkxεL()VεL2()2σ2N2α+4TVεL2()2σ2N2(α+δ)\displaystyle\;-\frac{3\sigma^{2}}{16}\int\limits_{0}^{T}\left\lVert V^{\varepsilon}_{x}*(\mu_{s}^{N,\varepsilon}(\omega)-\rho_{s}^{\varepsilon})\right\rVert_{L^{2}(\mathbb{R})}^{2}\,\mathrm{d}s+\frac{16T\left\lVert k_{x}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}\left\lVert V^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}}{\sigma^{2}N^{2\alpha}}+\frac{4T\left\lVert V^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}}{\sigma^{2}N^{2(\alpha+\delta)}}
+(4VxεL2()2N2ασ2+16VεL2()2Nσ2)0TkερsεL()2ds\displaystyle+\Big{(}\frac{4\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}}{N^{2\alpha}\sigma^{2}}+\frac{16\left\lVert V^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}}{N\sigma^{2}}\Big{)}\int\limits_{0}^{T}\left\lVert k^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}^{2}\,\mathrm{d}s

It remains to obtain an estimate on the complement of BsαB_{s}^{\alpha}.

On the set (Bsα)c(B_{s}^{\alpha})^{\mathrm{c}}: Applying Young’s inequality, multiple Hölder’s inequalities, the fact that ((Bsα)c)C(γ)Nγ\mathbb{P}((B_{s}^{\alpha})^{\mathrm{c}})\leq C(\gamma)N^{-\gamma}, we obtain

𝔼(sup0tT0t𝟙(Bsα)c|1N2i,j=1NVεx(Xsi,ε(ω))kε(Xsi,ε(ω)Xsj,ε(ω))Vεx((kερsε)ρsε)),\displaystyle\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int\limits_{0}^{t}\mathbbm{1}_{(B_{s}^{\alpha})^{\mathrm{c}}}\Big{|}\frac{1}{N^{2}}\sum\limits_{i,j=1}^{N}\Big{\langle}V^{\varepsilon}_{x}(\cdot-X_{s}^{i,\varepsilon}(\omega))k^{\varepsilon}(X_{s}^{i,\varepsilon}(\omega)-X_{s}^{j,\varepsilon}(\omega))-V^{\varepsilon}_{x}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon})),
Vε(μsN,ε(ω)ρsε)L2()|ds)\displaystyle\quad\;V^{\varepsilon}*(\mu_{s}^{N,\varepsilon}(\omega)-\rho_{s}^{\varepsilon})\Big{\rangle}_{L^{2}(\mathbb{R})}\Big{|}\,\mathrm{d}s\bigg{)}
\displaystyle\leq 1N2i,j=1N𝔼(sup0tT0t𝟙(Bsα)c|Vε(Xsi,ε(ω))kε(Xsi,ε(ω)Xsj,ε(ω))Vε((kερsε)ρsε)),\displaystyle\;\frac{1}{N^{2}}\sum\limits_{i,j=1}^{N}\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int\limits_{0}^{t}\mathbbm{1}_{(B_{s}^{\alpha})^{\mathrm{c}}}\Big{|}\Big{\langle}V^{\varepsilon}(\cdot-X_{s}^{i,\varepsilon}(\omega))k^{\varepsilon}(X_{s}^{i,\varepsilon}(\omega)-X_{s}^{j,\varepsilon}(\omega))-V^{\varepsilon}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon})),
Vεx(μsN,ε(ω)ρsε)L2()|ds)\displaystyle\quad V^{\varepsilon}_{x}*(\mu_{s}^{N,\varepsilon}(\omega)-\rho_{s}^{\varepsilon})\Big{\rangle}_{L^{2}(\mathbb{R})}\Big{|}\,\mathrm{d}s\bigg{)}
\displaystyle\leq 1N2i,j=1N𝔼(0T𝟙(Bsα)c(Vε(Xsi,ε(ω))kε(Xsi,ε(ω)Xsj,ε(ω))L2()2\displaystyle\;\frac{1}{N^{2}}\sum\limits_{i,j=1}^{N}\mathbb{E}\bigg{(}\int\limits_{0}^{T}\mathbbm{1}_{(B_{s}^{\alpha})^{\mathrm{c}}}\Big{(}\left\lVert V^{\varepsilon}(\cdot-X_{s}^{i,\varepsilon}(\omega))k^{\varepsilon}(X_{s}^{i,\varepsilon}(\omega)-X_{s}^{j,\varepsilon}(\omega))\right\rVert_{L^{2}(\mathbb{R})}^{2}
+Vε((kερsε)ρsε))L2()2)ds)+12𝔼(0T𝟙(Bsα)cVεx(μsN,ε(ω)ρsε)L2()2ds)\displaystyle\quad+\left\lVert V^{\varepsilon}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon}))\right\rVert_{L^{2}(\mathbb{R})}^{2}\Big{)}\,\mathrm{d}s\bigg{)}+\frac{1}{2}\mathbb{E}\bigg{(}\int\limits_{0}^{T}\mathbbm{1}_{(B_{s}^{\alpha})^{\mathrm{c}}}\left\lVert V^{\varepsilon}_{x}*(\mu_{s}^{N,\varepsilon}(\omega)-\rho_{s}^{\varepsilon})\right\rVert_{L^{2}(\mathbb{R})}^{2}\,\mathrm{d}s\bigg{)}
\displaystyle\leq 1Ni=1N𝔼(0T𝟙(Bsα)c(Vε(Xsi,ε(ω))L2()2kεL2\displaystyle\;\frac{1}{N}\sum\limits_{i=1}^{N}\mathbb{E}\bigg{(}\int\limits_{0}^{T}\mathbbm{1}_{(B_{s}^{\alpha})^{\mathrm{c}}}\Big{(}\left\lVert V^{\varepsilon}(\cdot-X_{s}^{i,\varepsilon}(\omega))\right\rVert_{L^{2}(\mathbb{R})}^{2}\left\lVert k^{\varepsilon}\right\rVert_{L^{\infty}}^{2}
+VεL2()2kερsεL()2ρsεL1()2)ds)+20T𝔼(𝟙(Bsα)cVεxL2()2)ds\displaystyle\quad+\left\lVert V^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}\left\lVert k^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}^{2}\left\lVert\rho_{s}^{\varepsilon}\right\rVert_{L^{1}(\mathbb{R})}^{2}\Big{)}\,\mathrm{d}s\bigg{)}+2\int\limits_{0}^{T}\mathbb{E}\bigg{(}\mathbbm{1}_{(B_{s}^{\alpha})^{\mathrm{c}}}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}\bigg{)}\,\mathrm{d}s
\displaystyle\leq  20T((Bsα)c)(VεL2()2kεL()2+VεxL2()2)ds\displaystyle\;2\int\limits_{0}^{T}\mathbb{P}\Big{(}(B_{s}^{\alpha})^{\mathrm{c}}\Big{)}\Big{(}\left\lVert V^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}\left\lVert k^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}^{2}+\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}\Big{)}\,\mathrm{d}s
\displaystyle\leq C(γ)TNγ(VεL2()2kεL()2+VεxL2()2).\displaystyle\;\frac{C(\gamma)T}{N^{\gamma}}\Big{(}\left\lVert V^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}\left\lVert k^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}^{2}+\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}\Big{)}.

Combined with the estimate on the set BαsB^{\alpha}_{s} , we obtained the result. ∎

Lemma 3.9 (Stochastic Remaining Term Inequality).

Let the assumptions of Theorem 3.3 hold true. Then

𝔼(sup0tT0t|1N2i,j=1NVεx(Xsi,ε)kε(Xsi,εXsj,ε)Vεx((kερsε)ρsε),\displaystyle\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int\limits_{0}^{t}\bigg{|}\bigg{\langle}\frac{1}{N^{2}}\sum\limits_{i,j=1}^{N}V^{\varepsilon}_{x}(\cdot-X_{s}^{i,\varepsilon})k^{\varepsilon}(X_{s}^{i,\varepsilon}-X_{s}^{j,\varepsilon})-V^{\varepsilon}_{x}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon}),
σNl=1N0sVεx(Xul)dBulL2()|ds)\displaystyle\quad\quad\frac{\sigma}{N}\sum\limits_{l=1}^{N}-\int\limits_{0}^{s}V^{\varepsilon}_{x}(\cdot-X_{u}^{l})\,\mathrm{d}B_{u}^{l}\bigg{\rangle}_{L^{2}(\mathbb{R})}\bigg{|}\,\mathrm{d}s\bigg{)}
(3.13) \displaystyle\leq 2σT32CBDG12Nα+12VεxL2()2kεxL()+σCBDG12T32Nα+δ+12VεxL2()2\displaystyle\;\frac{2\sigma T^{\frac{3}{2}}C_{\mathrm{BDG}}^{\frac{1}{2}}}{N^{\alpha+\frac{1}{2}}}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}\left\lVert k^{\varepsilon}_{x}\right\rVert_{L^{\infty}(\mathbb{R})}+\sigma\frac{C_{\mathrm{BDG}}^{\frac{1}{2}}T^{\frac{3}{2}}}{N^{\alpha+\delta+\frac{1}{2}}}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}
+(σCBDG12VεxxL2()Nα+12VεxL2()+σ2CBDG12VεxL2()2N)0TkερsεL()s12ds.\displaystyle+\Big{(}\sigma\frac{C_{\mathrm{BDG}}^{\frac{1}{2}}\left\lVert V^{\varepsilon}_{xx}\right\rVert_{L^{2}(\mathbb{R})}}{N^{\alpha+\frac{1}{2}}}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}+\sigma\frac{2C_{\mathrm{BDG}}^{\frac{1}{2}}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}}{N}\Big{)}\int\limits_{0}^{T}\left\lVert k^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}s^{\frac{1}{2}}\,\mathrm{d}s.
+2C(γ)CBDG12σN12+γVεxL2()2(kεL()23T32+0TkερsεL()s12ds).\displaystyle+\frac{2C(\gamma)C_{\mathrm{BDG}}^{\frac{1}{2}}\sigma}{N^{\frac{1}{2}+\gamma}}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}\Big{(}\left\lVert k^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}\frac{2}{3}T^{\frac{3}{2}}+\int\limits_{0}^{T}\left\lVert k^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}s^{\frac{1}{2}}\,\mathrm{d}s\Big{)}.
Proof.

We carry out a similar strategy as in the previous Lemma 3.8. Again, we want to split Ω\Omega into a good and bad set. Remember the definition of set BsαB_{s}^{\alpha} in (3.2), we do the estimates on BsαB_{s}^{\alpha} and its complement (Bsα)c(B_{s}^{\alpha})^{c} separately.

On the set BsαB_{s}^{\alpha}: Let ωBsα\omega\in B_{s}^{\alpha}, then we insert the i.i.d. process 𝐘N,ε\mathbf{Y}^{N,\varepsilon} and split the estimate further into two terms

1N2i,j=1NVεx(Xsi,ε(ω))kε(Xsi,ε(ω)Xsj,ε(ω))Vεx((kερsε)ρsε)),\displaystyle\;\frac{1}{N^{2}}\sum\limits_{i,j=1}^{N}\bigg{\langle}V^{\varepsilon}_{x}(\cdot-X_{s}^{i,\varepsilon}(\omega))k^{\varepsilon}(X_{s}^{i,\varepsilon}(\omega)-X_{s}^{j,\varepsilon}(\omega))-V^{\varepsilon}_{x}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon})),
σNl=1N0sVεx(Xul)dBulL2()\displaystyle\quad\quad\frac{\sigma}{N}\sum\limits_{l=1}^{N}-\int\limits_{0}^{s}V^{\varepsilon}_{x}(\cdot-X_{u}^{l})\,\mathrm{d}B_{u}^{l}\bigg{\rangle}_{L^{2}(\mathbb{R})}
=\displaystyle= 1N2i,j=1NVεx(Xsi,ε(ω))(kε(Xsi,ε(ω)Xsj,ε(ω))kε(Ysi,ε(ω)Ysj,ε(ω))),\displaystyle\;\frac{1}{N^{2}}\sum\limits_{i,j=1}^{N}\bigg{\langle}V^{\varepsilon}_{x}(\cdot-X_{s}^{i,\varepsilon}(\omega))(k^{\varepsilon}(X_{s}^{i,\varepsilon}(\omega)-X_{s}^{j,\varepsilon}(\omega))-k^{\varepsilon}(Y_{s}^{i,\varepsilon}(\omega)-Y_{s}^{j,\varepsilon}(\omega))),
σNl=1N0sVεx(Xul)dBulL2()\displaystyle\quad\quad\frac{\sigma}{N}\sum\limits_{l=1}^{N}-\int\limits_{0}^{s}V^{\varepsilon}_{x}(\cdot-X_{u}^{l})\,\mathrm{d}B_{u}^{l}\bigg{\rangle}_{L^{2}(\mathbb{R})}
+1N2i,j=1NVεx(Xsi,ε(ω))kε(Ysi,ε(ω)Ysj,ε(ω))Vεx((kερsε)ρsε)),\displaystyle+\frac{1}{N^{2}}\sum\limits_{i,j=1}^{N}\bigg{\langle}V^{\varepsilon}_{x}(\cdot-X_{s}^{i,\varepsilon}(\omega))k^{\varepsilon}(Y_{s}^{i,\varepsilon}(\omega)-Y_{s}^{j,\varepsilon}(\omega))-V^{\varepsilon}_{x}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon})),
σNl=1N0sVεx(Xul)dBulL2()\displaystyle\quad\quad\frac{\sigma}{N}\sum\limits_{l=1}^{N}-\int\limits_{0}^{s}V^{\varepsilon}_{x}(\cdot-X_{u}^{l})\,\mathrm{d}B_{u}^{l}\bigg{\rangle}_{L^{2}(\mathbb{R})}
=\displaystyle= IIs1(ω)+IIs2(ω).\displaystyle\;II_{s}^{1}(\omega)+II_{s}^{2}(\omega).

Further, utilizing the property of the set BsαB_{s}^{\alpha} and the Burkholder–Davis–Gundy inequality we obtain

(3.14) 𝔼(sup0tT0t|IIs1(ω)|𝟙(Bsα)ds)\displaystyle\;\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int\limits_{0}^{t}|II_{s}^{1}(\omega)|\mathbbm{1}_{(B_{s}^{\alpha})}\,\mathrm{d}s\bigg{)}
\displaystyle\leq 𝔼(0T1N2i,j=1N|Vεx(Xsi,ε(ω))(kε(Xsi,ε(ω)Xsj,ε(ω))kε(Ysi,ε(ω)Ysj,ε(ω)))|,\displaystyle\;\mathbb{E}\bigg{(}\int\limits_{0}^{T}\frac{1}{N^{2}}\sum\limits_{i,j=1}^{N}\bigg{\langle}\Big{|}V^{\varepsilon}_{x}(\cdot-X_{s}^{i,\varepsilon}(\omega))(k^{\varepsilon}(X_{s}^{i,\varepsilon}(\omega)-X_{s}^{j,\varepsilon}(\omega))-k^{\varepsilon}(Y_{s}^{i,\varepsilon}(\omega)-Y_{s}^{j,\varepsilon}(\omega)))\Big{|},
|σNl=1N0sVεx(Xul)dBul|L2()𝟙(Bsα)ds)\displaystyle\quad\bigg{|}\frac{\sigma}{N}\sum\limits_{l=1}^{N}\int\limits_{0}^{s}V^{\varepsilon}_{x}(\cdot-X_{u}^{l})\,\mathrm{d}B_{u}^{l}\bigg{|}\bigg{\rangle}_{L^{2}(\mathbb{R})}\mathbbm{1}_{(B_{s}^{\alpha})}\,\mathrm{d}s\bigg{)}
\displaystyle\leq  2𝔼(0T1Ni=1NkεxL()|Vεx(Xsi,ε(ω))|max1iN|Xsi,ε(ω)Ysi,ε(ω)|,\displaystyle\;2\mathbb{E}\bigg{(}\int\limits_{0}^{T}\frac{1}{N}\sum\limits_{i=1}^{N}\bigg{\langle}\left\lVert k^{\varepsilon}_{x}\right\rVert_{L^{\infty}(\mathbb{R})}|V^{\varepsilon}_{x}(\cdot-X_{s}^{i,\varepsilon}(\omega))|\max\limits_{1\leq i\leq N}|X_{s}^{i,\varepsilon}(\omega)-Y_{s}^{i,\varepsilon}(\omega)|,
|σNl=1N0sVεx(Xul)dBul|L2()𝟙(Bsα)ds)\displaystyle\quad\bigg{|}\frac{\sigma}{N}\sum\limits_{l=1}^{N}\int\limits_{0}^{s}V^{\varepsilon}_{x}(\cdot-X_{u}^{l})\,\mathrm{d}B_{u}^{l}\bigg{|}\bigg{\rangle}_{L^{2}(\mathbb{R})}\mathbbm{1}_{(B_{s}^{\alpha})}\,\mathrm{d}s\bigg{)}
\displaystyle\leq 2σkεxL()Nα1Ni=1N0T𝔼(|Vεx(Xsi,ε(ω))|,|1Nl=1N0sVεx(Xul)dBul|L2())ds\displaystyle\;\frac{2\sigma\left\lVert k^{\varepsilon}_{x}\right\rVert_{L^{\infty}(\mathbb{R})}}{N^{\alpha}}\frac{1}{N}\sum\limits_{i=1}^{N}\int\limits_{0}^{T}\mathbb{E}\bigg{(}\bigg{\langle}|V^{\varepsilon}_{x}(\cdot-X_{s}^{i,\varepsilon}(\omega))|,\bigg{|}\frac{1}{N}\sum\limits_{l=1}^{N}\int\limits_{0}^{s}V^{\varepsilon}_{x}(\cdot-X_{u}^{l})\,\mathrm{d}B_{u}^{l}\bigg{|}\bigg{\rangle}_{L^{2}(\mathbb{R})}\bigg{)}\,\mathrm{d}s
=\displaystyle= 2σkεxL()NαCBDG12VεxL2()2T321N12=2σT32CBDG12Nα+12VεxL2()2kεxL(),\displaystyle\;\frac{2\sigma\left\lVert k^{\varepsilon}_{x}\right\rVert_{L^{\infty}(\mathbb{R})}}{N^{\alpha}}C_{\mathrm{BDG}}^{\frac{1}{2}}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}T^{\frac{3}{2}}\frac{1}{N^{\frac{1}{2}}}=\frac{2\sigma T^{\frac{3}{2}}C_{\mathrm{BDG}}^{\frac{1}{2}}}{N^{\alpha+\frac{1}{2}}}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}\left\lVert k^{\varepsilon}_{x}\right\rVert_{L^{\infty}(\mathbb{R})},

where we have used the estimate

(3.15) 1Ni=1N0T𝔼(|Vεx(Xsi,ε(ω))|,|1Nl=1N0sVεx(Xul)dBul|L2())ds\displaystyle\;\frac{1}{N}\sum\limits_{i=1}^{N}\int\limits_{0}^{T}\mathbb{E}\bigg{(}\bigg{\langle}|V^{\varepsilon}_{x}(\cdot-X_{s}^{i,\varepsilon}(\omega))|,\bigg{|}\frac{1}{N}\sum\limits_{l=1}^{N}\int\limits_{0}^{s}V^{\varepsilon}_{x}(\cdot-X_{u}^{l})\,\mathrm{d}B_{u}^{l}\bigg{|}\bigg{\rangle}_{L^{2}(\mathbb{R})}\bigg{)}\,\mathrm{d}s
\displaystyle\leq 1Ni=1N0T𝔼(|Vεx(yXsi,ε(ω))|2)12𝔼(|1Nl=1N0sVεx(yXul)dBul|2)12dyds\displaystyle\;\frac{1}{N}\sum\limits_{i=1}^{N}\int\limits_{0}^{T}\int_{\mathbb{R}}\mathbb{E}(|V^{\varepsilon}_{x}(y-X_{s}^{i,\varepsilon}(\omega))|^{2})^{\frac{1}{2}}\mathbb{E}\bigg{(}\bigg{|}\frac{1}{N}\sum\limits_{l=1}^{N}\int\limits_{0}^{s}V^{\varepsilon}_{x}(y-X_{u}^{l})\,\mathrm{d}B_{u}^{l}\bigg{|}^{2}\bigg{)}^{\frac{1}{2}}\,\mathrm{d}y\,\mathrm{d}s
\displaystyle\leq 1Ni=1N0T𝔼(|Vεx(yXsi,ε(ω))|2)12CBDG12N𝔼(l=1N0s|Vεx(yXul)|2du)12dyds\displaystyle\;\frac{1}{N}\sum\limits_{i=1}^{N}\int\limits_{0}^{T}\int_{\mathbb{R}}\mathbb{E}(|V^{\varepsilon}_{x}(y-X_{s}^{i,\varepsilon}(\omega))|^{2})^{\frac{1}{2}}\frac{C_{\mathrm{BDG}}^{\frac{1}{2}}}{N}\mathbb{E}\bigg{(}\sum\limits_{l=1}^{N}\int\limits_{0}^{s}|V^{\varepsilon}_{x}(y-X_{u}^{l})|^{2}\,\mathrm{d}u\bigg{)}^{\frac{1}{2}}\,\mathrm{d}y\,\mathrm{d}s
\displaystyle\leq CBDG121N2i=1N0T(𝔼(|Vεx(yXsi,ε(ω))|2)dy)12\displaystyle\;C_{\mathrm{BDG}}^{\frac{1}{2}}\frac{1}{N^{2}}\sum\limits_{i=1}^{N}\int\limits_{0}^{T}\bigg{(}\int_{\mathbb{R}}\mathbb{E}(|V^{\varepsilon}_{x}(y-X_{s}^{i,\varepsilon}(\omega))|^{2})\,\mathrm{d}y\bigg{)}^{\frac{1}{2}}
(𝔼(l=1N0s|Vεx(yXul)|2du)dy)12ds\displaystyle\quad\cdot\bigg{(}\int_{\mathbb{R}}\mathbb{E}\bigg{(}\sum\limits_{l=1}^{N}\int\limits_{0}^{s}|V^{\varepsilon}_{x}(y-X_{u}^{l})|^{2}\,\mathrm{d}u\bigg{)}\,\mathrm{d}y\bigg{)}^{\frac{1}{2}}\,\mathrm{d}s
=\displaystyle= CBDG12VεxL2()2T321N12.\displaystyle\;C_{\mathrm{BDG}}^{\frac{1}{2}}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}T^{\frac{3}{2}}\frac{1}{N^{\frac{1}{2}}}.

This completes the estimate of IIs1(ω)II_{s}^{1}(\omega) on the set BsαB_{s}^{\alpha}. Next, for ωBsα\omega\in B_{s}^{\alpha} we rewrite IIs2(ω)II_{s}^{2}(\omega) in the following way

IIs2(ω)\displaystyle\;II_{s}^{2}(\omega)
=\displaystyle= 1N2i,j=1NVεx(Xsi,ε(ω))kε(Ysi,ε(ω)Ysj,ε(ω))Vεx((kερsε)ρsε)),\displaystyle\;\frac{1}{N^{2}}\sum\limits_{i,j=1}^{N}\bigg{\langle}V^{\varepsilon}_{x}(\cdot-X_{s}^{i,\varepsilon}(\omega))k^{\varepsilon}(Y_{s}^{i,\varepsilon}(\omega)-Y_{s}^{j,\varepsilon}(\omega))-V^{\varepsilon}_{x}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon})),
σNl=1N0sVεx(Xul)dBulL2()\displaystyle\quad\frac{\sigma}{N}\sum\limits_{l=1}^{N}-\int\limits_{0}^{s}V^{\varepsilon}_{x}(\cdot-X_{u}^{l})\,\mathrm{d}B_{u}^{l}\bigg{\rangle}_{L^{2}(\mathbb{R})}
=\displaystyle= σN2i,j=1NVεx(Xsi,ε(ω))(kε(Ysi,ε(ω)Ysj,ε(ω))(kερsε)(Ysi,ε(ω)))\displaystyle\;-\frac{\sigma}{N^{2}}\sum\limits_{i,j=1}^{N}\bigg{\langle}V^{\varepsilon}_{x}(\cdot-X_{s}^{i,\varepsilon}(\omega))(k^{\varepsilon}(Y_{s}^{i,\varepsilon}(\omega)-Y_{s}^{j,\varepsilon}(\omega))-(k^{\varepsilon}*\rho_{s}^{\varepsilon})(Y_{s}^{i,\varepsilon}(\omega)))
+(Vεx(Xsi,ε(ω))Vεx(Ysi,ε(ω)))(kερsε)(Ysi,ε(ω))\displaystyle\quad+(V^{\varepsilon}_{x}(\cdot-X_{s}^{i,\varepsilon}(\omega))-V^{\varepsilon}_{x}(\cdot-Y_{s}^{i,\varepsilon}(\omega)))(k^{\varepsilon}*\rho_{s}^{\varepsilon})(Y_{s}^{i,\varepsilon}(\omega))
+Vεx(Ysi,ε(ω))(kερsε)(Ysi,ε(ω))Vεx((kερsε)ρsε)),1Nl=1N0sVεx(Xul)dBulL2()\displaystyle\quad+V^{\varepsilon}_{x}(\cdot-Y_{s}^{i,\varepsilon}(\omega))(k^{\varepsilon}*\rho_{s}^{\varepsilon})(Y_{s}^{i,\varepsilon}(\omega))-V^{\varepsilon}_{x}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon})),\frac{1}{N}\sum\limits_{l=1}^{N}\int\limits_{0}^{s}V^{\varepsilon}_{x}(\cdot-X_{u}^{l})\,\mathrm{d}B_{u}^{l}\bigg{\rangle}_{L^{2}(\mathbb{R})}
=\displaystyle= σ(IIs21(ω)+IIs22(ω)+IIs23(ω)).\displaystyle\;\sigma(II_{s}^{21}(\omega)+II_{s}^{22}(\omega)+II_{s}^{23}(\omega)).

For the first term IIs21(ω)II_{s}^{21}(\omega), applying the the property of the set BsαB_{s}^{\alpha}, we find with the help of the estimate (3.15) for the stochastic term that

(3.16) 𝔼(sup0tT0t|IIs21(ω)|ds)\displaystyle\;\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int\limits_{0}^{t}|II_{s}^{21}(\omega)|\,\mathrm{d}s\bigg{)}
\displaystyle\leq 𝔼(sup0tT0t1Ni=1N|Vεx(Xsi,ε(ω))||1Nj=1Nkε(Ysi,ε(ω)Ysj,ε(ω))(kερsε)(Ysi,ε(ω))|,\displaystyle\;\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int\limits_{0}^{t}\frac{1}{N}\sum\limits_{i=1}^{N}\bigg{\langle}|V^{\varepsilon}_{x}(\cdot-X_{s}^{i,\varepsilon}(\omega))|\bigg{|}\frac{1}{N}\sum\limits_{j=1}^{N}k^{\varepsilon}(Y_{s}^{i,\varepsilon}(\omega)-Y_{s}^{j,\varepsilon}(\omega))-(k^{\varepsilon}*\rho_{s}^{\varepsilon})(Y_{s}^{i,\varepsilon}(\omega))\bigg{|},
|1Nl=1N0sVεx(Xul)dBul|L2()ds)\displaystyle\quad\;\bigg{|}\frac{1}{N}\sum\limits_{l=1}^{N}\int\limits_{0}^{s}V^{\varepsilon}_{x}(\cdot-X_{u}^{l})\,\mathrm{d}B_{u}^{l}\bigg{|}\bigg{\rangle}_{L^{2}(\mathbb{R})}\,\mathrm{d}s\bigg{)}
\displaystyle\leq N(α+δ)0T1Ni=1N𝔼(|Vεx(yXsi,ε(ω))1Nl=1N0sVεx(yXul)dBul|)dyds\displaystyle\;N^{-(\alpha+\delta)}\int\limits_{0}^{T}\frac{1}{N}\sum\limits_{i=1}^{N}\int_{\mathbb{R}}\mathbb{E}\bigg{(}\bigg{|}V^{\varepsilon}_{x}(y-X_{s}^{i,\varepsilon}(\omega))\frac{1}{N}\sum\limits_{l=1}^{N}\int\limits_{0}^{s}V^{\varepsilon}_{x}(y-X_{u}^{l})\,\mathrm{d}B_{u}^{l}\bigg{|}\bigg{)}\,\mathrm{d}y\,\mathrm{d}s
\displaystyle\leq N(α+δ)CBDG12VεxL2()2T321N12=CBDG12T32Nα+δ+12VεxL2()2,\displaystyle\;N^{-(\alpha+\delta)}C_{\mathrm{BDG}}^{\frac{1}{2}}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}T^{\frac{3}{2}}\frac{1}{N^{\frac{1}{2}}}=\frac{C_{\mathrm{BDG}}^{\frac{1}{2}}T^{\frac{3}{2}}}{N^{\alpha+\delta+\frac{1}{2}}}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2},

where we used Fubini’s Theorem in the second step. For the term IIs22(ω)II_{s}^{22}(\omega) compute

|IIs22(ω)|\displaystyle\;|II_{s}^{22}(\omega)|
\displaystyle\leq 1Ni=1N|Vεx(Xsi,ε(ω))Vεx(Ysi,ε(ω)))(kερsε)(Ysi,ε(ω)|,\displaystyle\;\frac{1}{N}\sum\limits_{i=1}^{N}\bigg{\langle}|V^{\varepsilon}_{x}(\cdot-X_{s}^{i,\varepsilon}(\omega))-V^{\varepsilon}_{x}(\cdot-Y_{s}^{i,\varepsilon}(\omega)))(k^{\varepsilon}*\rho_{s}^{\varepsilon})(Y_{s}^{i,\varepsilon}(\omega)|,
|1Nl=1N0sVεx(Xul)dBul|L2()\displaystyle\quad\quad\bigg{|}\frac{1}{N}\sum\limits_{l=1}^{N}\int\limits_{0}^{s}V^{\varepsilon}_{x}(\cdot-X_{u}^{l})\,\mathrm{d}B_{u}^{l}\bigg{|}\bigg{\rangle}_{L^{2}(\mathbb{R})}
\displaystyle\leq 1Ni=1N|01ddxVεx(yYsi,ε(ω)r(Xsi,ε(ω)Ysi,ε(ω)))(Ysi,ε(ω)Xsi,ε(ω))dr|,\displaystyle\;\frac{1}{N}\sum\limits_{i=1}^{N}\bigg{\langle}\bigg{|}\int\limits_{0}^{1}\frac{\mathrm{d}}{\mathrm{d}x}V^{\varepsilon}_{x}\Big{(}y-Y_{s}^{i,\varepsilon}(\omega)-r(X_{s}^{i,\varepsilon}(\omega)-Y_{s}^{i,\varepsilon}(\omega))\Big{)}(Y_{s}^{i,\varepsilon}(\omega)-X_{s}^{i,\varepsilon}(\omega))\,\mathrm{d}r\bigg{|},
|1Nl=1N0sVεx(Xul)dBul|L2()kερsεL()\displaystyle\quad\;\bigg{|}\frac{1}{N}\sum\limits_{l=1}^{N}\int\limits_{0}^{s}V^{\varepsilon}_{x}(\cdot-X_{u}^{l})\,\mathrm{d}B_{u}^{l}\bigg{|}\bigg{\rangle}_{L^{2}(\mathbb{R})}\left\lVert k^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}
\displaystyle\leq kερsεL()Nα1Ni=1N01|ddxVεx(yYsi,ε(ω)r(Xsi,ε(ω)Ysi,ε(ω)))|dr,\displaystyle\;\left\lVert k^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}N^{-\alpha}\frac{1}{N}\sum\limits_{i=1}^{N}\bigg{\langle}\int\limits_{0}^{1}\Big{|}\frac{\mathrm{d}}{\mathrm{d}x}V^{\varepsilon}_{x}\Big{(}y-Y_{s}^{i,\varepsilon}(\omega)-r(X_{s}^{i,\varepsilon}(\omega)-Y_{s}^{i,\varepsilon}(\omega))\Big{)}\Big{|}\,\mathrm{d}r,
|1Nl=1N0sVεx(Xul)dBul|L2()\displaystyle\quad\;\bigg{|}\frac{1}{N}\sum\limits_{l=1}^{N}\int\limits_{0}^{s}V^{\varepsilon}_{x}(\cdot-X_{u}^{l})\,\mathrm{d}B_{u}^{l}\bigg{|}\bigg{\rangle}_{L^{2}(\mathbb{R})}
\displaystyle\leq kερsεL()Nα1Ni=1N01|ddxVεx(yYsi,ε(ω)r(Xsi,ε(ω)Ysi,ε(ω)))|drL2()\displaystyle\;\left\lVert k^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}N^{-\alpha}\frac{1}{N}\sum\limits_{i=1}^{N}\bigg{\|}\int\limits_{0}^{1}\Big{|}\frac{\mathrm{d}}{\mathrm{d}x}V^{\varepsilon}_{x}\Big{(}y-Y_{s}^{i,\varepsilon}(\omega)-r(X_{s}^{i,\varepsilon}(\omega)-Y_{s}^{i,\varepsilon}(\omega))\Big{)}\Big{|}\,\mathrm{d}r\bigg{\|}_{L^{2}(\mathbb{R})}
1Nl=1N0sVεx(Xul)dBulL2()\displaystyle\quad\;\bigg{\|}\frac{1}{N}\sum\limits_{l=1}^{N}\int\limits_{0}^{s}V^{\varepsilon}_{x}(\cdot-X_{u}^{l})\,\mathrm{d}B_{u}^{l}\bigg{\|}_{L^{2}(\mathbb{R})}
\displaystyle\leq kερsεL()NαVεxxL2()1Nl=1N0sVεx(Xul)dBulL2(),\displaystyle\;\left\lVert k^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}N^{-\alpha}\left\lVert V^{\varepsilon}_{xx}\right\rVert_{L^{2}(\mathbb{R})}\bigg{\|}\frac{1}{N}\sum\limits_{l=1}^{N}\int\limits_{0}^{s}V^{\varepsilon}_{x}(\cdot-X_{u}^{l})\,\mathrm{d}B_{u}^{l}\bigg{\|}_{L^{2}(\mathbb{R})},

where we utilized the property of BsαB_{s}^{\alpha} in the third step, followed by the application of Hölder’s inequality and Minkowski’s inequality. Consequently, applying the Burkholder–Davis–Gundy inequality we obtain

(3.17) 𝔼(sup0tT0t|IIs22(ω)|ds)\displaystyle\;\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int\limits_{0}^{t}|II_{s}^{22}(\omega)|\,\mathrm{d}s\bigg{)}
\displaystyle\leq NαVεxxL2()𝔼(sup0tT0tkερsεL()1Nl=1N0sVεx(Xul)dBulL2()ds)\displaystyle\;N^{-\alpha}\left\lVert V^{\varepsilon}_{xx}\right\rVert_{L^{2}(\mathbb{R})}\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int\limits_{0}^{t}\left\lVert k^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}\bigg{\|}\frac{1}{N}\sum\limits_{l=1}^{N}\int\limits_{0}^{s}V^{\varepsilon}_{x}(\cdot-X_{u}^{l})\,\mathrm{d}B_{u}^{l}\bigg{\|}_{L^{2}(\mathbb{R})}\,\mathrm{d}s\bigg{)}
\displaystyle\leq Nα1VεxxL2()0TkερsεL()(𝔼(|l=1N0sVεx(yXul)dBul|2)dy)12ds\displaystyle\;N^{-\alpha-1}\left\lVert V^{\varepsilon}_{xx}\right\rVert_{L^{2}(\mathbb{R})}\int\limits_{0}^{T}\left\lVert k^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}\bigg{(}\int_{\mathbb{R}}\mathbb{E}\bigg{(}\bigg{|}\sum\limits_{l=1}^{N}\int\limits_{0}^{s}V^{\varepsilon}_{x}(y-X_{u}^{l})\,\mathrm{d}B_{u}^{l}\bigg{|}^{2}\bigg{)}\,\mathrm{d}y\bigg{)}^{\frac{1}{2}}\,\mathrm{d}s
\displaystyle\leq CBDG12Nα1VεxxL2()0TkερsεL()(𝔼(l=1N0s|Vεx(yXul)|2du)dy)12ds\displaystyle\;C_{\mathrm{BDG}}^{\frac{1}{2}}N^{-\alpha-1}\left\lVert V^{\varepsilon}_{xx}\right\rVert_{L^{2}(\mathbb{R})}\int\limits_{0}^{T}\left\lVert k^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}\bigg{(}\int_{\mathbb{R}}\mathbb{E}\bigg{(}\sum\limits_{l=1}^{N}\int\limits_{0}^{s}|V^{\varepsilon}_{x}(y-X_{u}^{l})|^{2}\,\mathrm{d}u\bigg{)}\,\mathrm{d}y\bigg{)}^{\frac{1}{2}}\,\mathrm{d}s
\displaystyle\leq CBDG12VεxxL2()Nα+12VεxL2()0TkερsεL()s12ds.\displaystyle\;\frac{C_{\mathrm{BDG}}^{\frac{1}{2}}\left\lVert V^{\varepsilon}_{xx}\right\rVert_{L^{2}(\mathbb{R})}}{N^{\alpha+\frac{1}{2}}}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}\int\limits_{0}^{T}\left\lVert k^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}s^{\frac{1}{2}}\,\mathrm{d}s.

For II23s(ω)II^{23}_{s}(\omega) we use again the Burkholder–Davis–Gundy inequality to estimate the stochastic integral and by the law of large number argument similar to the term I23I^{23} in Lemma 3.8, and obtain

(3.18) 𝔼(sup0tT0t|II23s(ω)|ds)\displaystyle\;\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int\limits_{0}^{t}|II^{23}_{s}(\omega)|\,\mathrm{d}s\bigg{)}
\displaystyle\leq 0T𝔼(|1Ni=1NVεx(yYsi,ε(ω))(kερsε)(Ysi,ε(ω))Vεx((kερsε)ρsε))(y)|,\displaystyle\;\int\limits_{0}^{T}\int_{\mathbb{R}}\mathbb{E}\bigg{(}\bigg{|}\frac{1}{N}\sum\limits_{i=1}^{N}V^{\varepsilon}_{x}(y-Y_{s}^{i,\varepsilon}(\omega))(k^{\varepsilon}*\rho_{s}^{\varepsilon})(Y_{s}^{i,\varepsilon}(\omega))-V^{\varepsilon}_{x}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon}))(y)\bigg{|},
|1Nl=1N0sVεx(yXul)dBul|)dyds\displaystyle\quad\quad\bigg{|}\frac{1}{N}\sum\limits_{l=1}^{N}\int\limits_{0}^{s}V^{\varepsilon}_{x}(y-X_{u}^{l})\,\mathrm{d}B_{u}^{l}\bigg{|}\bigg{)}\,\mathrm{d}y\,\mathrm{d}s
\displaystyle\leq 0T(𝔼(|1Ni=1NVεx(yYsi,ε(ω))(kερsε)(Ysi,ε(ω))Vεx((kερsε)ρsε))(y)|2)dy)12\displaystyle\;\int\limits_{0}^{T}\bigg{(}\int_{\mathbb{R}}\mathbb{E}\bigg{(}\bigg{|}\frac{1}{N}\sum\limits_{i=1}^{N}V^{\varepsilon}_{x}(y-Y_{s}^{i,\varepsilon}(\omega))(k^{\varepsilon}*\rho_{s}^{\varepsilon})(Y_{s}^{i,\varepsilon}(\omega))-V^{\varepsilon}_{x}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon}))(y)\bigg{|}^{2}\bigg{)}\,\mathrm{d}y\bigg{)}^{\frac{1}{2}}
(𝔼(|1Nl=1N0sVεx(yXul)dBul|2)dy)12ds\displaystyle\quad\quad\cdot\bigg{(}\int_{\mathbb{R}}\mathbb{E}\bigg{(}\bigg{|}\frac{1}{N}\sum\limits_{l=1}^{N}\int\limits_{0}^{s}V^{\varepsilon}_{x}(y-X_{u}^{l})\,\mathrm{d}B_{u}^{l}\bigg{|}^{2}\bigg{)}\,\mathrm{d}y\bigg{)}^{\frac{1}{2}}\,\mathrm{d}s
\displaystyle\leq  2CBDG120TN1/2kερsεL()VεxL2()s12N1/2VεxL2()ds\displaystyle\;2C_{\mathrm{BDG}}^{\frac{1}{2}}\int\limits_{0}^{T}N^{-1/2}\left\lVert k^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}s^{\frac{1}{2}}N^{-1/2}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}\,\mathrm{d}s
\displaystyle\leq 2CBDG12VεxL2()2N0TkερsεL()s12ds.\displaystyle\;\frac{2C_{\mathrm{BDG}}^{\frac{1}{2}}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}}{N}\int\limits_{0}^{T}\left\lVert k^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}s^{\frac{1}{2}}\,\mathrm{d}s.

This completes the estimate on the set BsαB_{s}^{\alpha}, namely

(3.19) 𝔼(sup0tT0t(|IIs1(ω)|+|IIs2(ω)|)𝟙(Bsα)ds)\displaystyle\;\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int\limits_{0}^{t}(|II_{s}^{1}(\omega)|+|II_{s}^{2}(\omega)|)\mathbbm{1}_{(B_{s}^{\alpha})}\,\mathrm{d}s\bigg{)}
\displaystyle\leq 2σT32CBDG12Nα+12VεxL2()2kεxL()+σCBDG12T32Nα+δ+12VεxL2()2\displaystyle\;\frac{2\sigma T^{\frac{3}{2}}C_{\mathrm{BDG}}^{\frac{1}{2}}}{N^{\alpha+\frac{1}{2}}}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}\left\lVert k^{\varepsilon}_{x}\right\rVert_{L^{\infty}(\mathbb{R})}+\sigma\frac{C_{\mathrm{BDG}}^{\frac{1}{2}}T^{\frac{3}{2}}}{N^{\alpha+\delta+\frac{1}{2}}}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}
+(σCBDG12VεxxL2()Nα+12VεxL2()+σ2CBDG12VεxL2()2N)0TkερsεL()s12ds.\displaystyle+\Big{(}\sigma\frac{C_{\mathrm{BDG}}^{\frac{1}{2}}\left\lVert V^{\varepsilon}_{xx}\right\rVert_{L^{2}(\mathbb{R})}}{N^{\alpha+\frac{1}{2}}}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}+\sigma\frac{2C_{\mathrm{BDG}}^{\frac{1}{2}}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}}{N}\Big{)}\int\limits_{0}^{T}\left\lVert k^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}s^{\frac{1}{2}}\,\mathrm{d}s.

On the set (Bsα)c(B_{s}^{\alpha})^{\mathrm{c}}: Using ((Bsα)c)C(γ)Nγ\mathbb{P}((B_{s}^{\alpha})^{\mathrm{c}})\leq C(\gamma)N^{-\gamma} for all γ>0\gamma>0 by Assumption 2.6, the Burkholder–Davis–Gundy inequality, Hölder’s inequality, we obtain

𝔼(sup0tT0t|1N2i,j=1NVεx(Xsi,ε)kε(Xsi,εXsj,ε)Vεx((kερsε)ρsε),\displaystyle\;\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int\limits_{0}^{t}\bigg{|}\bigg{\langle}\frac{1}{N^{2}}\sum\limits_{i,j=1}^{N}V^{\varepsilon}_{x}(\cdot-X_{s}^{i,\varepsilon})k^{\varepsilon}(X_{s}^{i,\varepsilon}-X_{s}^{j,\varepsilon})-V^{\varepsilon}_{x}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon}),
σNl=1N0sVεx(Xul)dBulL2()|𝟙(Bsα)cds)\displaystyle\quad\quad\frac{\sigma}{N}\sum\limits_{l=1}^{N}-\int\limits_{0}^{s}V^{\varepsilon}_{x}(\cdot-X_{u}^{l})\,\mathrm{d}B_{u}^{l}\bigg{\rangle}_{L^{2}(\mathbb{R})}\bigg{|}\mathbbm{1}_{(B_{s}^{\alpha})^{\mathrm{c}}}\,\mathrm{d}s\bigg{)}
\displaystyle\leq σN2i,j=1N0T𝔼(𝟙(Bsα)c|(Vεx(yXsi,ε)kε(Xsi,εXsj,ε)Vεx((kερsε)ρsε))(y)\displaystyle\;\frac{\sigma}{N^{2}}\sum\limits_{i,j=1}^{N}\int\limits_{0}^{T}\int_{\mathbb{R}}\mathbb{E}\bigg{(}\mathbbm{1}_{(B_{s}^{\alpha})^{\mathrm{c}}}\bigg{|}(V^{\varepsilon}_{x}(y-X_{s}^{i,\varepsilon})k^{\varepsilon}(X_{s}^{i,\varepsilon}-X_{s}^{j,\varepsilon})-V^{\varepsilon}_{x}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon}))(y)
1Nl=1N0sVεx(yXul)dBul|)dyds\displaystyle\quad\cdot\;\frac{1}{N}\sum\limits_{l=1}^{N}\int\limits_{0}^{s}V^{\varepsilon}_{x}(y-X_{u}^{l})\,\mathrm{d}B_{u}^{l}\bigg{|}\bigg{)}\,\mathrm{d}y\,\mathrm{d}s
\displaystyle\leq σN2i,j=1N0T(𝔼(𝟙(Bsα)c|(Vεx(yXsi,ε)kε(Xsi,εXsj,ε)Vεx((kερsε)(y)ρsε)(y)|2)dy)12\displaystyle\;\frac{\sigma}{N^{2}}\sum\limits_{i,j=1}^{N}\int\limits_{0}^{T}\bigg{(}\int_{\mathbb{R}}\mathbb{E}\bigg{(}\mathbbm{1}_{(B_{s}^{\alpha})^{\mathrm{c}}}|(V^{\varepsilon}_{x}(y-X_{s}^{i,\varepsilon})k^{\varepsilon}(X_{s}^{i,\varepsilon}-X_{s}^{j,\varepsilon})-V^{\varepsilon}_{x}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})(y)\rho_{s}^{\varepsilon})(y)|^{2}\bigg{)}\mathrm{d}y\bigg{)}^{\frac{1}{2}}
(𝔼(|1Nl=1N0sVεx(yXul)dBul|2)dy)12ds\displaystyle\quad\cdot\bigg{(}\int_{\mathbb{R}}\mathbb{E}\bigg{(}\bigg{|}\frac{1}{N}\sum\limits_{l=1}^{N}\int\limits_{0}^{s}V^{\varepsilon}_{x}(y-X_{u}^{l})\,\mathrm{d}B_{u}^{l}\bigg{|}^{2}\bigg{)}\,\mathrm{d}y\bigg{)}^{\frac{1}{2}}\,\mathrm{d}s
\displaystyle\leq 2CBDG12σNi=1N0T(𝔼(𝟙(Bsα)c(|(Vεx(yXsi,ε)|2kεL()2+|Vεx((kερsε)ρsε)(y)|2))dy)12\displaystyle\frac{2C_{\mathrm{BDG}}^{\frac{1}{2}}\sigma}{N}\sum\limits_{i=1}^{N}\int\limits_{0}^{T}\bigg{(}\int_{\mathbb{R}}\mathbb{E}\bigg{(}\mathbbm{1}_{(B_{s}^{\alpha})^{\mathrm{c}}}(|(V^{\varepsilon}_{x}(y-X_{s}^{i,\varepsilon})|^{2}\left\lVert k^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}^{2}+|V^{\varepsilon}_{x}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon})(y)|^{2})\bigg{)}\,\mathrm{d}y\bigg{)}^{\frac{1}{2}}
1N(𝔼(l=1N0T|Vεx(yXul)|2du)dy)12ds\displaystyle\quad\cdot\frac{1}{N}\bigg{(}\int_{\mathbb{R}}\mathbb{E}\bigg{(}\sum\limits_{l=1}^{N}\int\limits_{0}^{T}|V^{\varepsilon}_{x}(y-X_{u}^{l})|^{2}\,\mathrm{d}u\bigg{)}\,\mathrm{d}y\bigg{)}^{\frac{1}{2}}\,\mathrm{d}s
\displaystyle\leq 2CBDG12σN12VεxL2()0T𝔼(𝟙(Bsα)c(VεxL2()2kεL()2+Vεx((kερsε)ρsε)L2()2))12s12ds\displaystyle\;\frac{2C_{\mathrm{BDG}}^{\frac{1}{2}}\sigma}{N^{\frac{1}{2}}}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}\int\limits_{0}^{T}\mathbb{E}\bigg{(}\mathbbm{1}_{(B_{s}^{\alpha})^{\mathrm{c}}}\big{(}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}\left\lVert k^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}^{2}+\left\lVert V^{\varepsilon}_{x}*((k^{\varepsilon}*\rho_{s}^{\varepsilon})\rho_{s}^{\varepsilon})\right\rVert_{L^{2}(\mathbb{R})}^{2}\big{)}\bigg{)}^{\frac{1}{2}}s^{\frac{1}{2}}\,\mathrm{d}s
\displaystyle\leq 2C(γ)CBDG12σN12+γVεxL2()2(kεL()23T32+0TkερsεL()s12ds).\displaystyle\;\frac{2C(\gamma)C_{\mathrm{BDG}}^{\frac{1}{2}}\sigma}{N^{\frac{1}{2}+\gamma}}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}\Big{(}\left\lVert k^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}\frac{2}{3}T^{\frac{3}{2}}+\int\limits_{0}^{T}\left\lVert k^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}s^{\frac{1}{2}}\,\mathrm{d}s\Big{)}.

This completes the estimate on the set (Bsα)c(B_{s}^{\alpha})^{\mathrm{c}} and we have shown our Lemma. ∎

Lemma 3.10 (Stochastic Integral Inequality).

Under the assumptions of Theorem 3.3 we have the following L2L^{2}-estimate for the stochastic integral,

(3.20) 4σ2𝔼(sup0tT0tσNi=1N0sVεxx(Xui)dBuL2()2ds)\displaystyle 4\sigma^{2}\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int\limits_{0}^{t}\bigg{\|}\frac{\sigma}{N}\sum\limits_{i=1}^{N}\int\limits_{0}^{s}V^{\varepsilon}_{xx}(\cdot-X_{u}^{i})\,\mathrm{d}B_{u}\bigg{\|}_{L^{2}(\mathbb{R})}^{2}\,\mathrm{d}s\bigg{)}\leq T32NVεxxL2()2.\displaystyle\;\frac{T^{\frac{3}{2}}}{N}\left\lVert V^{\varepsilon}_{xx}\right\rVert_{L^{2}(\mathbb{R})}^{2}.
Proof.

An application of the Burkholder–Davis–Gundy inequality implies

𝔼(sup0tT0t1Ni=1N0sVεxx(Xui)dBuL2()2ds)\displaystyle\;\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\int\limits_{0}^{t}\left\lVert\frac{1}{N}\sum\limits_{i=1}^{N}\int\limits_{0}^{s}V^{\varepsilon}_{xx}(\cdot-X_{u}^{i})\,\mathrm{d}B_{u}\right\rVert_{L^{2}(\mathbb{R})}^{2}\,\mathrm{d}s\bigg{)}
\displaystyle\leq 0T𝔼(|1Ni=1N0sVεxx(yXui)dBu|2)dy)ds\displaystyle\;\int\limits_{0}^{T}\int_{\mathbb{R}}\mathbb{E}\bigg{(}\bigg{|}\frac{1}{N}\sum\limits_{i=1}^{N}\int\limits_{0}^{s}V^{\varepsilon}_{xx}(y-X_{u}^{i})\,\mathrm{d}B_{u}\bigg{|}^{2}\bigg{)}\,\mathrm{d}y\bigg{)}\,\mathrm{d}s
\displaystyle\leq 1N20T𝔼(i=1N0s|Vεxx(yXui)|2du)dy)dsT32NVεxxL2()2.\displaystyle\;\frac{1}{N^{2}}\int\limits_{0}^{T}\int_{\mathbb{R}}\mathbb{E}\bigg{(}\sum\limits_{i=1}^{N}\int\limits_{0}^{s}|V^{\varepsilon}_{xx}(y-X_{u}^{i})|^{2}\,\mathrm{d}u\bigg{)}\,\mathrm{d}y\bigg{)}\,\mathrm{d}s\leq\;\frac{T^{\frac{3}{2}}}{N}\left\lVert V^{\varepsilon}_{xx}\right\rVert_{L^{2}(\mathbb{R})}^{2}.

Continuation of the proof of theorem 3.3.

We are ready to input the estimates from above lemmata in the the inequality (3.4). We find

𝔼(sup0tTVεμtN,εVερtεL2()2)+σ28𝔼(0TVxεμsN,εVxερsεL2()ds)\displaystyle\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\left\lVert V^{\varepsilon}*\mu_{t}^{N,\varepsilon}-V^{\varepsilon}*\rho_{t}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}\bigg{)}+\frac{\sigma^{2}}{8}\mathbb{E}\bigg{(}\int\limits_{0}^{T}\left\lVert V_{x}^{\varepsilon}*\mu_{s}^{N,\varepsilon}-V_{x}^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}\,\mathrm{d}s\bigg{)}
\displaystyle\leq 2NVεL2()2+16TkxεL()2VεL2()2σ2N2α+4TVεL2()2σ2N2(α+δ)\displaystyle\;\frac{2}{N}\left\lVert V^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}+\frac{16T\left\lVert k_{x}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}^{2}\left\lVert V^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}}{\sigma^{2}N^{2\alpha}}+\frac{4T\left\lVert V^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}}{\sigma^{2}N^{2(\alpha+\delta)}}
+(4VεxL2()2N2ασ2+16VεL2()2Nσ2)0TkερsεL()2ds\displaystyle+\Big{(}\frac{4\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}}{N^{2\alpha}\sigma^{2}}+\frac{16\left\lVert V^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}}{N\sigma^{2}}\Big{)}\int\limits_{0}^{T}\left\lVert k^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}^{2}\,\mathrm{d}s
+C(γ)TNγ(VεL2()2kεL()2+VεxL2()2)\displaystyle+\frac{C(\gamma)T}{N^{\gamma}}\Big{(}\left\lVert V^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}\left\lVert k^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}^{2}+\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}\Big{)}
+2σT32CBDG12Nα+12VεxL2()2kεxL()+σCBDG12T32Nα+δ+12VεxL2()2\displaystyle+\frac{2\sigma T^{\frac{3}{2}}C_{\mathrm{BDG}}^{\frac{1}{2}}}{N^{\alpha+\frac{1}{2}}}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}\left\lVert k^{\varepsilon}_{x}\right\rVert_{L^{\infty}(\mathbb{R})}+\sigma\frac{C_{\mathrm{BDG}}^{\frac{1}{2}}T^{\frac{3}{2}}}{N^{\alpha+\delta+\frac{1}{2}}}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}
+(σCBDG12VεxxL2()Nα+12VεxL2()+σ2CBDG12VεxL2()2N)0TkερsεL()s12ds.\displaystyle+\Big{(}\sigma\frac{C_{\mathrm{BDG}}^{\frac{1}{2}}\left\lVert V^{\varepsilon}_{xx}\right\rVert_{L^{2}(\mathbb{R})}}{N^{\alpha+\frac{1}{2}}}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}+\sigma\frac{2C_{\mathrm{BDG}}^{\frac{1}{2}}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}}{N}\Big{)}\int\limits_{0}^{T}\left\lVert k^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}s^{\frac{1}{2}}\,\mathrm{d}s.
+2C(γ)CBDG12σN12+γVεxL2()2(kεL()23T32+0TkερsεL()s12ds).\displaystyle+\frac{2C(\gamma)C_{\mathrm{BDG}}^{\frac{1}{2}}\sigma}{N^{\frac{1}{2}+\gamma}}\left\lVert V^{\varepsilon}_{x}\right\rVert_{L^{2}(\mathbb{R})}^{2}\Big{(}\left\lVert k^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}\frac{2}{3}T^{\frac{3}{2}}+\int\limits_{0}^{T}\left\lVert k^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}s^{\frac{1}{2}}\,\mathrm{d}s\Big{)}.
+T32NVεxxL2()2.\displaystyle+\frac{T^{\frac{3}{2}}}{N}\left\lVert V^{\varepsilon}_{xx}\right\rVert_{L^{2}(\mathbb{R})}^{2}.

The above estimate is the most general one we obtain. In the following we simplify it to derive a usable estimates. In the process we may loose some convergence rate, depending on the concrete problem at hand. Noticing that by mass conservation

kερsεL2(0,T;L())kεL()ρsεL2(0,T;L1())TkεL(),\left\lVert k^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{2}(0,T;L^{\infty}(\mathbb{R}))}\leq\left\lVert k^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}\left\lVert\rho_{s}^{\varepsilon}\right\rVert_{L^{2}(0,T;L^{1}(\mathbb{R}))}\leq T\left\lVert k^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})},

by keeping all the NN and ε\varepsilon dependent terms and put all the other constants into a universal constant CC, which depends on TT, σ\sigma, γ\gamma, CBDGC_{\mathrm{BDG}}, we obtain

𝔼(sup0tTVεμtN,εVερtεL2()2)+σ28𝔼(0TVxεμsN,εVxερsεL2()ds)\displaystyle\;\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\left\lVert V^{\varepsilon}*\mu_{t}^{N,\varepsilon}-V^{\varepsilon}*\rho_{t}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}\bigg{)}+\frac{\sigma^{2}}{8}\mathbb{E}\bigg{(}\int\limits_{0}^{T}\left\lVert V_{x}^{\varepsilon}*\mu_{s}^{N,\varepsilon}-V_{x}^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}\,\mathrm{d}s\bigg{)}
\displaystyle\leq CN(VεH1()2kεL2+VεxxL2()2)+CVεH1()2(1+kεL()2)Nγ\displaystyle\;\frac{C}{N}(\left\lVert V^{\varepsilon}\right\rVert_{H^{1}(\mathbb{R})}^{2}\left\lVert k^{\varepsilon}\right\rVert_{L^{\infty}}^{2}+\left\lVert V^{\varepsilon}_{xx}\right\rVert_{L^{2}(\mathbb{R})}^{2})+\frac{C\left\lVert V^{\varepsilon}\right\rVert_{H^{1}(\mathbb{R})}^{2}(1+\left\lVert k^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}^{2})}{N^{\gamma}}
+kεxL()2VεL2()2+kεL()2VxεL2()2+VεL2()2N2α\displaystyle+\frac{\left\lVert k^{\varepsilon}_{x}\right\rVert_{L^{\infty}(\mathbb{R})}^{2}\left\lVert V^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}+\left\lVert k^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}^{2}\left\lVert V_{x}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}+\left\lVert V^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}}{N^{2\alpha}}
+CVxεL2()2(1+kxεL())+VxεL2()VεxxL2()kεL(d)Nα+12.\displaystyle+C\frac{\left\lVert V_{x}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}(1+\left\lVert k_{x}^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})})+\left\lVert V_{x}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}\left\lVert V^{\varepsilon}_{xx}\right\rVert_{L^{2}(\mathbb{R})}\left\lVert k^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R}^{d})}}{N^{\alpha+\frac{1}{2}}}.

In the above estimates, α(0,12)\alpha\in(0,\frac{1}{2}) and δ>0\delta>0 are also used and the Theorem is proven. ∎

In the our main setting kε=WεVεk^{\varepsilon}=W^{\varepsilon}*V^{\varepsilon} we provide the following rough estimate.

Corollary 3.11.

Let kε=WεVεk^{\varepsilon}=W^{\varepsilon}*V^{\varepsilon} and Wε,VεW^{\varepsilon},V^{\varepsilon} be admissible with rates aW,aVa_{W},a_{V}. If Theorem 3.3 holds, then

𝔼(sup0tTVεμtN,εVερtεL2()2)+σ28𝔼(0TVxεμsN,εVxερsεL2()ds)\displaystyle\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\left\lVert V^{\varepsilon}*\mu_{t}^{N,\varepsilon}-V^{\varepsilon}*\rho_{t}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}\bigg{)}+\frac{\sigma^{2}}{8}\mathbb{E}\bigg{(}\int\limits_{0}^{T}\left\lVert V_{x}^{\varepsilon}*\mu_{s}^{N,\varepsilon}-V_{x}^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}\,\mathrm{d}s\bigg{)}
CNε2aW+4aV+CN2αε2aW+4aV+CNα+12εaW+3aV+CNγε2aW+4aV.\displaystyle\leq\frac{C}{N\varepsilon^{2a_{W}+4a_{V}}}+\frac{C}{N^{2\alpha}\varepsilon^{2a_{W}+4a_{V}}}+\frac{C}{N^{\alpha+\frac{1}{2}}\varepsilon^{a_{W}+3a_{V}}}+\frac{C}{N^{\gamma}\varepsilon^{2a_{W}+4a_{V}}}.
Proof.

Estimating all norms of VεV^{\varepsilon} by VεH2()CεaV\left\lVert V^{\varepsilon}\right\rVert_{H^{2}(\mathbb{R})}\leq C\varepsilon^{-a_{V}} and using Young’s inequality to find

kεL()+kεxL()2WεL2()VεH2()CεaWaV.\displaystyle\left\lVert k^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}+\left\lVert k^{\varepsilon}_{x}\right\rVert_{L^{\infty}(\mathbb{R})}\leq 2\left\lVert W^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}\left\lVert V^{\varepsilon}\right\rVert_{H^{2}(\mathbb{R})}\leq C\varepsilon^{-a_{W}-a_{V}}.

Hence the right hand side of the main inequality in Theorem 3.3 can be estimated by

CNε2aW+4aV+CN2αε2aW+4aV+CNα+12εaW+3aV+CNγε2aW+4aV.\frac{C}{N\varepsilon^{2a_{W}+4a_{V}}}+\frac{C}{N^{2\alpha}\varepsilon^{2a_{W}+4a_{V}}}+\frac{C}{N^{\alpha+\frac{1}{2}}\varepsilon^{a_{W}+3a_{V}}}+\frac{C}{N^{\gamma}\varepsilon^{2a_{W}+4a_{V}}}.

Now, that we have proven our main estimate, we are ready to demonstrate the relative entropy estimates by combining Theorem 3.3 and Lemma 3.1. We start with the first main result of this paper

Proof of Theorem 2.9.

We combine the assumptions of Theorem 3.3 and the results from Lemma 3.1, Theorem 3.3 and Corollary 3.11, to find a small β1βα\beta_{1}\leq\beta_{\alpha} so that for 0<ββ10<\beta\leq\beta_{1}, ε=Nβ\varepsilon=N^{-\beta} and small 0<λ10<\lambda\ll 1

N(ρtN,ε|ρtN,ε)CN12+λ=o(1N).{\mathcal{H}}_{N}(\rho_{t}^{N,\varepsilon}|\rho_{t}^{\otimes N,\varepsilon})\leq\;\frac{C}{N^{\frac{1}{2}+\lambda}}=o\bigg{(}\frac{1}{\sqrt{N}}\bigg{)}.

This allows us to demonstrate strong convergence in the L([0,T];L1())L^{\infty}([0,T];L^{1}(\mathbb{R}))-norm. Indeed, let us recall the Csiszár–Kullback–Pinsker inequality [Vil09, Chapter 22], which states that for any mm\in\mathbb{N} and function f,g:mf,g\colon\mathbb{R}^{m}\to\mathbb{R} we have

(3.21) fgL1(m)2mm(f|g)\left\lVert f-g\right\rVert_{L^{1}(\mathbb{R}^{m})}\leq\sqrt{2m{\mathcal{H}}_{m}(f\;|\;g)}

and the relative entropy inequaity [DMM01, Lemma 3.9]

(3.22) m(ρN,m,εt|ρm,εt)2N(ρN,εt|ρtN,ε){\mathcal{H}}_{m}(\rho^{N,m,\varepsilon}_{t}\;|\;\rho^{\otimes m,\varepsilon}_{t})\leq 2{\mathcal{H}}_{N}(\rho^{N,\varepsilon}_{t}\;|\;\rho_{t}^{\otimes N,\varepsilon})

for mNm\leq N. Consequently,

ρtN,2,ερtερtεL1(2)222(ρtN,1,ε|ρtε)4N(ρtN,ε|ρtN,ε)=o(1N).\left\lVert\rho_{t}^{N,2,\varepsilon}-\rho_{t}^{\varepsilon}\otimes\rho_{t}^{\varepsilon}\right\rVert_{L^{1}(\mathbb{R}^{2})}^{2}\leq 2{\mathcal{H}}_{2}(\rho_{t}^{N,1,\varepsilon}|\rho_{t}^{\varepsilon})\leq 4{\mathcal{H}}_{N}(\rho_{t}^{N,\varepsilon}|\rho_{t}^{\otimes N,\varepsilon})=o\bigg{(}\frac{1}{\sqrt{N}}\bigg{)}.

In the case kε=(WεVε)xk^{\varepsilon}=(W^{\varepsilon}*V^{\varepsilon})_{x} the estimate (2.18) is derived analogously. The key is to recognize that we actually derived an estimate on the derivative of VεV^{\varepsilon}, which we have not used so far. In the case of kε=(WεVε)xk^{\varepsilon}=(W^{\varepsilon}*V^{\varepsilon})_{x} we utilize it and as a result we obtain the same convergence rates. The estimate for the modulated energy follows also directly from equality (3.2), Young’s inequality and an application of Theorem 3.3 for VεV^{\varepsilon} and W^ε\hat{W}^{\varepsilon} under the assumption that Wε,VεW^{\varepsilon},V^{\varepsilon} are strongly admissible. ∎

3.3. Special Choices of WεW^{\varepsilon} and VεV^{\varepsilon}

We present a series of corollaries for Theorem 3.3 for different choices of VεV^{\varepsilon}. In most applications we want to take a mollified sequence. In the special case Vε=JεV^{\varepsilon}=J^{\varepsilon} we obtain the following corollary.

Corollary 3.12.

Suppose Theorem 3.3 holds true. Let Vε=JεV^{\varepsilon}=J^{\varepsilon} be a mollification, then for ε=Nβ\varepsilon=N^{-\beta} with some β<βα\beta<\beta_{\alpha} and kεL()εak\left\lVert k^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}\leq\varepsilon^{-a_{k}}, kεxL()εak1\left\lVert k^{\varepsilon}_{x}\right\rVert_{L^{\infty}(\mathbb{R})}\leq\varepsilon^{-a_{k}-1} for some ak>0a_{k}>0, then we obtain the following L2L^{2}-estimate ,

𝔼(sup0tTJεμtN,εJερtεL2()2)+σ22𝔼(0TJxεμsN,εJxερsεL2()ds)\displaystyle\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\left\lVert J^{\varepsilon}*\mu_{t}^{N,\varepsilon}-J^{\varepsilon}*\rho_{t}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}\bigg{)}+\frac{\sigma^{2}}{2}\mathbb{E}\bigg{(}\int\limits_{0}^{T}\left\lVert J_{x}^{\varepsilon}*\mu_{s}^{N,\varepsilon}-J_{x}^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}\,\mathrm{d}s\bigg{)}
CN2αβ+CNα+1/24β+CN2αβ(2ak+2)β+CNα+123β(ak+1)β\displaystyle\quad\leq\;\frac{C}{N^{2\alpha-\beta}}+\frac{C}{N^{\alpha+1/2-4\beta}}+\frac{C}{N^{2\alpha-\beta-(2a_{k}+2)\beta}}+\frac{C}{N^{\alpha+\frac{1}{2}-3\beta-(a_{k}+1)\beta}}
+CN2α3βakβ+CNγN(3+2ak)β\displaystyle\quad\quad+\frac{C}{N^{2\alpha-3\beta-a_{k}\beta}}+\frac{C}{N^{\gamma}N^{(3+2a_{k})\beta}}
CN2αβ(2ak+2)β+CNα+123β(ak+1)β+CN2α3βakβ+CNγN(3+2ak)β\displaystyle\quad\leq\;\frac{C}{N^{2\alpha-\beta-(2a_{k}+2)\beta}}+\frac{C}{N^{\alpha+\frac{1}{2}-3\beta-(a_{k}+1)\beta}}+\frac{C}{N^{2\alpha-3\beta-a_{k}\beta}}+\frac{C}{N^{\gamma}N^{(3+2a_{k})\beta}}

for a constant CC, which depends on TT, σ\sigma, γ\gamma, CBDGC_{\mathrm{BDG}}. In particular if kL()k\in L^{\infty}(\mathbb{R}) and kε=(ζε(Jεk))Jεk^{\varepsilon}=(\zeta^{\varepsilon}(J^{\varepsilon}*k))*J^{\varepsilon} the above estimate holds with ε=Nβ\varepsilon=N^{-\beta} and ak=0a_{k}=0.

Proof.

If Vε=JεV^{\varepsilon}=J^{\varepsilon}, we obtains easily that

JεHm()\displaystyle\left\lVert J^{\varepsilon}\right\rVert_{H^{m}(\mathbb{R})} =1ε1[(1+|ξε|2)m2[J](ξ)](ε)L2()\displaystyle=\frac{1}{\varepsilon}\left\lVert\mathcal{F}^{-1}\bigg{[}\bigg{(}1+\bigg{|}\frac{\xi}{\varepsilon}\bigg{|}^{2}\bigg{)}^{\frac{m}{2}}\mathcal{F}[J](\xi)\bigg{]}\bigg{(}\frac{\cdot}{\varepsilon}\bigg{)}\right\rVert_{L^{2}(\mathbb{R})}
=1ε1/2+m1[(ε2+|ξ|2)m2[J](ξ)]L2()Cε1/2+m.\displaystyle=\frac{1}{\varepsilon^{1/2+m}}\left\lVert\mathcal{F}^{-1}[(\varepsilon^{2}+|\xi|^{2})^{\frac{m}{2}}\mathcal{F}[J](\xi)]\right\rVert_{L^{2}(\mathbb{R})}\leq\frac{C}{\varepsilon^{1/2+m}}.

Therefore, we obtained with ε=Nβ\varepsilon=N^{-\beta},

𝔼(sup0tTJεμtN,εJερtεL2()2)+σ22𝔼(0TJxεμsN,εJxερsεL2()ds)\displaystyle\mathbb{E}\bigg{(}\sup\limits_{0\leq t\leq T}\left\lVert J^{\varepsilon}*\mu_{t}^{N,\varepsilon}-J^{\varepsilon}*\rho_{t}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}\bigg{)}+\frac{\sigma^{2}}{2}\mathbb{E}\bigg{(}\int\limits_{0}^{T}\left\lVert J_{x}^{\varepsilon}*\mu_{s}^{N,\varepsilon}-J_{x}^{\varepsilon}*\rho_{s}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}\,\mathrm{d}s\bigg{)}
CN2αβ+CNα+1/24β+CN2αβε2ak+2+CNα+123βεak+1+CN2α3βεak+CNγε3+2ak\displaystyle\quad\leq\;\frac{C}{N^{2\alpha-\beta}}+\frac{C}{N^{\alpha+1/2-4\beta}}+\frac{C}{N^{2\alpha-\beta}\varepsilon^{2a_{k}+2}}+\frac{C}{N^{\alpha+\frac{1}{2}-3\beta}\varepsilon^{a_{k}+1}}+\frac{C}{N^{2\alpha-3\beta}\varepsilon^{a_{k}}}+\frac{C}{N^{\gamma}\varepsilon^{3+2a_{k}}}

The second claim follows by Young’s inequality and the scaling of the mollifier. More precisely,

JεWm,1()\displaystyle\left\lVert J^{\varepsilon}\right\rVert_{W^{m,1}(\mathbb{R})} =1ε1[(1+|ξ|2)m2[J](ξ)]L1()Cεm\displaystyle=\frac{1}{\varepsilon}\left\lVert\mathcal{F}^{-1}[(1+|\xi|^{2})^{\frac{m}{2}}\mathcal{F}[J](\xi)]\right\rVert_{L^{1}(\mathbb{R})}\leq\frac{C}{\varepsilon^{m}}

Corollary 3.13.

Suppose Assumptions 2.52.6 hold for α(14,12)\alpha\in(\frac{1}{4},\frac{1}{2}) and suppose the bounded force kk has the approximation kε=WεVεk^{\varepsilon}=W^{\varepsilon}*V^{\varepsilon} with Wε=ζε(kJε)W^{\varepsilon}=\zeta^{\varepsilon}(k*J^{\varepsilon}) and Vε=JεV^{\varepsilon}=J^{\varepsilon}. Then for ε=Nβ\varepsilon=N^{-\beta} and β<min(α6,110(4α1),βα)\beta<\min\bigg{(}\frac{\alpha}{6},\frac{1}{10}(4\alpha-1),\beta_{\alpha}\bigg{)}, there exists an 0<λ10<\lambda\ll 1 such that

supt[0,T]N(ρtN,ε|ρtN,ε)CN12+λ\sup\limits_{t\in[0,T]}{\mathcal{H}}_{N}(\rho_{t}^{N,\varepsilon}|\rho_{t}^{\otimes N,\varepsilon})\leq\frac{C}{N^{\frac{1}{2}+\lambda}}

for a constant CC, which depends on TT, σ\sigma, γ\gamma, CBDGC_{\mathrm{BDG}}.

Proof.

Let us start by estimating WεL2()2\left\lVert W^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2} in inequality (3.3),

WεL2()2=|ζε(x)|2|k(xy)Jε(y)dy|2dx4ε2kL()=4N2βkL()\left\lVert W^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}^{2}=\int_{\mathbb{R}}|\zeta^{\varepsilon}(x)|^{2}\bigg{|}\int_{\mathbb{R}}k(x-y)J^{\varepsilon}(y)\,\mathrm{d}y\bigg{|}^{2}\,\mathrm{d}x\leq 4\varepsilon^{-2}\left\lVert k\right\rVert_{L^{\infty}(\mathbb{R})}=4N^{2\beta}\left\lVert k\right\rVert_{L^{\infty}(\mathbb{R})}

Now, applying Corollary 3.12 to inequality (3.3), keeping track of the powers, we obtain the result. More precisely, notice that we have to first fix α(14,12)\alpha\in(\frac{1}{4},\frac{1}{2}), then choos a βJ\beta_{J} such that the terms for 0<ββJ0<\beta\leq\beta_{J} smaller than CN12+λ\frac{C}{N^{\frac{1}{2}+\lambda}}, and then fix γ\gamma such that the estimate holds. ∎

Next, we provide a similar corollary in the case the force kk is a potential and has a convolution structure.

Corollary 3.14.

Suppose Assumptions 2.52.6 hold for α(14,12)\alpha\in(\frac{1}{4},\frac{1}{2}). Let the force kk be given by a potential k=(WV)xk=(W*V)_{x} with W,VH12()W,V\in H^{\frac{1}{2}}(\mathbb{R}) and its approximation is given by kε=((WJε)(VJε))xk^{\varepsilon}=((W*J^{\varepsilon})*(V*J^{\varepsilon}))_{x}. Then for ε=Nβ\varepsilon=N^{-\beta} and β<min(13,α14,βα)\beta<\min\bigg{(}\frac{1}{3},\alpha-\frac{1}{4},\beta_{\alpha}\bigg{)}, there exists a 0<λ10<\lambda\ll 1 such that

supt[0,T]N(ρtN,ε|ρtN,ε)\displaystyle\sup\limits_{t\in[0,T]}{\mathcal{H}}_{N}(\rho_{t}^{N,\varepsilon}|\rho_{t}^{\otimes N,\varepsilon}) CN12+λ,\displaystyle\leq\frac{C}{N^{\frac{1}{2}+\lambda}},
supt[0,T]|𝒦N(ρtN,ε|ρtN,ε)|\displaystyle\sup\limits_{t\in[0,T]}|{\mathcal{K}}_{N}(\rho_{t}^{N,\varepsilon}|\rho_{t}^{\otimes N,\varepsilon})| CN12+λ.\displaystyle\leq\frac{C}{N^{\frac{1}{2}+\lambda}}.

for a constant CC, which depends on TT, σ\sigma, γ\gamma, CBDGC_{\mathrm{BDG}}.

Proof.

Since WH12()W\in H^{\frac{1}{2}}(\mathbb{R}) we know that WJεL2()W*J^{\varepsilon}\in L^{2}(\mathbb{R}) and therefore we only need to estimate the L2L^{2}-norm for Vεx=(VJε)xV^{\varepsilon}_{x}=(V*J^{\varepsilon})_{x} in inequality (3.3) to obtain the convergence rates. We emphasize that in Theorem 3.3 we also obtained an estimate on the gradient VεV^{\varepsilon}. Consequently, we can use Theorem 3.3 for the function Vε=VJεV^{\varepsilon}=V*J^{\varepsilon}. By the estimate

JεWm,1()\displaystyle\left\lVert J^{\varepsilon}\right\rVert_{W^{m,1}(\mathbb{R})} =1ε1[(1+|ξ|2)m2[J](ξ)]L1()Cεm\displaystyle=\frac{1}{\varepsilon}\left\lVert\mathcal{F}^{-1}[(1+|\xi|^{2})^{\frac{m}{2}}\mathcal{F}[J](\xi)]\right\rVert_{L^{1}(\mathbb{R})}\leq\frac{C}{\varepsilon^{m}}
JεxWm,1()\displaystyle\left\lVert J^{\varepsilon}_{x}\right\rVert_{W^{m,1}(\mathbb{R})} =1ε1[(1+|ξ|2)m2[Jx](ξ)]L1()Cεm+1\displaystyle=\frac{1}{\varepsilon}\left\lVert\mathcal{F}^{-1}[(1+|\xi|^{2})^{\frac{m}{2}}\mathcal{F}[J_{x}](\xi)]\right\rVert_{L^{1}(\mathbb{R})}\leq\frac{C}{\varepsilon^{m+1}}

for any m0m\geq 0 we know that

VεL2()VL2()C\displaystyle\left\lVert V^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}\leq\left\lVert V\right\rVert_{L^{2}(\mathbb{R})}\leq C
VxεL2()VH12()JεW12,1()Cε12\displaystyle\left\lVert V_{x}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}\leq\left\lVert V\right\rVert_{H^{\frac{1}{2}}(\mathbb{R})}\left\lVert J^{\varepsilon}\right\rVert_{W^{\frac{1}{2},1}(\mathbb{R})}\leq\frac{C}{\varepsilon^{\frac{1}{2}}}
(3.23) VxxεL2()VH12()JxεW12,1()Cε32\displaystyle\left\lVert V_{xx}^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R})}\leq\left\lVert V\right\rVert_{H^{\frac{1}{2}}(\mathbb{R})}\left\lVert J_{x}^{\varepsilon}\right\rVert_{W^{\frac{1}{2},1}(\mathbb{R})}\leq\frac{C}{\varepsilon^{\frac{3}{2}}}
kεL()WεH12()VεH12()C\displaystyle\left\lVert k^{\varepsilon}\right\rVert_{L^{\infty}(\mathbb{R})}\leq\left\lVert W^{\varepsilon}\right\rVert_{H^{\frac{1}{2}}(\mathbb{R})}\left\lVert V^{\varepsilon}\right\rVert_{H^{\frac{1}{2}}(\mathbb{R})}\leq C
kεxL()WεH1()VεH1()Cε.\displaystyle\left\lVert k^{\varepsilon}_{x}\right\rVert_{L^{\infty}(\mathbb{R})}\leq\left\lVert W^{\varepsilon}\right\rVert_{H^{1}(\mathbb{R})}\left\lVert V^{\varepsilon}\right\rVert_{H^{1}(\mathbb{R})}\leq\frac{C}{\varepsilon}.

Plugging in all estimates with ε=Nβ\varepsilon=N^{-\beta} into Theorem 3.3 and having equality (3.2) in mind we obtain the rate of β\beta and the estimate on the modulated energy. ∎

We have now shown in two cases how to derive explicit estimates on the relative entropy N(ρtN,ε|ρtN,ε){\mathcal{H}}_{N}(\rho_{t}^{N,\varepsilon}|\rho_{t}^{\otimes N,\varepsilon}) with the help of Theorem 3.3. In general, if the function Wε,VεW^{\varepsilon},V^{\varepsilon} have low regularity, we need to mollify them to make them admissible. Hence, we borrow the necessary regularity from JJ and consequently, get higher rates of NN in our estimates. Compare for instance Corollary 3.13 and Corollary 3.14. Therefore the estimate by using the regularity of the JεJ^{\varepsilon} term, will lead to weaker convergence rates. The benefit is of course that one does not require a potential field and the convolution structure of the potential.

4. De-regularization of the high dimensional PDE and the limiting PDE

The goal of this section is to prove Theorem 2.11, i.e. the strong form of propagation of chaos on the PDE level in the L1L^{1}-norm. For the de-regularization of Liouville equation (2.8) we need kL()k\in L^{\infty}(\mathbb{R}). We take the following approximation kε=(ζε(kJε))Jε)k^{\varepsilon}=(\zeta^{\varepsilon}(k*J^{\varepsilon}))*J^{\varepsilon}). We need convergence results between ρtN,1,ε\rho_{t}^{N,1,\varepsilon} and ρtN,1\rho_{t}^{N,1} as well as ρtε\rho_{t}^{\varepsilon} and ρt\rho_{t}. The latter convergence was shown in [CNP23, Section 3]. More precisely,

(4.1) limε0ρερL1([0,T];L1())=0.\lim\limits_{\varepsilon\to 0}\left\lVert\rho^{\varepsilon}-\rho\right\rVert_{L^{1}([0,T];L^{1}(\mathbb{R}))}=0.

It remains to show that the approximated Liouville equation converges in entropy to the Liouville equation. An application of inequality (3.21) implies also the L1L^{1}-convergence.

Lemma 4.1.

Let kL()k\in L^{\infty}(\mathbb{R}), ρN,ε\rho^{N,\varepsilon} be the solution of the regularized Liouville equation (2.8) and ρN\rho^{N} the solution of the Liouville equation (2.6). Then, we have

supt[0,T]N(ρN,εt|ρNt)+σ24Ni=1N0TNρsN,ε|xilog(ρN,εsρNs)|2dXNds\displaystyle\;\sup\limits_{t\in[0,T]}{\mathcal{H}}_{N}(\rho^{N,\varepsilon}_{t}\;|\;\rho^{N}_{t})+\frac{\sigma^{2}}{4N}\sum\limits_{i=1}^{N}\int\limits_{0}^{T}\int\limits_{\mathbb{R}^{N}}\rho_{s}^{N,\varepsilon}\bigg{|}\partial_{x_{i}}\log\bigg{(}\frac{\rho^{N,\varepsilon}_{s}}{\rho^{N}_{s}}\bigg{)}\bigg{|}^{2}\,\mathrm{d}\mathrm{X}^{N}\,\mathrm{d}s
\displaystyle\leq CTkL()2supt[0,T]N(ρN,εt|ρN,εt)+2CkL()2ρsερsL1([0,T];L1())\displaystyle\;CT\left\lVert k\right\rVert_{L^{\infty}(\mathbb{R})}^{2}\sup\limits_{t\in[0,T]}\sqrt{{\mathcal{H}}_{N}(\rho^{N,\varepsilon}_{t}\;|\;\rho^{\otimes N,\varepsilon}_{t})}+2C\left\lVert k\right\rVert_{L^{\infty}(\mathbb{R})}^{2}\left\lVert\rho_{s}^{\varepsilon}-\rho_{s}\right\rVert_{L^{1}([0,T];L^{1}(\mathbb{R}))}
+0T2|k(x1x2)kε(x1x2)|2ρs(x1)ρs(x2)dx1dx2ds.\displaystyle+\int\limits_{0}^{T}\int\limits_{\mathbb{R}^{2}}|k(x_{1}-x_{2})-k^{\varepsilon}(x_{1}-x_{2})|^{2}\rho_{s}(x_{1})\rho_{s}(x_{2})\,\mathrm{d}x_{1}\,\mathrm{d}x_{2}\,\mathrm{d}s.

In particular, the last term vanishes by dominated convergence.

Proof.

We start by computing the time derivative of ρN,εtlog(ρN,εtρNt)\rho^{N,\varepsilon}_{t}\log\bigg{(}\frac{\rho^{N,\varepsilon}_{t}}{\rho^{N}_{t}}\bigg{)}. We have

N(ρN,εt|ρNt)=N(ρN,ε|ρN)(0)+0tddsN(ρN,εs|ρNs)ds\displaystyle{\mathcal{H}}_{N}(\rho^{N,\varepsilon}_{t}\;|\;\rho^{N}_{t})={\mathcal{H}}_{N}(\rho^{N,\varepsilon}\;|\;\rho^{N})(0)+\int\limits_{0}^{t}\frac{\,\mathrm{d}}{\,\mathrm{d}s}{\mathcal{H}}_{N}(\rho^{N,\varepsilon}_{s}\;|\;\rho^{N}_{s})\,\mathrm{d}s
=\displaystyle= σ22Ni=1N0tNxiρsN,εxilog(ρN,εsρNs)dXNds\displaystyle\;-\frac{\sigma^{2}}{2N}\sum\limits_{i=1}^{N}\int\limits_{0}^{t}\int\limits_{\mathbb{R}^{N}}\partial_{x_{i}}\rho_{s}^{N,\varepsilon}\partial_{x_{i}}\log\bigg{(}\frac{\rho^{N,\varepsilon}_{s}}{\rho^{N}_{s}}\bigg{)}\,\mathrm{d}\mathrm{X}^{N}\,\mathrm{d}s
1Ni=1N0tN(ρN,εs1Nj=1Nkε(xixj))xilog(ρN,εsρNs)dXNds\displaystyle-\frac{1}{N}\sum\limits_{i=1}^{N}\int\limits_{0}^{t}\int\limits_{\mathbb{R}^{N}}\Bigg{(}\rho^{N,\varepsilon}_{s}\frac{1}{N}\sum\limits_{j=1}^{N}k^{\varepsilon}(x_{i}-x_{j})\Bigg{)}\partial_{x_{i}}\log\bigg{(}\frac{\rho^{N,\varepsilon}_{s}}{\rho^{N}_{s}}\bigg{)}\,\mathrm{d}\mathrm{X}^{N}\,\mathrm{d}s
+σ22Ni=1N0tNxiρsNxi(ρN,εsρNs)dXNds\displaystyle+\frac{\sigma^{2}}{2N}\sum\limits_{i=1}^{N}\int\limits_{0}^{t}\int\limits_{\mathbb{R}^{N}}\partial_{x_{i}}\rho_{s}^{N}\partial_{x_{i}}\bigg{(}\frac{\rho^{N,\varepsilon}_{s}}{\rho^{N}_{s}}\bigg{)}\,\mathrm{d}\mathrm{X}^{N}\,\mathrm{d}s
+1Ni=1N0tN(ρNs1Nj=1Nk(xixj))xi(ρN,εsρNs)dXNds\displaystyle+\frac{1}{N}\sum\limits_{i=1}^{N}\int\limits_{0}^{t}\int\limits_{\mathbb{R}^{N}}\Bigg{(}\rho^{N}_{s}\frac{1}{N}\sum\limits_{j=1}^{N}k(x_{i}-x_{j})\Bigg{)}\partial_{x_{i}}\bigg{(}\frac{\rho^{N,\varepsilon}_{s}}{\rho^{N}_{s}}\bigg{)}\,\mathrm{d}\mathrm{X}^{N}\,\mathrm{d}s
=\displaystyle= σ22Ni=1N0tNρsN,ε|xilog(ρN,εsρNs)|2dXNds\displaystyle\;-\frac{\sigma^{2}}{2N}\sum\limits_{i=1}^{N}\int\limits_{0}^{t}\int\limits_{\mathbb{R}^{N}}\rho_{s}^{N,\varepsilon}\bigg{|}\partial_{x_{i}}\log\bigg{(}\frac{\rho^{N,\varepsilon}_{s}}{\rho^{N}_{s}}\bigg{)}\bigg{|}^{2}\,\mathrm{d}\mathrm{X}^{N}\,\mathrm{d}s
+1N2i,j=1N0tN(k(xixj)kε(xixj))ρN,εsxilog(ρN,εsρNs)dXNds\displaystyle+\frac{1}{N^{2}}\sum\limits_{i,j=1}^{N}\int\limits_{0}^{t}\int\limits_{\mathbb{R}^{N}}(k(x_{i}-x_{j})-k^{\varepsilon}(x_{i}-x_{j}))\rho^{N,\varepsilon}_{s}\partial_{x_{i}}\log\bigg{(}\frac{\rho^{N,\varepsilon}_{s}}{\rho^{N}_{s}}\bigg{)}\,\mathrm{d}\mathrm{X}^{N}\,\mathrm{d}s
\displaystyle\leq σ24Ni=1N0tNρsN,ε|xilog(ρN,εsρNs)|2dXNds\displaystyle\;-\frac{\sigma^{2}}{4N}\sum\limits_{i=1}^{N}\int\limits_{0}^{t}\int\limits_{\mathbb{R}^{N}}\rho_{s}^{N,\varepsilon}\bigg{|}\partial_{x_{i}}\log\bigg{(}\frac{\rho^{N,\varepsilon}_{s}}{\rho^{N}_{s}}\bigg{)}\bigg{|}^{2}\,\mathrm{d}\mathrm{X}^{N}\,\mathrm{d}s
+1σ2N2i,j=1N0tN|k(xixj)kε(xixj)|2ρN,εsdXNds.\displaystyle+\frac{1}{\sigma^{2}N^{2}}\sum\limits_{i,j=1}^{N}\int\limits_{0}^{t}\int\limits_{\mathbb{R}^{N}}|k(x_{i}-x_{j})-k^{\varepsilon}(x_{i}-x_{j})|^{2}\rho^{N,\varepsilon}_{s}\,\mathrm{d}\mathrm{X}^{N}\,\mathrm{d}s.

Now, it is enough to show, that the last term vanishes for NN\to\infty and consequently for ε0\varepsilon\to 0. We start by using the fact that the particle system (2.3) is exchangeable. We obtain

1σ2N2i,j=1N0tN|k(xixj)kε(xixj)|2ρN,εs(XN)dXNds\displaystyle\;\frac{1}{\sigma^{2}N^{2}}\sum\limits_{i,j=1}^{N}\int\limits_{0}^{t}\int\limits_{\mathbb{R}^{N}}|k(x_{i}-x_{j})-k^{\varepsilon}(x_{i}-x_{j})|^{2}\rho^{N,\varepsilon}_{s}(\mathrm{X}^{N})\,\mathrm{d}\mathrm{X}^{N}\,\mathrm{d}s
=\displaystyle= 1σ20t2|k(x1x2)kε(x1x2)|2ρN,2,εs(x1,x2)dx1dx2ds.\displaystyle\;\frac{1}{\sigma^{2}}\int\limits_{0}^{t}\int\limits_{\mathbb{R}^{2}}|k(x_{1}-x_{2})-k^{\varepsilon}(x_{1}-x_{2})|^{2}\rho^{N,2,\varepsilon}_{s}(x_{1},x_{2})\,\mathrm{d}x_{1}\,\mathrm{d}x_{2}\,\mathrm{d}s.

Hence, we obtained an expression in which the dimension does not change in the limit. By applying mass conservation, the Csiszár–Kullback–Pinsker inequality (3.21) and inequality (3.22) we further estimate the term

0t2|k(x1x2)kε(x1x2)|2ρN,2,εs(x1,x2)dx1dx2ds\displaystyle\;\int\limits_{0}^{t}\int\limits_{\mathbb{R}^{2}}|k(x_{1}-x_{2})-k^{\varepsilon}(x_{1}-x_{2})|^{2}\rho^{N,2,\varepsilon}_{s}(x_{1},x_{2})\,\mathrm{d}x_{1}\,\mathrm{d}x_{2}\,\mathrm{d}s
=\displaystyle= 0t2|k(x1x2)kε(x1x2)|2(ρN,2,εs(ρsερsε)(x1,x2))dx1dx2ds\displaystyle\;\int\limits_{0}^{t}\int\limits_{\mathbb{R}^{2}}|k(x_{1}-x_{2})-k^{\varepsilon}(x_{1}-x_{2})|^{2}(\rho^{N,2,\varepsilon}_{s}-(\rho_{s}^{\varepsilon}\otimes\rho_{s}^{\varepsilon})(x_{1},x_{2}))\,\mathrm{d}x_{1}\,\mathrm{d}x_{2}\,\mathrm{d}s
+0t2|k(x1x2)kε(x1x2)|2(ρsερsε)(x1,x2)dx1dx2ds\displaystyle+\int\limits_{0}^{t}\int\limits_{\mathbb{R}^{2}}|k(x_{1}-x_{2})-k^{\varepsilon}(x_{1}-x_{2})|^{2}(\rho_{s}^{\varepsilon}\otimes\rho_{s}^{\varepsilon})(x_{1},x_{2})\,\mathrm{d}x_{1}\,\mathrm{d}x_{2}\,\mathrm{d}s
\displaystyle\leq CkL()20tρN,2,εsρ2,εL1(2)ds+0t2|k(x1x2)kε(x1x2)|2\displaystyle\;C\left\lVert k\right\rVert_{L^{\infty}(\mathbb{R})}^{2}\int\limits_{0}^{t}\left\lVert\rho^{N,2,\varepsilon}_{s}-\rho^{\otimes 2,\varepsilon}\right\rVert_{L^{1}(\mathbb{R}^{2})}\,\mathrm{d}s+\int\limits_{0}^{t}\int\limits_{\mathbb{R}^{2}}|k(x_{1}-x_{2})-k^{\varepsilon}(x_{1}-x_{2})|^{2}
((ρsε(x1)ρs(x1))ρsε(x2)+ρs(x1)(ρsε(x2)ρs(x2))+ρs(x1)ρs(x2))dx1dx2ds\displaystyle\quad\quad\bigg{(}(\rho_{s}^{\varepsilon}(x_{1})-\rho_{s}(x_{1}))\rho_{s}^{\varepsilon}(x_{2})+\rho_{s}(x_{1})(\rho_{s}^{\varepsilon}(x_{2})-\rho_{s}(x_{2}))+\rho_{s}(x_{1})\rho_{s}(x_{2})\bigg{)}\,\mathrm{d}x_{1}\,\mathrm{d}x_{2}\,\mathrm{d}s
\displaystyle\leq CTkL()2supt[0,T]N(ρN,εt|ρN,εt)+2CkL()2ρsερsL1([0,T];L1())\displaystyle\;CT\left\lVert k\right\rVert_{L^{\infty}(\mathbb{R})}^{2}\sup\limits_{t\in[0,T]}\sqrt{{\mathcal{H}}_{N}(\rho^{N,\varepsilon}_{t}\;|\;\rho^{\otimes N,\varepsilon}_{t})}+2C\left\lVert k\right\rVert_{L^{\infty}(\mathbb{R})}^{2}\left\lVert\rho_{s}^{\varepsilon}-\rho_{s}\right\rVert_{L^{1}([0,T];L^{1}(\mathbb{R}))}
+0t2|k(x1x2)kε(x1x2)|2ρs(x1)ρs(x2)dx1dx2ds.\displaystyle+\int\limits_{0}^{t}\int\limits_{\mathbb{R}^{2}}|k(x_{1}-x_{2})-k^{\varepsilon}(x_{1}-x_{2})|^{2}\rho_{s}(x_{1})\rho_{s}(x_{2})\,\mathrm{d}x_{1}\,\mathrm{d}x_{2}\,\mathrm{d}s.

Plugging this estimate into our above entropy calculation and taking the supremum in time proves the lemma. ∎

Combing both implies the strong convergence on the PDE-level of any observable ρN,m\rho^{N,m} to the law ρm\rho^{\otimes m} in the L1()L^{1}(\mathbb{R})-norm.

Proof of theorem 2.11.

For kL()k\in L^{\infty}(\mathbb{R}), let kε=(ζε(kJε))Jεk^{\varepsilon}=(\zeta^{\varepsilon}(k*J^{\varepsilon}))*J^{\varepsilon}, therefore we take Wε=ζε(kεJε)W^{\varepsilon}=\zeta^{\varepsilon}(k^{\varepsilon}*J^{\varepsilon}) and Vε=JεV^{\varepsilon}=J^{\varepsilon} with ε=ε(N)=Nβ\varepsilon=\varepsilon(N)=N^{-\beta}. By assumption of the Theorem (see also [CNP23, Theorem 6.1]), there exists a βα(0,12)\beta_{\alpha}\in(0,\frac{1}{2}) such that for all ββα\beta\leq\beta_{\alpha} the convergence in probability, Assumption 2.5, and the law of large numbers, Assumption 2.6 both hold. Therefore we can apply the result from Corollary 3.13 for 0<β<min(13,α14,βα)0<\beta<\min\bigg{(}\frac{1}{3},\alpha-\frac{1}{4},\beta_{\alpha}\bigg{)} and obtain the convergence of the relative entropy (ρN,εt|ρtN,ε){\mathcal{H}}(\rho^{N,\varepsilon}_{t}\;|\;\rho_{t}^{\otimes N,\varepsilon}) to zero. We can even get better convergence rate β\beta, since we are not interested in the order of convergence of the relative entropy. Applying (3.21) and  (3.22), we obtain

ρN,mρmL1([0,T];L1(m))\displaystyle\;\left\lVert\rho^{N,m}-\rho^{\otimes m}\right\rVert_{L^{1}([0,T];L^{1}(\mathbb{R}^{m}))}
\displaystyle\leq ρN,mρN,m,εL1([0,T];L1(m))+ρN,m,ερm,εL1([0,T];L1(m))\displaystyle\;\left\lVert\rho^{N,m}-\rho^{N,m,\varepsilon}\right\rVert_{L^{1}([0,T];L^{1}(\mathbb{R}^{m}))}+\left\lVert\rho^{N,m,\varepsilon}-\rho^{\otimes m,\varepsilon}\right\rVert_{L^{1}([0,T];L^{1}(\mathbb{R}^{m}))}
+ρm,ερmL1([0,T];L1(m))\displaystyle+\left\lVert\rho^{\otimes m,\varepsilon}-\rho^{\otimes m}\right\rVert_{L^{1}([0,T];L^{1}(\mathbb{R}^{m}))}
\displaystyle\leq 0T2mm(ρN,m,εt|ρtN,m)+2mm(ρN,m,εt|ρm,εt)+ρm,εtρtmL1(m)dt\displaystyle\;\int\limits_{0}^{T}\sqrt{2m{\mathcal{H}}_{m}(\rho^{N,m,\varepsilon}_{t}\;|\;\rho_{t}^{N,m})}+\sqrt{2m{\mathcal{H}}_{m}(\rho^{N,m,\varepsilon}_{t}\;|\;\rho^{\otimes m,\varepsilon}_{t})}+\left\lVert\rho^{\otimes m,\varepsilon}_{t}-\rho_{t}^{\otimes m}\right\rVert_{L^{1}(\mathbb{R}^{m})}\,\mathrm{d}t
\displaystyle\leq 0T4mN(ρN,εt|ρtN)+4mN(ρN,εt|ρN,εt)+ρm,εtρmtL1(m)dt.\displaystyle\;\int\limits_{0}^{T}\sqrt{4m{\mathcal{H}}_{N}(\rho^{N,\varepsilon}_{t}\;|\;\rho_{t}^{N})}+\sqrt{4m{\mathcal{H}}_{N}(\rho^{N,\varepsilon}_{t}\;|\;\rho^{\otimes N,\varepsilon}_{t})}+\left\lVert\rho^{\otimes m,\varepsilon}_{t}-\rho^{\otimes m}_{t}\right\rVert_{L^{1}(\mathbb{R}^{m})}\,\mathrm{d}t.

As mentioned the second term converges to zero. For the first term we use the inequality in Lemma 4.1 together with the fact that the N(ρN,εt|ρtN,ε){\mathcal{H}}_{N}(\rho^{N,\varepsilon}_{t}\;|\;\rho_{t}^{\otimes N,\varepsilon}) converges to zero and the dominated convergence to obtain

lim supN0T4mN(ρN,εt|ρtN)dt\displaystyle\limsup\limits_{N\to\infty}\int\limits_{0}^{T}\sqrt{4m{\mathcal{H}}_{N}(\rho^{N,\varepsilon}_{t}\;|\;\rho_{t}^{N})}\,\mathrm{d}t C(kL(),m)lim supN0TρεtρtL1()12dt\displaystyle\leq C(\left\lVert k\right\rVert_{L^{\infty}(\mathbb{R})},m)\limsup\limits_{N\to\infty}\int\limits_{0}^{T}\left\lVert\rho^{\varepsilon}_{t}-\rho_{t}\right\rVert_{L^{1}(\mathbb{R})}^{\frac{1}{2}}\,\mathrm{d}t
C(kL(),m,T)lim supN(0TρεtρtL1()dt)12\displaystyle\leq C(\left\lVert k\right\rVert_{L^{\infty}(\mathbb{R})},m,T)\limsup\limits_{N\to\infty}\bigg{(}\int\limits_{0}^{T}\left\lVert\rho^{\varepsilon}_{t}-\rho_{t}\right\rVert_{L^{1}(\mathbb{R})}\,\mathrm{d}t\bigg{)}^{\frac{1}{2}}
=0.\displaystyle=0.

where the last equality follows by (4.1). Consequently, it remains to show that the third term vanishes, i.e.

(4.2) lim supNρm,εtρmtL1([0,T];L1(m))=0.\limsup\limits_{N\to\infty}\left\lVert\rho^{\otimes m,\varepsilon}_{t}-\rho^{\otimes m}_{t}\right\rVert_{L^{1}([0,T];L^{1}(\mathbb{R}^{m}))}=0.

Again this follows by (4.1) and an induction argument. Indeed let us assume m=2m=2, then by mass conservation we have

ρ2,εtρ2tL1([0,T];L1(2))\displaystyle\left\lVert\rho^{\otimes 2,\varepsilon}_{t}-\rho^{\otimes 2}_{t}\right\rVert_{L^{1}([0,T];L^{1}(\mathbb{R}^{2}))}
=0T2|(ρεt(x1)ρt(x1))ρεt(x2)+ρt(x1)(ρεt(x2)ρt(x2))|dx1dx2dt\displaystyle\quad=\int\limits_{0}^{T}\int_{\mathbb{R}^{2}}|(\rho^{\varepsilon}_{t}(x_{1})-\rho_{t}(x_{1}))\rho^{\varepsilon}_{t}(x_{2})+\rho_{t}(x_{1})(\rho^{\varepsilon}_{t}(x_{2})-\rho_{t}(x_{2}))|\,\mathrm{d}x_{1}\,\mathrm{d}x_{2}\,\mathrm{d}t
2ρεtρtL1([0,T];L1())N0,\displaystyle\quad\leq 2\left\lVert\rho^{\varepsilon}_{t}-\rho_{t}\right\rVert_{L^{1}([0,T];L^{1}(\mathbb{R}))}\xrightarrow{N\to\infty}\quad 0,

which proves the initial case for the induction. Now, by the same argument one can prove the induction step and therefore equation (4.2). ∎

5. Application

We provide some examples for which Theorem 2.9 can be shown with the same techniques developed in Section 3. In particular we demonstrate the convergence in relative entropy in the attractive Coulomb case on the whole space. Note that the rate of converges may vary across these examples. As stated in Remark 3.4 we only need the existence of approximated PDE (2.9), the particle system (2.4), the convergence in probability of the particle system 𝐗N\mathbf{X}^{N} to the mean-field limit 𝐘N\mathbf{Y}^{N}(Assumption (2.5)) and the law of large numbers (Assumption 2.6). Since we are working on the regularized level, we can often assume the existence of the above results.

Although the result in Theorem 2.9 also works with rotational field, it worth to study directly a convolution type of potential field to achieve better cut-off rate, in other words, to allow bigger β\beta. For a given potential field, the challenging part is to find a convolution structure for the potential described in Section 3. The first idea to obtain interesting kernels, beside the Delta-Distribution, which was given in [Oel87], is to look at infinite divisible distributions. Assume that kε=Uxk^{\varepsilon}=U_{x} is infinitely divisible. Then, there exists a VεV^{\varepsilon} such that Uε=VεVεU^{\varepsilon}=V^{\varepsilon}*V^{\varepsilon}. Hence, if we can approximate the antiderivative of our kernel by a infinitely divisible distribution (multiplied by a constant if necessary) we are able to find candidates for interesting kernels.

Another powerful tool is the Fourier analysis. On the Fourier side the equation kε=VεWεk^{\varepsilon}=V^{\varepsilon}*W^{\varepsilon} becomes

(Wε)=(Vε)(Wε),\mathcal{F}(W^{\varepsilon})=\mathcal{F}(V^{\varepsilon})\mathcal{F}({W}^{\varepsilon}),

which can be explored. In particular for singular kernels we have representations of the Fourier transforms, see for instance [Ste70]. Consequently, we can use this approach to obtain a wide range of interesting examples used in biology or physics.

In the rest of the section we provide some fascinating examples for which the case of convolution structure in Theorem 3.3 can be obtained.

5.1. Uniform bounded confidence model

Let V(x)=i𝟙[R2,R2](x)V(x)=i\mathbbm{1}_{\big{[}-\frac{R}{2},\frac{R}{2}\big{]}}(x) be a complex-valued function. Then

{VV:x{0ifx>|R|,xRifRx0,xRif 0<xR.\displaystyle\begin{cases}V*V\colon&\mathbb{R}\to\mathbb{R}\\ &x\to\begin{cases}0\quad&\mathrm{if}\;x>|R|,\\ -x-R\quad&\mathrm{if}\;-R\leq x\leq 0,\\ x-R\quad&\mathrm{if}\;0<x\leq R.\end{cases}\end{cases}

is a Lipschitz-continuous function with bounded support. Furthermore, we have

(VV)=𝟙[R,0]+𝟙[0,R]=:kUa.e.\nabla(V*V)=-\mathbbm{1}_{[-R,0]}+\mathbbm{1}_{[0,R]}=:k_{\scriptscriptstyle{U}}\quad\text{a.e.}

Consequently, the uniform bounded confidence model, satisfies the assumption of Section 3 with the usual mollification approximation. Also, it is well known that the indicator function 𝟙[R2,R2]Hs()\mathbbm{1}_{\big{[}-\frac{R}{2},\frac{R}{2}\big{]}}\in H^{s}(\mathbb{R}) for all s<1/2s<1/2. We also have the convergence in probability by [CNP23, Lemma 4.7, Theorem 6.1.]. Hence, we obtain the following proposition

Proposition 5.1.

Let kUk_{\scriptscriptstyle{U}} be given above, then the first marginal (ρtN,1,t0)(\rho_{t}^{N,1},t\geq 0) of the law of the system 𝐗N\mathbf{X}^{N} converges to the law (ρt,t0)(\rho_{t},t\geq 0) of 𝐘N\mathbf{Y}^{N} in the L1([0,T];L1())L^{1}([0,T];L^{1}(\mathbb{R}))-norm.

5.2. Parabolic-Elliptic Keller–Segel System

In this subsection we provide an approximation for the elliptic-parabolic Keller–Segel model [KS70] in d\mathbb{R}^{d}. The underlying PDE is given by

{tρt=σ22Δρt(χρtct)Δct=ρt\displaystyle\begin{cases}\partial_{t}\rho_{t}&=\frac{\sigma^{2}}{2}\Delta\rho_{t}-\nabla\cdot(\chi\rho_{t}\nabla c_{t})\\ -\Delta c_{t}&=\rho_{t}\end{cases}

for χ,σ>0\chi,\sigma>0. Decoupling the above system by setting ct=Φρtc_{t}=\Phi*\rho_{t} with Φ\Phi being the fundamental solution of the Laplace equation we can formally derive the above equation from the particle system (2.3) with the interaction force kernel k=Φk=-\nabla\Phi. In particular, if d2d\geq 2 we have

Φ(x)={12πlog(|x|),x0,ifd=2,1d(d2)λ(B1(0))1|x|d2,x0,ifd3,\displaystyle\Phi(x)=\begin{cases}-\frac{1}{2\pi}\log(|x|),&\quad x\neq 0,\quad\mathrm{if}\;d=2,\\ \frac{1}{d(d-2)\lambda(B_{1}(0))}\frac{1}{|x|^{d-2}},&\quad x\neq 0,\quad\mathrm{if}\;d\geq 3,\end{cases}

is the fundamental solution of the Laplace equation.

In the following we present two approaches to mollify our kernel. For the first approach, let us define a mollification kernel JKSJ_{KS}, which satisfies JKS0J_{KS}\geq 0, JKSL1(d)=1\left\lVert J_{KS}\right\rVert_{L^{1}(\mathbb{R}^{d})}=1 and supp(JKS)B(0,1/2)\mathrm{supp}(J_{KS})\subset B(0,1/2) and is infinitely differentiable. As always we set JKSε(x)=1εdJKS(xε)J_{KS}^{\varepsilon}(x)=\frac{1}{\varepsilon^{d}}J_{KS}\big{(}\frac{x}{\varepsilon}\big{)}. Then kε=(JεKSΦJεKS)k^{\varepsilon}=-\nabla(J^{\varepsilon}_{KS}*\Phi*J^{\varepsilon}_{KS}) satisfies all properties of [HLL19a, Theorem 2.1]. Hence, the convergence in probability Assumption 2.5, the law of large numbers Assumption 2.6 is satisfied.

Hence, under consideration of Remark 2.10 we can obtain a relative entropy convergence results on the approximated dd-dimensional attractive Keller–Segel system on the whole space d\mathbb{R}^{d}. We formulate the following proposition as combination of Lemma 3.1 and Theorem 3.3.

Proposition 5.2.

Let kε=(WεVε)xk^{\varepsilon}=-\nabla(W^{\varepsilon}*V^{\varepsilon})_{x} with Wε=JεKSΦW^{\varepsilon}=J^{\varepsilon}_{KS}*\Phi, Vε=JεKSV^{\varepsilon}=J^{\varepsilon}_{KS}. Let ρN,ε\rho^{N,\varepsilon} be the solution of the Lioville equation (2.8) and ρε\rho^{\varepsilon} be the solution to regularized Keller–Segel equation, i.e. to the PDE (2.9). Then, there exists a β>0\beta>0 depending on the dimension dd such that for ε=ε(N)=Nβ\varepsilon=\varepsilon(N)=N^{-\beta} there exists a λ>0\lambda>0 such that

supt[0,T]N(ρN,ε(N)t|ρN,ε(N)t)CNλ.\sup\limits_{t\in[0,T]}{\mathcal{H}}_{N}(\rho^{N,\varepsilon(N)}_{t}\;|\;\rho^{\otimes N,\varepsilon(N)}_{t})\leq CN^{-\lambda}.
Remark 5.3.

By going through the proof of Theorem 3.3 one can obtain a convergence rate and precise condition for β\beta. Furthermore, by inequality (3.21) we have proven convergence of the L1(d)L^{1}(\mathbb{R}^{d})-norm of the marginals

limNρtN,2,ε(N)ρtε(N)ρtε(N)L1(d)=0.\lim\limits_{N\to\infty}\left\lVert\rho_{t}^{N,2,\varepsilon(N)}-\rho_{t}^{\varepsilon(N)}\otimes\rho_{t}^{\varepsilon(N)}\right\rVert_{L^{1}(\mathbb{R}^{d})}=0.

It is also well-known that under additional assumptions on the initial data ρ0\rho_{0} and in the sub-critical regime χ<8π\chi<8\pi in the case d=2d=2 the density ρtε(N)\rho_{t}^{\varepsilon(N)} converges in L1(d)L^{1}(\mathbb{R}^{d}). Hence, we have shown that in the sub-critical case the density of the two marginal ρtN,2,ε(N)\rho_{t}^{N,2,\varepsilon(N)} converges in L1(d)L^{1}(\mathbb{R}^{d}) to the solution of the Keller–Segel equation.

In the case d3d\geq 3 we can obtain an even better approximation, which has a symmetric convolution structure given by kε=(VεVε)k^{\varepsilon}=-\nabla(V^{\varepsilon}*V^{\varepsilon}). Indeed, define the approximation of Φ\Phi as Φε:=JεKSΦJεKS\Phi^{\varepsilon}:=J^{\varepsilon}_{KS}*\Phi*J^{\varepsilon}_{KS}. Then, for cα=πα/2Γ(α/2)c_{\alpha}=\pi^{-\alpha/2}\Gamma(\alpha/2) and

(5.1) Vε=c2cd2d(d2)λ(B1(0))1(|ξ|1(JKSε)(ξ))V^{\varepsilon}=\sqrt{\frac{c_{2}}{c_{d-2}d(d-2)\lambda(B_{1}(0))}}\mathcal{F}^{-1}(|\xi|^{-1}\mathcal{F}(J_{KS}^{\varepsilon})(\xi))

we have

(5.2) Φε=VεVε.\Phi^{\varepsilon}=V^{\varepsilon}*V^{\varepsilon}.

More precisely, for fix ε>0\varepsilon>0 we have JKSεLp(d)J_{KS}^{\varepsilon}\in L^{p}(\mathbb{R}^{d}) for all p1p\geq 1. Hence the Fourier transform (JKSε)\mathcal{F}(J_{KS}^{\varepsilon}) is well-defined and by the Hardy–Littlewood–Sobolev inequality [Ste70, Chapter 5, Theorem 1][LL01, Corollary 5.10] ΦεL2()\Phi^{\varepsilon}\in L^{2}(\mathbb{R}) and the Fourier transform exists. Similar ||1(JKSε)()L2(d)|\cdot|^{-1}\mathcal{F}(J_{KS}^{\varepsilon})(\cdot)\in L^{2}(\mathbb{R}^{d}). A simple calculation shows

||112(JKSε)()L2()2\displaystyle\left\lVert|\cdot|^{-1}\mathcal{F}^{\frac{1}{2}}(J_{KS}^{\varepsilon})(\cdot)\right\rVert_{L^{2}(\mathbb{R})}^{2} =d|ξ|2|(JKSε)(ξ)|dξ\displaystyle=\int_{\mathbb{R}^{d}}|\xi|^{-2}|\mathcal{F}(J_{KS}^{\varepsilon})(\xi)|\,\mathrm{d}\xi
(JKSε)L()B1(0)|ξ|2dξ+B1(0)c|(JKSε)(ξ)|dξ<\displaystyle\leq\left\lVert\mathcal{F}(J_{KS}^{\varepsilon})\right\rVert_{L^{\infty}(\mathbb{R})}\int_{B_{1}(0)}|\xi|^{-2}\,\mathrm{d}\xi+\int_{B_{1}(0)^{\mathrm{c}}}|\mathcal{F}(J_{KS}^{\varepsilon})(\xi)|\,\mathrm{d}\xi<\infty

since d>2d>2 and (JKSε)\mathcal{F}(J_{KS}^{\varepsilon}) is a Schwartz function. As a result, to verify (5.2) we need to show

(Φε)=(Vε)2,\mathcal{F}(\Phi^{\varepsilon})=\mathcal{F}(V^{\varepsilon})^{2},

where the right-hand is square integrable. Now, by [LL01, Corollary 5.10] we have

(Vε)2(ξ)=c2cd2d(d2)λ(B1(0))|ξ|2(JKSε)(ξ)=(Φε)(ξ),\mathcal{F}(V^{\varepsilon})^{2}(\xi)=\frac{c_{2}}{c_{d-2}d(d-2)\lambda(B_{1}(0))}|\xi|^{-2}\mathcal{F}(J_{KS}^{\varepsilon})(\xi)=\mathcal{F}(\Phi^{\varepsilon})(\xi),

where the left-hand side is in L2(d)L^{2}(\mathbb{R}^{d}) by similar arguments as before. Therefore, (5.2) is proven and we can find an appropriate approximation for the Keller–Segel interaction kernel. In particular, we can derive similar estimates to (3.23) with the help of Fourier analysis and the Hardy–Littlewood–Sobolev inequality [Ste70, Chapter  5. Theorem 1]. Clearly, this estimates will now depend on the dimension dd and therefore the convergence rate parameters also depend on the dimension dd.

Proposition 5.4.

Let d3d\geq 3 and kε=(WεVε)xk^{\varepsilon}=-\nabla(W^{\varepsilon}*V^{\varepsilon})_{x} with Wε,VεW^{\varepsilon},V^{\varepsilon} defined by the same expression (5.1). Then the conclusion of Proposition 5.2 holds. Additionally we have the modulated energy estimate

supt[0,T]|𝒦N(ρtN,ε|ρtN,ε)|CNλ\sup\limits_{t\in[0,T]}|{\mathcal{K}}_{N}(\rho_{t}^{N,\varepsilon}|\rho_{t}^{\otimes N,\varepsilon})|\leq CN^{-\lambda}

for some C>0,λ>0C>0,\lambda>0.

Another approach to approximate the Coulomb kernel kk is by utilizing the following approximation in dimension d3d\geq 3,

(5.3) Φε(x)=1d(d2)λ(B1(0))hε(y)|xy|d2dy,\Phi^{\varepsilon}(x)=\frac{1}{d(d-2)\lambda(B_{1}(0))}\int_{\mathbb{R}}\frac{h^{\varepsilon}(y)}{|x-y|^{d-2}}\,\mathrm{d}y,

where

(5.4) hε(y)=1(4πε)d/2exp(|y|24ε)h^{\varepsilon}(y)=\frac{1}{(4\pi\varepsilon)^{d/2}}\exp\bigg{(}-\frac{|y|^{2}}{4\varepsilon}\bigg{)}

is the Weierstrass kernel. Indeed, we note first that the square root 12(hε)\mathcal{F}^{\frac{1}{2}}(h^{\varepsilon}) is well-defined since the Fourier transform of a Gaussian is still a Gaussian or in other words the normal distribution is infinitely divisible. More precisly, by [LL01, Theorem 5.2] we have

(hε)(ξ)=exp(4επ2|ξ|2).\mathcal{F}(h^{\varepsilon})(\xi)=\exp\bigg{(}-4\varepsilon\pi^{2}|\xi|^{2}\bigg{)}.

Hence, similar to the first approximation, we obtain Proposition 5.4. By using the Weierstrass kernel over an abstract mollification kernel we obtain explicit sharp convergence rates. For instance, using Plancherel theorem we obtain

VεL2(d)2\displaystyle\left\lVert\nabla V^{\varepsilon}\right\rVert_{L^{2}(\mathbb{R}^{d})}^{2} =c2cd2d(d2)λ(B1(0))1(|ξ|112(hε)(ξ))()L2(d)2\displaystyle=\frac{c_{2}}{c_{d-2}d(d-2)\lambda(B_{1}(0))}\left\lVert\nabla\mathcal{F}^{-1}(|\xi|^{-1}\mathcal{F}^{\frac{1}{2}}(h^{\varepsilon})(\xi))(\cdot)\right\rVert_{L^{2}(\mathbb{R}^{d})}^{2}
=c2cd2d(d2)λ(B1(0))di=1d|xi1(|ξ|112(hε)(ξ))(x)|2dx\displaystyle=\frac{c_{2}}{c_{d-2}d(d-2)\lambda(B_{1}(0))}\int_{\mathbb{R}^{d}}\sum\limits_{i=1}^{d}|\partial_{x_{i}}\mathcal{F}^{-1}(|\xi|^{-1}\mathcal{F}^{\frac{1}{2}}(h^{\varepsilon})(\xi))(x)|^{2}\,\mathrm{d}x
=c2cd2d(d2)λ(B1(0))di=1d|1(2πiξi|ξ|112(hε)(ξ))(x)|2dx\displaystyle=\frac{c_{2}}{c_{d-2}d(d-2)\lambda(B_{1}(0))}\int_{\mathbb{R}^{d}}\sum\limits_{i=1}^{d}|\mathcal{F}^{-1}(2\pi i\xi_{i}|\xi|^{-1}\mathcal{F}^{\frac{1}{2}}(h^{\varepsilon})(\xi))(x)|^{2}\,\mathrm{d}x
=2πc2cd2d(d2)λ(B1(0))i=1dd|ξiξ112(hε)(ξ)|2dξ\displaystyle=\frac{2\pi c_{2}}{c_{d-2}d(d-2)\lambda(B_{1}(0))}\sum\limits_{i=1}^{d}\int_{\mathbb{R}^{d}}|\xi_{i}\xi^{-1}\mathcal{F}^{\frac{1}{2}}(h^{\varepsilon})(\xi)|^{2}\,\mathrm{d}\xi
=2πc2cd2d(d2)λ(B1(0))i=1dd|ξiξ|1exp(2επ2|ξ|2)|2dξ\displaystyle=\frac{2\pi c_{2}}{c_{d-2}d(d-2)\lambda(B_{1}(0))}\sum\limits_{i=1}^{d}\int_{\mathbb{R}^{d}}|\xi_{i}\xi|^{-1}\exp\bigg{(}-2\varepsilon\pi^{2}|\xi|^{2}\bigg{)}|^{2}\,\mathrm{d}\xi
=2πc2cd2d(d2)λ(B1(0))εd/2i=1dd|ξiξ1exp(2π2|ξ|2)|2dξ\displaystyle=\frac{2\pi c_{2}}{c_{d-2}d(d-2)\lambda(B_{1}(0))\varepsilon^{d/2}}\sum\limits_{i=1}^{d}\int_{\mathbb{R}^{d}}\bigg{|}\xi_{i}\xi^{-1}\exp\bigg{(}-2\pi^{2}|\xi|^{2}\bigg{)}\bigg{|}^{2}\,\mathrm{d}\xi
2πc2cd2(d2)λ(B1(0))εd/2dexp(4π2|ξ|2)|dξ\displaystyle\leq\frac{2\pi c_{2}}{c_{d-2}(d-2)\lambda(B_{1}(0))\varepsilon^{d/2}}\int_{\mathbb{R}^{d}}\exp\bigg{(}-4\pi^{2}|\xi|^{2}\bigg{)}|\,\mathrm{d}\xi
=c22d1πd/21cd2(d2)λ(B1(0))εd/2.\displaystyle=\frac{c_{2}}{2^{d-1}\pi^{d/2-1}c_{d-2}(d-2)\lambda(B_{1}(0))}\varepsilon^{-d/2}.
Remark 5.5.

The above potential is attractive and therefore as far as we know regularization/approximation is necessary to obtain a solution of the underlying Lioville equation on d\mathbb{R}^{d}. Nevertheless, one can obtain tightness of the empirical measure in the super subcritical regime [FJ17]. Our approach provides propagation of chaos of the intermediate system on the level of the relative entropy. Hence, it can be used as a tool to develop further results on the propagation of chaos for the Keller–Segel model without regularization.

5.3. Parabolic-Elliptic Keller–Segel System with Bessel potential

Let us recall the parabolic-elliptic Keller–Segel model [KS70] in d\mathbb{R}^{d} given by

{tρt=σ22Δρt(ρtct)ct=Δct+ρt.\displaystyle\begin{cases}\partial_{t}\rho_{t}&=\frac{\sigma^{2}}{2}\Delta\rho_{t}-\nabla\cdot(\rho_{t}\nabla c_{t})\\ c_{t}&=\Delta c_{t}+\rho_{t}.\end{cases}

Again solving the second equation by setting

ct=(IΔ)1ρt=Gρtc_{t}=(I-\Delta)^{-1}\rho_{t}=G*\rho_{t}

with the L1L^{1} function GG defined by

(5.5) G(x):=1[(1+4π2|ξ|2)1](x)=1(4π)d/20exp(t|x|24t)tn2dt,G(x):=\mathcal{F}^{-1}[(1+4\pi^{2}|\xi|^{2})^{-1}](x)=\frac{1}{(4\pi)^{d/2}}\int\limits_{0}^{\infty}\exp\bigg{(}-t-\frac{|x|^{2}}{4t}\bigg{)}t^{-\frac{n}{2}}\,\mathrm{d}t,

we can decouple the system and obtain an analogous result by using the following approximations of GG,

(5.6) Gε(x)=Ghε(x),G^{\varepsilon}(x)=G*h^{\varepsilon}(x),

or

(5.7) Gε(x)=GJε(x),G^{\varepsilon}(x)=G*J^{\varepsilon}(x),

where hεh^{\varepsilon} is the Weierstrass kernel given by (5.4). Setting

Vε(x)=1[(1+4π2|ξ|2)1/212[hε](ξ)](x),V^{\varepsilon}(x)=\mathcal{F}^{-1}[(1+4\pi^{2}|\xi|^{2})^{-1/2}\mathcal{F}^{\frac{1}{2}}[h^{\varepsilon}](\xi)](x),

it can be shown similar to the elliptic-parabolic Keller–Segel model that

Gε=VεVε.G^{\varepsilon}=V^{\varepsilon}*V^{\varepsilon}.

Consequently, we obtain the analogous result.

Proposition 5.6.

Let kε=Gεk^{\varepsilon}=-\nabla G^{\varepsilon} with GεG^{\varepsilon} defined by (5.5) (5.6) or (5.7) and suppose the Assumptions 2.5 and 2.6 hold. Moreover, for this kεk^{\varepsilon} let ρN,ε\rho^{N,\varepsilon} be the solution of the Lioville equation (2.8) and ρε\rho^{\varepsilon} be the solution to regularized Keller–Segel equation, i.e. to the PDE (2.9). Then, there exists a β>0\beta>0 depending on the dimension dd such that for ε=ε(N)=Nβ\varepsilon=\varepsilon(N)=N^{-\beta} there exists a λ>0\lambda>0 such that

supt[0,T]N(ρN,ε(N)t|ρN,ε(N)t)+supt[0,T]|𝒦N(ρtN,ε|ρtN,ε)|CNλ.\sup\limits_{t\in[0,T]}{\mathcal{H}}_{N}(\rho^{N,\varepsilon(N)}_{t}\;|\;\rho^{\otimes N,\varepsilon(N)}_{t})+\sup\limits_{t\in[0,T]}|{\mathcal{K}}_{N}(\rho_{t}^{N,\varepsilon}|\rho_{t}^{\otimes N,\varepsilon})|\leq CN^{-\lambda}.
Remark 5.7.

By going through the proof of Theorem 3.3 one can obtain a convergence rate and precise condition for β\beta. We also assumed the convergence probability, since we can not reference a concrete result. Nevertheless, we think that this assumption should be true for good enough initial data.

References

  • [AF03] Robert A. Adams and John J. F. Fournier, Sobolev spaces, second ed., Pure and Applied Mathematics (Amsterdam), vol. 140, Elsevier/Academic Press, Amsterdam, 2003.
  • [BJS22] Didier Bresch, Pierre-Emmanuel Jabin, and Juan Soler, A new approach to the mean-field limit of Vlasov–Fokker–Planck equations, ArXiv preprint arXiv:2203.15747 (2022).
  • [BJW19] Didier Bresch, Pierre-Emmanuel Jabin, and Zhenfu Wang, On mean-field limits and quantitative estimates with a large class of singular kernels: application to the Patlak-Keller-Segel model, C. R. Math. Acad. Sci. Paris 357 (2019), no. 9, 708–720.
  • [BJW20] by same author, Mean-field limit and quantitative estimates with singular attractive kernels, ArXiv preprint arXiv:2011.08022 (2020).
  • [BR20] Viorel Barbu and Michael Röckner, From nonlinear Fokker-Planck equations to solutions of distribution dependent SDE, Ann. Probab. 48 (2020), no. 4, 1902–1920.
  • [Bre11] Haim Brezis, Functional analysis, Sobolev spaces and partial differential equations, Universitext, Springer, New York, 2011.
  • [CCH14] José Antonio Carrillo, Young-Pil Choi, and Maxime Hauray, The derivation of swarming models: mean-field limit and Wasserstein distances, Collective dynamics from bacteria to crowds, CISM Courses and Lect., vol. 553, Springer, Vienna, 2014, pp. 1–46.
  • [CCS19] José A. Carrillo, Young-Pil Choi, and Samir Salem, Propagation of chaos for the Vlasov-Poisson-Fokker-Planck equation with a polynomial cut-off, Commun. Contemp. Math. 21 (2019), no. 4, 1850039, 28.
  • [CH23] Alexandra Holzinger Chen, Li and Xiaokai Huo, Quantitative convergence in relative entropy for a moderately interacting particle system on d\mathbb{R}^{d}, ArXiv preprint arXiv:2311.01980 (2023).
  • [CLPY20] Li Chen, Xin Li, Peter Pickl, and Qitao Yin, Combined mean field limit and non-relativistic limit of Vlasov-Maxwell particle system to Vlasov-Poisson system, J. Math. Phys. 61 (2020), no. 6, 061903, 21.
  • [CNP23] Li Chen, Paul Nikolaev, and David J. Prömel, Well-posedness of diffusion-aggregation equations with bounded kernels and their mean-field approximations, ArXiv preprint arXiv:2310.13463 (2023).
  • [DMM01] P. Del Moral and L. Miclo, Genealogies and increasing propagation of chaos for Feynman-Kac and genetic models, Ann. Appl. Probab. 11 (2001), no. 4, 1166–1198.
  • [DTGC14] Pia Domschke, Dumitru Trucu, Alf Gerisch, and Mark A. J. Chaplain, Mathematical modelling of cancer invasion: implications of cell adhesion variability for tumour infiltrative growth patterns, J. Theoret. Biol. 361 (2014), 41–60.
  • [Eva15] Lawrence C. Evans, Partial differential equations, second edition, reprinted with corrections ed., Graduate studies in mathematics; volume 19, Providence, Rhode Island, 2015.
  • [FHS19] Razvan C. Fetecau, Hui Huang, and Weiran Sun, Propagation of chaos for the Keller-Segel equation over bounded domains, J. Differential Equations 266 (2019), no. 4, 2142–2174.
  • [FJ17] Nicolas Fournier and Benjamin Jourdain, Stochastic particle approximation of the Keller-Segel equation and two-dimensional generalization of Bessel processes, Ann. Appl. Probab. 27 (2017), no. 5, 2807–2861.
  • [FW23] Xuanrui Feng and Zhenfu Wang, Quantitative Propagation of Chaos for 2D Viscous Vortex Model on the Whole Space, ArXiv preprint arXiv:2310.05156 (2023).
  • [Gär88] Jürgen Gärtner, On the McKean-Vlasov limit for interacting diffusions, Math. Nachr. 137 (1988), 197–248.
  • [GQ15] David Godinho and Cristobal Quiñinao, Propagation of chaos for a subcritical Keller–Segel model, Annales de l’Institut Henri Poincaré, Probabilités et Statistiques 51 (2015), no. 3, 965–992.
  • [Han22] Yi Han, Entropic propagation of chaos for mean field diffusion with lpl^{p} interactions via hierarchy, linear growth and fractional noise, ArXiv preprint arXiv:2205.02772 (2022).
  • [HKPZ19] Seung-Yeal Ha, Jeongho Kim, Peter Pickl, and Xiongtao Zhang, A probabilistic approach for the mean-field limit to the Cucker-Smale model with a singular communication, Kinet. Relat. Models 12 (2019), no. 5, 1045–1067.
  • [HL09] Seung-Yeal Ha and Jian-Guo Liu, A simple proof of the Cucker-Smale flocking dynamics and mean-field limit, Commun. Math. Sci. 7 (2009), no. 2, 297–325.
  • [HLL19a] Hui Huang, Jian-Guo Liu, and Jianfeng Lu, Learning interacting particle systems: diffusion parameter estimation for aggregation equations, Math. Models Methods Appl. Sci. 29 (2019), no. 1, 1–29.
  • [HLL19b] by same author, Learning interacting particle systems: diffusion parameter estimation for aggregation equations, Math. Models Methods Appl. Sci. 29 (2019), no. 1, 1–29.
  • [HLP20] Hui Huang, Jian-Guo Liu, and Peter Pickl, On the mean-field limit for the Vlasov-Poisson-Fokker-Planck system, J. Stat. Phys. 181 (2020), no. 5, 1915–1965.
  • [HM14] Maxime Hauray and Stéphane Mischler, On Kac’s chaos and related problems, J. Funct. Anal. 266 (2014), no. 10, 6055–6157.
  • [Hol23] Alexandra Holzinger, Rigorous derivations of diffusion systems from moderately interacting particle models, Dissertation, Technische Universität Wien]. reposiTUm. https://doi.org/10.34726/hss.2023.112186 (2023).
  • [Hor04] Dirk Horstmann, From 1970 until present: the Keller-Segel model in chemotaxis and its consequences. II, Jahresber. Deutsch. Math.-Verein. 106 (2004), no. 2, 51–69.
  • [Hos20] Noorazar Hossein, Recent advances in opinion propagation dynamics: a 2020 survey, The European Physical Journal Plus 135 (2020), no. 6, 1–20.
  • [HP09] T. Hillen and K. J. Painter, A user’s guide to PDE models for chemotaxis, J. Math. Biol. 58 (2009), no. 1-2, 183–217.
  • [HRZ22] Zimo Hao, Michael Röckner, and Xicheng Zhang, Strong convergence of propagation of chaos for McKean–Vlasov SDEs with singular interactions, ArXiv preprint arXiv:2204.07952 (2022).
  • [Jab14] Pierre-Emmanuel Jabin, A review of the mean field limits for Vlasov equations, Kinet. Relat. Models 7 (2014), no. 4, 661–711.
  • [JW16] Pierre-Emmanuel Jabin and Zhenfu Wang, Mean field limit and propagation of chaos for Vlasov systems with bounded forces, J. Funct. Anal. 271 (2016), no. 12, 3588–3627.
  • [JW18] by same author, Quantitative estimates of propagation of chaos for stochastic systems with W1,W^{-1,\infty} kernels, Invent. Math. 214 (2018), no. 1, 523–591.
  • [KS70] Evelyn F. Keller and Lee A. Segel, Initiation of slime mold aggregation viewed as an instability, J. Theoret. Biol. 26 (1970), no. 3, 399–415.
  • [Lac23] Daniel Lacker, Hierarchies, entropy, and quantitative propagation of chaos for mean field diffusions, Probab. Math. Phys. 4 (2023), no. 2, 377–432.
  • [LL01] Elliott H. Lieb and Michael Loss, Analysis, second ed., Graduate Studies in Mathematics, vol. 14, American Mathematical Society, Providence, RI, 2001.
  • [LLY19] Lei Li, Jian-Guo Liu, and Pu Yu, On the mean field limit for Brownian particles with Coulomb interaction in 3D, J. Math. Phys. 60 (2019), no. 11, 111501, 34.
  • [Lor07] Jan Lorenz, Continuous Opinion Dynamics Under Bounded Confidence: A Survey, International Journal of Modern Physics C 18 (2007), no. 12, 1819–1838.
  • [LP17] Dustin Lazarovici and Peter Pickl, A mean field limit for the Vlasov–Poisson system, Arch. Ration. Mech. Anal. 225 (2017), no. 3, 1201–1231.
  • [LY19] Jian-Guo Liu and Rong Yang, Propagation of chaos for the Keller-Segel equation with a logarithmic cut-off, Methods Appl. Anal. 26 (2019), no. 4, 319–347.
  • [McK67] H. P. McKean, Jr., Propagation of chaos for a class of non-linear parabolic equations, Stochastic Differential Equations (Lecture Series in Differential Equations, Session 7, Catholic Univ., 1967), Air Force Office Sci. Res., Arlington, Va., 1967, pp. 41–57.
  • [NRS22] Quoc-Hung Nguyen, Matthew Rosenzweig, and Sylvia Serfaty, Mean-field limits of Riesz-type singular flows, Ars Inven. Anal. (2022), Paper No. 4, 45.
  • [Oel84] Karl Oelschläger, A martingale approach to the law of large numbers for weakly interacting stochastic processes, Ann. Probab. 12 (1984), no. 2, 458–479.
  • [Oel87] Karl Oelschläger, A fluctuation theorem for moderately interacting diffusion processes, Probab. Theory Related Fields 74 (1987), no. 4, 591–616.
  • [ORT20] Christian Olivera, Alexandre Richard, and Milica Tomasevic, Quantitative approximation of nonlinear fokker-planck equations with singular potential by interacting particle systems.
  • [Osa87] Hirofumi Osada, Propagation of chaos for the two-dimensional Navier-Stokes equation, Probabilistic methods in mathematical physics (Katata/Kyoto, 1985), Academic Press, Boston, MA, 1987, pp. 303–334.
  • [Rom18] Marco Romito, A simple method for the existence of a density for stochastic evolutions with rough coefficients, Electron. J. Probab. 23 (2018), Paper no. 113, 43.
  • [RS23] Matthew Rosenzweig and Sylvia Serfaty, Global-in-time mean-field convergence for singular Riesz-type diffusive flows, Ann. Appl. Probab. 33 (2023), no. 2, 754–798.
  • [Ser20] Sylvia Serfaty, Mean field limit for Coulomb-type flows, Duke Math. J. 169 (2020), no. 15, 2887–2935.
  • [Ste70] Elias M. Stein, Singular integrals and differentiability properties of functions, Princeton Mathematical Series, No. 30, Princeton University Press, Princeton, N.J., 1970.
  • [Szn91] Alain-Sol Sznitman, Topics in propagation of chaos, École d’Été de Probabilités de Saint-Flour XIX—1989, Lecture Notes in Math., vol. 1464, Springer, Berlin, 1991, pp. 165–251.
  • [TBL06] Chad M. Topaz, Andrea L. Bertozzi, and Mark A. Lewis, A nonlocal continuum model for biological aggregation, Bull. Math. Biol. 68 (2006), no. 7, 1601–1623.
  • [Tri83] Hans Triebel, Theory of function spaces, Monographs in Mathematics, vol. 78, Birkhäuser Verlag, Basel, 1983.
  • [Vil09] Cédric Villani, Optimal transport, Grundlehren der mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 338, Springer-Verlag, Berlin, 2009, Old and new.
  • [WYW06] Zhuoqun Wu, Jingxue Yin, and Chunpeng Wang, Elliptic & parabolic equations, World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ, 2006.
  • [Yos80] Kôsaku Yosida, Functional analysis, sixth ed., Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 123, Springer-Verlag, Berlin-New York, 1980.