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

Existence of martingale solutions for stochastic flocking models with local alignment

Arnaud Debussche, Angelo Rosello
Univ Rennes, CNRS, IRMAR - UMR 6625, F- 35000 Rennes, France
Abstract

We establish the existence of martingale solutions to a class of stochastic conservation equations. The underlying models correspond to random perturbations of kinetic models for collective motion such as the Cucker-Smale [6, 7] and Motsch-Tadmor [16] models. By regularizing the coefficients, we first construct approximate solutions obtained as the mean-field limit of the corresponding particle systems. We then establish the compactness in law of this family of solutions by relying on a stochastic averaging lemma. This extends the results obtained in [12, 11] in the deterministic case.   
  
Keywords: Stochastic partial differential equations, mean-field limit, collective motion.

1 Introduction, main results

1.1 Collective motion with local alignment

The emergence of a consensus or ordered motion amongst a population of interacting agents has been drawing a fair amount of attention among the scientific community in recent years. This phenomenon, consistently observed in nature, from schooling fish to swarming bacteria, is usually referred to as flocking. One of the earliest and most commonly studied mathematical models describing this kind of behavior is the celebrated Cucker-Smale model, introduced in [6, 7]. In this model, agents interact in a mean-field manner: for 1iN1\leq i\leq N, denoting by Xi,N,Vi,NdX^{i,N},V^{i,N}\in\mathbb{R}^{d} the position and velocity of the ii-th individual, the evolution of the system is given by

{ddtXti,N=Vti,N,ddtVti,N=1Nj=1Nψ(Xti,NXtj,N)(Vtj,NVti,N),\left\{\begin{array}[]{l}\displaystyle{\frac{d}{dt}}X^{i,N}_{t}=V^{i,N}_{t},\\ \displaystyle{\frac{d}{dt}}V^{i,N}_{t}=\frac{1}{N}\sum_{j=1}^{N}\psi(X^{i,N}_{t}-X^{j,N}_{t})(V_{t}^{j,N}-V^{i,N}_{t}),\end{array}\right.

where the weight function ψ:d+\psi:\mathbb{R}^{d}\to\mathbb{R}^{+} is even, typically of the form

ψ(xy)=λ(1+|xy|2)γ,λ,γ>0.\displaystyle\psi(x-y)=\frac{\lambda}{(1+|x-y|^{2})^{\gamma}},\hskip 14.22636pt\lambda,\gamma>0.

Equivalently, one may consider the conservation equation

tf+vxf+v(LCS[f]f)=0\displaystyle\partial_{t}f+v\cdot\nabla_{x}f+\nabla_{v}\cdot(L^{CS}[f]f)=0 (1.1)

where the Cucker-Smale alignment term LCS[f]L^{CS}[f] is given by the convolution

LCS[f](x,v)=d×dψ(xy)(wv)f(y,w)𝑑y𝑑w.\displaystyle L^{CS}[f](x,v)=\int_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\psi(x-y)(w-v)f(y,w)dydw. (1.2)

Equation (1.1) is naturally associated to the particle system, since it is satisfied by the empirical measure

μtN=1Ni=1Nδ(Xti,N,Vti,N)\mu^{N}_{t}=\frac{1}{N}\sum_{i=1}^{N}\delta_{(X_{t}^{i,N},V_{t}^{i,N})}

in the sense of distributions. In [16], Motsch and Tadmor brought to light several drawbacks regarding the physical relevance of the Cucker-Smale model when confronted to strongly non-homegeneous distributions of agents, due to the normalizing constant 1N\frac{1}{N} in the alignment force, which involves the whole group of individuals. To remedy these issues, they proposed a new model where the influence between two agents is normalized by the total influence:

{ddtXti,N=Vti,N,ddtVti,N=1j=1Nϕ(Xti,NXtj,N)j=1Nϕ(Xti,NXtj,N)(Vtj,NVti,N).\left\{\begin{array}[]{l}\displaystyle{\frac{d}{dt}}X^{i,N}_{t}=V^{i,N}_{t},\\ \displaystyle{\frac{d}{dt}}V^{i,N}_{t}=\frac{1}{\sum_{j=1}^{N}\phi(X^{i,N}_{t}-X^{j,N}_{t})}\sum_{j=1}^{N}\phi(X^{i,N}_{t}-X^{j,N}_{t})(V_{t}^{j,N}-V^{i,N}_{t}).\end{array}\right.

Considering some weight function ϕ:d+\phi:\mathbb{R}^{d}\to\mathbb{R}^{+} with compact support, we may naturally consider a hybrid model, letting the Cucker-Smale forcing dictate the long-range interaction and the Motsch-Tadmor term dictate the short-range interaction:

tf+vxf+v((LCS[f]+LMT[f])f)=0\displaystyle\partial_{t}f+v\cdot\nabla_{x}f+\nabla_{v}\cdot((L^{CS}[f]+L^{MT}[f])f)=0 (1.3)

where the LMT[f]L^{MT}[f] is given by

LMT[f](x,v)=d×dϕ(xy)(wv)f(y,w)𝑑y𝑑wd×dϕ(xy)f(y,w)𝑑y𝑑w.\displaystyle L^{MT}[f](x,v)=\frac{\displaystyle{\int_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\phi(x-y)(w-v)\,f(y,w)dydw}}{\displaystyle{\int_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\phi(x-y)f(y,w)dydw}}. (1.4)

Note that the Motsch-Tadmor forcing term can also be written

LMT[f](x,v)=u[f](x)v,L^{MT}[f](x,v)=u[f](x)-v,

expressing the alignment of the speed with the local average velocity u[f]u[f], defined as

u[f](x)=d×dϕ(xy)wf(y,w)𝑑y𝑑wd×dϕ(xy)f(y,w)𝑑y𝑑w.\displaystyle u[f](x)=\frac{\displaystyle{\int_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\phi(x-y)w\,f(y,w)dydw}}{\displaystyle{\int_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\phi(x-y)f(y,w)dydw}}.

As suggested in [12], we may also consider (1.3) in the singular limit where the weight function ϕ\phi governing the short-range interaction converges to the Dirac function δ0\delta_{0}, leading to

tf+vxf+v((LCS[f]+LSLA[f])f)=0.\displaystyle\partial_{t}f+v\cdot\nabla_{x}f+\nabla_{v}\cdot((L^{CS}[f]+L^{SLA}[f])f)=0. (1.5)

In (1.5), the Strong Local Alignment term LSLA[f]L^{SLA}[f] is given by

LSLA[f](x,v)=u0[f](x)v,\displaystyle L^{SLA}[f](x,v)=u_{0}[f](x)-v, (1.6)

where the local velocity u0[f]u_{0}[f] is given by

u0[f](x)=dwf(x,w)𝑑wdf(x,w)𝑑w.\displaystyle u_{0}[f](x)=\displaystyle{\frac{\int_{\mathbb{R}^{d}}wf(x,w)dw}{\int_{\mathbb{R}^{d}}f(x,w)dw}}.

The existence of solutions to the kinetic equations (1.3) and (1.5) has been established in [12]. Moreover, in [11], the authors rigorously explore the limit ϕδ0\phi\to\delta_{0}: considering some ϕ1Cc(d)\phi_{1}\in C_{c}(\mathbb{R}^{d}) and weight functions of the form ϕr(x)=rdϕ1(x/r)\phi^{r}(x)=r^{-d}\phi_{1}(x/r), solutions (fr)r0(f^{r})_{r\geq 0} of (1.3) converge (up to some subsequence) to a solution ff of (1.5).

In order to take into account random phenomena emerging from the environment, or unpredictable interactions between the agents, it is rather natural to perturb the deterministic equations (1.3) and (1.5) with some noise, driven by a Wiener process dW(z)=kKk[f](z)dβtkdW(z)=\sum_{k}K_{k}[f](z)d\beta^{k}_{t}, leading to the stochastic conservation equation

dft+[vxft+v((L[ft]ft)]dt+kv(Kk[ft]ft)dβtk=0df_{t}+\Big{[}v\cdot\nabla_{x}f_{t}+\nabla_{v}\cdot((L[f_{t}]f_{t})\Big{]}dt+\sum_{k}\nabla_{v}\cdot(K_{k}[f_{t}]f_{t})\circ d\beta_{t}^{k}=0

with L[f]=LCS[f]+LMT[f]L[f]=L^{CS}[f]+L^{MT}[f] or L[f]=LCS[f]+LSLA[f]L[f]=L^{CS}[f]+L^{SLA}[f]. Here, for simplicity purposes, we choose to only consider a "one-dimensional" noise driven by a real valued Brownian motion β=(βt)t0\beta=(\beta_{t})_{t\geq 0}, leading to equations

dft+[vxft+v((LCS[ft]+LMT[ft])ft)]dt+v(K[ft]ft)dβt=0\displaystyle df_{t}+\Big{[}v\cdot\nabla_{x}f_{t}+\nabla_{v}\cdot((L^{CS}[f_{t}]+L^{MT}[f_{t}])f_{t})\Big{]}dt+\nabla_{v}\cdot(K[f_{t}]f_{t})\circ d\beta_{t}=0 (1.7)

and

dft+[vxft+v((LCS[ft]+LSLA[ft])ft)]dt+v(K[ft]ft)dβt=0.\displaystyle df_{t}+\Big{[}v\cdot\nabla_{x}f_{t}+\nabla_{v}\cdot((L^{CS}[f_{t}]+L^{SLA}[f_{t}])f_{t})\Big{]}dt+\nabla_{v}\cdot(K[f_{t}]f_{t})\circ d\beta_{t}=0. (1.8)

The methods developed in the present paper shall not rely on this particular form, so that the results may be easily generalized to SPDEs with multiple Brownian motions. Note that these stochastic conservation equations are written in Stratonovitch form, since it is the most physically relevant form. As in the deterministic case, equation (1.7) is naturally associated to the stochastic particle system

{dXti,N=Vti,Ndt,dVti,N=L[μtN](Xti,N,Vti,N)dt+K[μtN](Xti,N,Vti,N)dβt\left\{\begin{array}[]{l}dX^{i,N}_{t}=V^{i,N}_{t}dt,\\ dV^{i,N}_{t}=L[\mu^{N}_{t}](X^{i,N}_{t},V^{i,N}_{t})dt+K[\mu^{N}_{t}](X^{i,N}_{t},V^{i,N}_{t})\circ d\beta_{t}\end{array}\right. (1.9)

where μtN=1Ni=1Nδ(Xti,N,Vti,N)\mu^{N}_{t}=\frac{1}{N}\sum_{i=1}^{N}\delta_{(X^{i,N}_{t},V^{i,N}_{t})} and L[μ]=LCS[μ]+LMT[μ]L[\mu]=L^{CS}[\mu]+L^{MT}[\mu].

The mean-field convergence of (1.9) to the corresponding limiting SPDE has been studied in the litterature in the case of the Cucker-Smale interaction, that is with L[μ]=LCS[μ]L[\mu]=L^{CS}[\mu]. The diffusion coefficient K[μ](x,v)=D(vvc)K[\mu](x,v)~{}=~{}D(v-v_{c}) for some constant vcdv_{c}\in\mathbb{R}^{d} is considered in [1] ; the coefficient K[μ](x,v)=2σ(vv¯)K[\mu](x,v)=\sqrt{2\sigma}(v-\bar{v}), where v¯=w𝑑μ(y,w)\bar{v}=\int wd\mu(y,w), is looked upon in [4] and [10] ; some more general (non linear) diffusion coefficients are considered in [18].

As for the Motsch-Tadmor model, that is for L[μ]=LMT[μ]L[\mu]=L^{MT}[\mu], the flocking phenomenon for the particle system (1.9) (alignment of speeds, distance between the individuals bounded over time) is studied in [15] in the case of a multiplicative noise K[μ](x,v)=DvK[\mu](x,v)=Dv. However, even in the deterministic case, due to the singular ratio involved in the non-linear term LMT[μ]L^{MT}[\mu], the mean-field limit of the Motsch-Tadmor particle system is a very delicate question (as suggested in [16] and [17]) to which the authors could not find a proper answer in the litterature. It should also be noted that the strong local alignment term LSLA[f]L^{SLA}[f] given by (1.6) is ill-defined when ff is a general measure, so that the particle system associated with equation (1.8) cannot in fact be written.

In the present work, we shall consider a diffusion coefficient of the form

K[f](x,v)=F(x)+d×dψ~(xy)(wv)f(y,w)𝑑y𝑑w\displaystyle K[f](x,v)=F(x)+\int_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\widetilde{\psi}(x-y)(w-v)f(y,w)dydw (1.10)

which corresponds to some random environmental forcing F(x)dβtF(x)\circ d\beta_{t}, as well as a random perturbation ψψ+ψ~dβt\psi\leftarrow\psi+\widetilde{\psi}\circ d\beta_{t} of the weight function involved in the Cucker-Smale alignment term (1.2) (as considered in [4] and [18] in the case of the Cucker-Smale model only). Note that the choice F=0F=0 and ψ~=2σ\widetilde{\psi}=\sqrt{2\sigma} leads in particular to the simpler coefficient K[μ](x,v)=2σ(vv¯)K[\mu](x,v)=\sqrt{2\sigma}(v-\bar{v}). The arguments developed in the present work also easily apply to the case K[μ](x,v)=DvK[\mu](x,v)=Dv looked upon in [15].

In this paper, we extend the work developed in [12] and [11] in the deterministic case to establish the existence of martingale solutions (see Definition 1.1 below) for the stochastic conservation equations (1.7) and (1.8). To this intent, we start by regularizing the coefficients: in section 2, we prove the existence of a unique solution of equation (1.7) with regularized coefficients, which is naturally constructed as the mean-field limit of the corresponding stochastic particle system. Then, in section 3, we prove the tightness of these approximate solutions with respect to the regularizing parameter, and rigorously pass to the limit in the martingale problem associated with (1.7).

1.2 Assumptions and main results

The weight functions ψ,ψ~:d+\psi,\widetilde{\psi}:\mathbb{R}^{d}\to\mathbb{R}^{+} involved in (1.2) and (1.10) are assumed to satisfy

|ψ(x)|+|ψ~(x)|1,|xαψ(x)|+|xαψ~(x)|1 for 1|α|4.\displaystyle|\psi(x)|+|\widetilde{\psi}(x)|\lesssim 1,\hskip 14.22636pt|\partial^{\alpha}_{x}\psi(x)|+|\partial^{\alpha}_{x}\widetilde{\psi}(x)|\lesssim 1\text{ for $1\leq|\alpha|\leq 4$}. (1.11)

The weight function ϕ:d+\phi:\mathbb{R}^{d}\to\mathbb{R}^{+} involved in (1.4) is assumed to be smooth and compactly supported around zero: ϕC(d)\phi\in C^{\infty}(\mathbb{R}^{d}) and for some 0<r1<r2<0<r_{1}<r_{2}<\infty,

infxB(0,r1)ϕ(x)>0,Supp(ϕ)B(0,r2).\displaystyle\inf_{x\in B(0,r_{1})}\phi(x)>0,\hskip 28.45274ptSupp(\phi)\subset B(0,r_{2}). (1.12)

The forcing FF involved in (1.10) is assumed to be smooth and sublinear:

|F(x)|1+|x|.\displaystyle|F(x)|\lesssim 1+|x|.

Simple calculations show that the proper Itô form associated with SPDE (1.7) is

dft+v((LCS[ft]+LMT[ft]+S[ft])ft)dt+v(K[ft]ft)dβt=12v2(K[ft]K[ft]Tft)dt\displaystyle df_{t}+\nabla_{v}\cdot\Big{(}\Big{(}L^{CS}[f_{t}]+L^{MT}[f_{t}]+S[f_{t}]\Big{)}f_{t}\Big{)}dt+\nabla_{v}\cdot\left(K[f_{t}]f_{t}\right)d\beta_{t}=\frac{1}{2}\nabla_{v}^{2}(K[f_{t}]K[f_{t}]^{T}f_{t})dt

where we have used the notation

v2(K[f]K[f]Tf)=1i,jdvivj2[K[f]iK[f]jf]\nabla_{v}^{2}(K[f]K[f]^{T}f)=\sum_{1\leq i,j\leq d}\partial^{2}_{v_{i}v_{j}}\Big{[}K[f]^{i}K[f]^{j}f\Big{]}

and the additional drift forcing term S[f]S[f] is given by

S[f](z)=122dψ~(xy)(K[f](y,w)K[f](x,v))f(y,w)𝑑y𝑑w.\displaystyle S[f](z)=\frac{1}{2}\int_{\mathbb{R}^{2d}}\widetilde{\psi}(x-y)\Big{(}K[f](y,w)-K[f](x,v)\Big{)}f(y,w)dydw.

This motivates the following definition.

Definition 1.1.

Let T>0T>0 and (Ω,,(t)0tT,,β)(\Omega,{\cal F},({\cal F}_{t})_{0\leq t\leq T},\mathbb{P},\beta) be a filtered probability space equipped with an (t)({\cal F}_{t})-brownian motion β\beta. Let f0:2d+f_{0}:\mathbb{R}^{2d}\to\mathbb{R}_{+} with 2df0(z)𝑑z=1\int_{\mathbb{R}^{2d}}f_{0}(z)dz=1.
A process fL(Ω;L([0,T];L1(2d)))f\in L^{\infty}\Big{(}\Omega;L^{\infty}([0,T];L^{1}(\mathbb{R}^{2d}))\Big{)} with

[2d|f(ω)(t,z)|𝑑z1,dt-a.e],-a.s\displaystyle\Big{[}\int_{\mathbb{R}^{2d}}|f(\omega)(t,z)|dz\leq 1,\;\;dt\text{-a.e}\Big{]},\;\;\mathbb{P}\text{-a.s} (1.13)

satisfying the estimate

𝔼[0T2d(1+|v|2)|f(t,z)|𝑑z𝑑t]<\displaystyle\mathbb{E}\Big{[}\int_{0}^{T}\int_{\mathbb{R}^{2d}}(1+|v|^{2})|f(t,z)|dzdt\Big{]}<\infty (1.14)

is said to be a solution of (1.7) on (Ω,β)(\Omega,\beta) (with initial value f0f_{0}) when, for any test function ΨCc(2d)\Psi\in C^{\infty}_{c}(\mathbb{R}^{2d}), the process (f(t),Ψ)0tT(\langle f(t),\Psi\rangle)_{0\leq t\leq T} is adapted with a continuous version and satisfies

f(t),Ψ=f0,Ψ+0t<[f(s)]Ψ,f(s)>ds+0t<K[f(s)]vΨ,f(s)>dβs,t[0,T]\langle f(t),\Psi\rangle=\langle f_{0},\Psi\rangle+\int_{0}^{t}\bigl{<}{\cal L}[f(s)]\Psi,f(s)\bigr{>}ds+\int_{0}^{t}\bigl{<}K[f(s)]\cdot\nabla_{v}\Psi,f(s)\bigr{>}d\beta_{s},\;\;\;t\in[0,T]

where [f]{\cal L}[f] denotes the second order operator

[f]Ψ=vxΨ+(LCS[f]+LMT[f]+S[f])vΨ+121i,jdK[f]iK[f]jvivj2Ψ.\displaystyle{\cal L}[f]\Psi=v\cdot\nabla_{x}\Psi+\Big{(}L^{CS}[f]+L^{MT}[f]+S[f]\Big{)}\cdot\nabla_{v}\Psi+\frac{1}{2}\sum_{1\leq i,j\leq d}K[f]^{i}K[f]^{j}\partial^{2}_{v_{i}v_{j}}\Psi. (1.15)

If there exists some probability tuple (Ω¯,¯,(¯t)0tT,¯,β¯)(\overline{\Omega},\overline{\cal F},(\overline{\cal F}_{t})_{0\leq t\leq T},\overline{\mathbb{P}},\overline{\beta}) and a solution f¯\overline{f} of (1.7) on (Ω¯,β¯)(\overline{\Omega},\overline{\beta}), we say that equation (1.7){(\ref{chap3-spde})} has a martingale solution (in the sense of [8], Chapter 8).

Estimates (1.13) and (1.14) are quite natural since we expect solutions of (1.7) to be densities. On can easily check that these estimates guarantee that the process (f(t),Ψ)t0(\langle f(t),\Psi\rangle)_{t\geq 0} is well defined, the stochastic integral being a square integrable martingale. Solutions of equation (1.8) are defined similarly. We now state our main results.

Theorem 1 (Stochastic Motsch-Tadmor flocking).

Let f0:2d+f_{0}:\mathbb{R}^{2d}\to\mathbb{R}_{+} with 2df0(z)𝑑z=1\int_{\mathbb{R}^{2d}}f_{0}(z)dz=1 such that, for some δ>1\delta>1 and θ(0,1)\theta\in(0,1),

2d|f0(z)|p𝑑z+2d(|x|δ+|v|k)f0(z)𝑑z<,p=1+1θ,k>max(d+2,4)1θ.\displaystyle\int_{\mathbb{R}^{2d}}|f_{0}(z)|^{p}dz+\int_{\mathbb{R}^{2d}}(|x|^{\delta}+|v|^{k})f_{0}(z)dz<\infty,\hskip 14.22636ptp=1+\frac{1}{\theta},\;\;k>\frac{\max(d+2,4)}{1-\theta}. (1.16)

Then there exists a martingale solution ff of equation (1.7) with initial data f0f_{0}.

We also prove the existence of a martingale solution of (1.8), which can be constructed as a weak limit of solutions of (1.7) as the function ϕ\phi involved in the Motsch-Tadmor alignment term (1.4) properly approaches the Dirac function δ0\delta_{0}, similarly to the result established in [11].

Theorem 2 (Stochastic flocking with strong local alignment).

Let ϕ1Cc(d)\phi_{1}\in C_{c}^{\infty}(\mathbb{R}^{d}) satisfy (1.12). Let us consider the sequence of functions (ϕr)r>0(\phi_{r})_{r>0} given by

xd,ϕr(x)=rdϕ1(x/r).\forall x\in\mathbb{R}^{d},\;\;\;\phi_{r}(x)=r^{d}\phi_{1}(x/r).

Assuming (1.16), the martingale solution frf^{r} of equation (1.7) with ϕ=ϕr\phi=\phi_{r} constructed in Theorem 1 satisfies, along some subsequence rn0r_{n}\to 0,

frnf in law, in C([0,T];HW1σ(2d)) for all σ>0,f^{r_{n}}\to f\text{ in law},\text{ in }C([0,T];H^{-\sigma}_{W^{-1}}(\mathbb{R}^{2d}))\text{ for all $\sigma>0$},

where ff defines a martingale solution of (1.8). The weighted Sobolev space HW1σ(2d)H^{-\sigma}_{W^{-1}}(\mathbb{R}^{2d}), with W1(z)=(1+|z|)1W^{-1}(z)=(1+|z|)^{-1} is introduced in (3.34) below.

Note that, although ff only converges weakly in space, the use of a stochastic averaging lemma (developed in Proposition 3.4 below) will guarantee that integrated quantities of the form

ρφ=dφ(v)f𝑑v\rho_{\varphi}=\int_{\mathbb{R}^{d}}\varphi(v)fdv

converge in the strong space L2([0,T];L2(d))L^{2}([0,T];L^{2}(\mathbb{R}^{d})). This allows to properly pass to the limit in the non-linear equation (1.7).

Acknowledgment

A. Debussche and A. Rosello are partially supported by the French government thanks to the "Investissements d’Avenir" program ANR-11-LABX-0020-01.

2 Regularized equation

In this section, we prove existence and uniqueness of a solution of equation (1.7) with regularized coefficients. This solution will be naturally obtained as the mean-field limit of the corresponding particle system.

For R>0R>0 let us introduce smooth, compactly supported truncation functions

χRCc(d;),θRCc(d;d)\chi_{R}\in C^{\infty}_{c}(\mathbb{R}^{d};\mathbb{R}),\hskip 14.22636pt\theta_{R}\in C^{\infty}_{c}(\mathbb{R}^{d};\mathbb{R}^{d})

satisfying

{χR(x)=1 if |x|R,|χR(x)|1,θR(v)=v if |v|R,|θR(v)||v|,|vθR(v)|1 uniformly in R>0.\left\{\begin{array}[]{l}\chi_{R}(x)=1\;\text{ if }\;|x|\leq R,\hskip 28.45274pt|\chi_{R}(x)|\leq 1,\vspace{1mm}\\ \theta_{R}(v)=v\;\text{ if }\;|v|\leq R,\hskip 28.45274pt|\theta_{R}(v)|\leq|v|,\vspace{1mm}\\ |\nabla_{v}\cdot\theta_{R}(v)|\lesssim 1\;\;\text{ uniformly in }R>0.\end{array}\right. (2.1)

We may then introduce the following regularized coefficients:

{LRCS[μ](z)=χR(xy)ψ(xy)θR(wv)𝑑μ(y,w),uR[μ](x)=d×dϕ(xy)θR(w)𝑑μ(y,w)R1+d×dϕ(xy)𝑑μ(y,w),LRMT[μ](z)=uR[μ](x)v,KR[μ](z)=χR(x)F(x)+d×dχR(xy)ψ~(xy)θR(wv)𝑑μ(y,w),SR[μ](z)=122dψ~(xy)(KR[μ](y,w)KR[μ](x,v))𝑑μ(y,w).\left\{\begin{array}[]{l}L^{CS}_{R}[\mu](z)=\displaystyle{\int\chi_{R}(x-y)\psi(x-y)\theta_{R}(w-v)d\mu(y,w),\vspace{2mm}}\\ u_{R}[\mu](x)=\frac{\displaystyle{\int_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\phi(x-y)\theta_{R}(w)\,d\mu(y,w)}}{\displaystyle{R^{-1}+\int_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\phi(x-y)d\mu(y,w)}},\vspace{2mm}\\ L^{MT}_{R}[\mu](z)=u_{R}[\mu](x)-v,\vspace{2mm}\\ K_{R}[\mu](z)=\displaystyle{\chi_{R}(x)F(x)+\int_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\chi_{R}(x-y)\widetilde{\psi}(x-y)\theta_{R}(w-v)d\mu(y,w),\vspace{2mm}}\\ S_{R}[\mu](z)=\displaystyle{\frac{1}{2}\int_{\mathbb{R}^{2d}}\widetilde{\psi}(x-y)\Big{(}K_{R}[\mu](y,w)-K_{R}[\mu](x,v)\Big{)}d\mu(y,w).}\end{array}\right. (2.2)

Simple calculations show that, for fixed RR, these regularized coefficients are globally Lipschitz continuous in the following sense: for all z,z2dz,z^{\prime}\in\mathbb{R}^{2d}, μ,ν𝒫(2d)\mu,\nu\in{\cal P}(\mathbb{R}^{2d}),

|LRCS[μ](z)LRCS[ν](z)|+|LRMT[μ](z)LRMT[ν](z)||zz|+W1[μ,ν],|KR[μ](z)KR[ν](z)|+|SR[μ](z)SR[ν](z)||zz|+W1[μ,ν]\begin{array}[]{l}\Big{|}L^{CS}_{R}[\mu](z)-L^{CS}_{R}[\nu](z^{\prime})\Big{|}+\Big{|}L^{MT}_{R}[\mu](z)-L^{MT}_{R}[\nu](z^{\prime})\Big{|}\lesssim|z-z^{\prime}|+W_{1}[\mu,\nu],\\ \Big{|}K_{R}[\mu](z)-K_{R}[\nu](z^{\prime})\Big{|}+\Big{|}S_{R}[\mu](z)-S_{R}[\nu](z^{\prime})\Big{|}\lesssim|z-z^{\prime}|+W_{1}[\mu,\nu]\end{array} (2.3)

where the constants involved in \lesssim in (2.3) depend on RR, and W1W_{1} denotes the Wasserstein distance. The assumptions made also guarantee the uniform sub-linearity of some coefficients: for all z2dz\in\mathbb{R}^{2d}, μ𝒫(2d)\mu\in{\cal P}(\mathbb{R}^{2d}),

|LRCS[μ](z)|+|KR[μ](z)|+|SR[μ](z)|1+|z|+|z|𝑑μ(z)\displaystyle\Big{|}L^{CS}_{R}[\mu](z)\Big{|}+\Big{|}K_{R}[\mu](z)\Big{|}+\Big{|}S_{R}[\mu](z)\Big{|}\lesssim 1+|z|+\int|z^{\prime}|d\mu(z^{\prime}) (2.4)

where the constant involved in \lesssim in (2.4) does not depend on RR.

2.1 Mean-field limit of the associated particle system

We may now consider the associated mean-field particle system on Zi,N=(Xi,N,Vi,N)2dZ^{i,N}=(X^{i,N},V^{i,N})\in\mathbb{R}^{2d}, i{1,,N}i\in\{1,\ldots,N\}:

{dXti,N=Vti,NdtdVti,N=(LRCS[μtN]+LRMT[μtN]+SR[μtN])(Zti,N)dt+KR[μtN](Zti,N)dβt,\left\{\begin{array}[]{l }dX^{i,N}_{t}=V^{i,N}_{t}dt\\ dV^{i,N}_{t}=\Big{(}L^{CS}_{R}[\mu^{N}_{t}]+L^{MT}_{R}[\mu^{N}_{t}]+S_{R}[\mu^{N}_{t}]\Big{)}(Z^{i,N}_{t})dt+K_{R}[\mu^{N}_{t}](Z^{i,N}_{t})d\beta_{t},\end{array}\right. (2.5)

where μtN\mu_{t}^{N} denotes the empirical measure

μtN=1Ni=1NδZti,N.\displaystyle\mu_{t}^{N}=\frac{1}{N}\sum_{i=1}^{N}\delta_{Z^{i,N}_{t}}.

From the sub-linearity of the coefficients, we easily deduce the following result.

Proposition 2.1.

For any T>0T>0 and z0i,N2dz_{0}^{i,N}\in\mathbb{R}^{2d}, i{1,,N}i\in\{1,\ldots,N\}, the SDE system (2.5) with initial condition Z0i,N=z0i,NZ_{0}^{i,N}=z_{0}^{i,N} has a unique global solution (Zti,N)t[0,T]i{1,,N}(Z^{i,N}_{t})_{t\in[0,T]}^{i\in\{1,\ldots,N\}} which satisfies the following estimates: for any p1p\geq 1,

𝔼[supt[0,T]2d|z|p𝑑μtN(z)]1+2d|z|p𝑑μ0N(z),\displaystyle\mathbb{E}\Big{[}\sup_{t\in[0,T]}\int_{\mathbb{R}^{2d}}|z|^{p}d\mu^{N}_{t}(z)\Big{]}\lesssim 1+\int_{\mathbb{R}^{2d}}|z|^{p}d\mu_{0}^{N}(z),
𝔼[supt[0,T]|Zti,N|p]1+|z0i,N|p+2d|z|p𝑑μ0N(z).\displaystyle\mathbb{E}\Big{[}\sup_{t\in[0,T]}\Big{|}Z^{i,N}_{t}\Big{|}^{p}\Big{]}\lesssim 1+|z_{0}^{i,N}|^{p}+\int_{\mathbb{R}^{2d}}|z|^{p}d\mu_{0}^{N}(z).

The constants involved in \lesssim depend on R,pR,p and TT only.

Proof.

The coefficients of (2.5) being locally Lipschitz-continuous, the local existence and uniqueness of solutions is guaranteed. The estimates of Proposition 2.1 should first be established with the stopping time

τM=inf{t0,max1iN|Zi,N|M}T,\tau_{M}=\inf\left\{t\geq 0,\;\;\max_{1\leq i\leq N}\Big{|}Z^{i,N}\Big{|}\geq M\right\}\wedge T,

which should then be sent to TT, as MM goes to infinity. For the sake of simplicity, we omit this stopping time in the following. Given the sub-linearity of the coefficients, Itô’s formula gives

|Zti,N|p=|z0i,N|p+0tAsi𝑑s+0tBsi𝑑βs\displaystyle\Big{|}Z^{i,N}_{t}\Big{|}^{p}=|z_{0}^{i,N}|^{p}+\int_{0}^{t}A^{i}_{s}ds+\int_{0}^{t}B^{i}_{s}d\beta_{s} (2.6)

with

Ati1+|Zti,N|p+|z|p𝑑μtN,Bti=p|Zti,N|p2Zti,NKR[μtN](Zti,N).\displaystyle A^{i}_{t}\lesssim 1+|Z^{i,N}_{t}|^{p}+\int|z|^{p}d\mu^{N}_{t},\hskip 28.45274ptB^{i}_{t}=p\Big{|}Z_{t}^{i,N}\Big{|}^{p-2}Z_{t}^{i,N}\cdot K_{R}[\mu_{t}^{N}](Z_{t}^{i,N}).

Averaging (2.6) over ii, we are led to

|z|p𝑑μtN=|z|p𝑑μ0N+0t(1Ni=1NAsi)𝑑s+0t(1Ni=1NBsi)𝑑βs\displaystyle\int|z|^{p}d\mu^{N}_{t}=\int|z|^{p}d\mu_{0}^{N}+\int_{0}^{t}\Big{(}\frac{1}{N}\sum_{i=1}^{N}A^{i}_{s}\Big{)}ds+\int_{0}^{t}\Big{(}\frac{1}{N}\sum_{i=1}^{N}B^{i}_{s}\Big{)}d\beta_{s} (2.7)

from which we easily deduce, for all 0rT0\leq r\leq T

𝔼[supt[0,r]|z|p𝑑μtN]1+𝔼[|z|p𝑑μ0N]\displaystyle\mathbb{E}\Big{[}\sup_{t\in[0,r]}\int|z|^{p}d\mu^{N}_{t}\Big{]}\lesssim 1+\mathbb{E}\Big{[}\int|z|^{p}d\mu_{0}^{N}\Big{]} +0r𝔼[supt[0,s]|z|p𝑑μtN]𝑑s\displaystyle+\int_{0}^{r}\mathbb{E}\Big{[}\sup_{t\in[0,s]}\int|z|^{p}d\mu_{t}^{N}\Big{]}ds
+𝔼[supt[0,r]0t(1Ni=1NBsi)𝑑βs].\displaystyle+\mathbb{E}\Big{[}\sup_{t\in[0,r]}\int_{0}^{t}\Big{(}\frac{1}{N}\sum_{i=1}^{N}B^{i}_{s}\Big{)}d\beta_{s}\Big{]}.

Burkholder-Davis-Gundy’s inequality (from [3]) gives

𝔼[supt[0,r]0t(1Ni=1NBsi)𝑑βs]\displaystyle\mathbb{E}\Big{[}\sup_{t\in[0,r]}\int_{0}^{t}\Big{(}\frac{1}{N}\sum_{i=1}^{N}B^{i}_{s}\Big{)}d\beta_{s}\Big{]} 𝔼[(0r|1Ni=1NBsi|2𝑑s)1/2]𝔼[(0r|1+|z|p𝑑μsN|2𝑑s)1/2]\displaystyle\lesssim\mathbb{E}\Big{[}\Big{(}\int_{0}^{r}\Big{|}\frac{1}{N}\sum_{i=1}^{N}B^{i}_{s}\Big{|}^{2}ds\Big{)}^{1/2}\Big{]}\lesssim\mathbb{E}\Big{[}\Big{(}\int_{0}^{r}\Big{|}1+\int|z|^{p}d\mu_{s}^{N}\Big{|}^{2}ds\Big{)}^{1/2}\Big{]}
1+𝔼[(supt[0,r]|z|p𝑑μtN)1/2(0r|z|p𝑑μsN𝑑s)1/2]\displaystyle\lesssim 1+\mathbb{E}\Big{[}\Big{(}\sup_{t\in[0,r]}\int|z|^{p}d\mu_{t}^{N}\Big{)}^{1/2}\Big{(}\int_{0}^{r}\int|z|^{p}d\mu_{s}^{N}ds\Big{)}^{1/2}\Big{]}
1+𝔼[supt[0,r]|z|p𝑑μtN]1/20r𝔼[supt[0,s]|z|p𝑑μtN]1/2\displaystyle\lesssim 1+\mathbb{E}\Big{[}\sup_{t\in[0,r]}\int|z|^{p}d\mu_{t}^{N}\Big{]}^{1/2}\int_{0}^{r}\mathbb{E}\Big{[}\sup_{t\in[0,s]}\int|z|^{p}d\mu^{N}_{t}\Big{]}^{1/2}

so that we may come back to (2.7) and get

𝔼[supt[0,r]|z|p𝑑μtN]1+𝔼[|z|p𝑑μ0N]+0r𝔼[supt[0,s]|z|p𝑑μtN]𝑑s.\displaystyle\mathbb{E}\Big{[}\sup_{t\in[0,r]}\int|z|^{p}d\mu^{N}_{t}\Big{]}\lesssim 1+\mathbb{E}\Big{[}\int|z|^{p}d\mu_{0}^{N}\Big{]}+\int_{0}^{r}\mathbb{E}\Big{[}\sup_{t\in[0,s]}\int|z|^{p}d\mu_{t}^{N}\Big{]}ds.

We may now apply Gronwall’s lemma to derive the first estimate. Coming back to (2.6), a similar reasoning leads to the second estimate.

Let us now consider the space of trajectories

𝒞=C([0,T];2d),z=supt[0,T]|zt|{\cal C}=C([0,T];\mathbb{R}^{2d}),\hskip 14.22636pt\|z\|_{\infty}=\sup_{t\in[0,T]}|z_{t}|

and view the empirical measure as a (random) probability over 𝒞{\cal C}:

μN=1Ni=1NδZi,N𝒫(𝒞).\displaystyle\mu^{N}=\frac{1}{N}\sum_{i=1}^{N}\delta_{Z^{i,N}}\in{\cal P}({\cal C}).

More precisely, for p1p\geq 1, we introduce the Wasserstein space

𝒫p(𝒞)={μ𝒫(𝒞),z𝒞zpdμ<}{\cal P}_{p}({\cal C})=\left\{\mu\in{\cal P}({\cal C}),\;\;\int_{z\in{\cal C}}\|z\|_{\infty}^{p}d\mu<\infty\right\}

equipped with the usual distance

Wp[μ,ν]=infπΠ(μ,ν)((z1,z2)𝒞2z1z2p𝑑π(z1,z2))1/pW_{p}[\mu,\nu]=\inf_{\pi\in\Pi(\mu,\nu)}\Big{(}\int_{(z_{1},z_{2})\in{\cal C}^{2}}\|z_{1}-z_{2}\|_{\infty}^{p}d\pi(z_{1},z_{2})\Big{)}^{1/p}

where

Π(μ,ν)={π𝒫(𝒞2),z2𝒞π(dz1,dz2)=μ(dz1),z1𝒞π(dz1,dz2)=ν(dz2)}.\Pi(\mu,\nu)=\left\{\pi\in{\cal P}({\cal C}^{2}),\;\;\int_{z_{2}\in{\cal C}}\pi(dz_{1},dz_{2})=\mu(dz_{1}),\;\int_{z_{1}\in{\cal C}}\pi(dz_{1},dz_{2})=\nu(dz_{2})\right\}.

We may now state the following mean-field limit result.

Proposition 2.2.

Let p>1p>1. As NN\to\infty, provided that μ0Nμ0\mu_{0}^{N}\to\mu_{0} in 𝒫p(2d){\cal P}_{p}(\mathbb{R}^{2d}), we have μNμ in Lp(Ω;𝒫p(𝒞))\mu^{N}\to\mu\text{ in }L^{p}(\Omega;{\cal P}_{p}({\cal C})) where μ\mu solves the regularized SPDE

dμt+vxμtdt+v((LRCS[μt]+LRMT[μt]+SR[μt])μt)dt+v(KR[μt]μt)dβt\displaystyle d\mu_{t}+v\cdot\nabla_{x}\mu_{t}dt+\nabla_{v}\cdot\Big{(}\Big{(}L^{CS}_{R}[\mu_{t}]+L^{MT}_{R}[\mu_{t}]+S_{R}[\mu_{t}]\Big{)}\mu_{t}\Big{)}dt+\nabla_{v}\cdot\left(K_{R}[\mu_{t}]\mu_{t}\right)d\beta_{t}
=12v(v(KR[μt]KR[μt]Tμt))dt\displaystyle\hskip 170.71652pt=\frac{1}{2}\nabla_{v}\cdot\Big{(}\nabla_{v}\cdot(K_{R}[\mu_{t}]K_{R}[\mu_{t}]^{T}\mu_{t})\Big{)}dt (2.8)

in the following sense: denoting the operator

R[μ]Ψ=vxΨ+(LRCS[μ]+LRMT[μ]+SR[μ])vΨ+121i,jdKR[μ]iKR[μ]jvivj2Ψ\displaystyle{\cal L}_{R}[\mu]\Psi=v\cdot\nabla_{x}\Psi+\Big{(}L_{R}^{CS}[\mu]+L_{R}^{MT}[\mu]+S_{R}[\mu]\Big{)}\cdot\nabla_{v}\Psi+\frac{1}{2}\sum_{1\leq i,j\leq d}K_{R}[\mu]^{i}K_{R}[\mu]^{j}\partial^{2}_{v_{i}v_{j}}\Psi (2.9)

we have, for any test function ΨCc(2d)\Psi\in C_{c}^{\infty}(\mathbb{R}^{2d}),

μt,Ψ=μ0,Ψ+0t<R[μs]Ψ,μs>ds+0t<KR[μs]vΨ,μs>dβs,t[0,T].\displaystyle\langle\mu_{t},\Psi\rangle=\langle\mu_{0},\Psi\rangle+\int_{0}^{t}\Bigl{<}{\cal L}_{R}[\mu_{s}]\Psi,\mu_{s}\Bigr{>}ds+\int_{0}^{t}\Bigl{<}K_{R}[\mu_{s}]\cdot\nabla_{v}\Psi,\mu_{s}\Bigr{>}d\beta_{s},\;\;\;t\in[0,T].

More precisely, μ\mu is the unique element of L2(Ω;𝒫p(𝒞))L^{2}(\Omega;{\cal P}_{p}({\cal C})) given by the push-forward measure of the initial data by the non-linear characteristics:

μ=(Zμ)μ0 in 𝒫(𝒞),a.s\displaystyle\mu=(Z^{\mu})^{*}\mu_{0}\text{ in }{\cal P}({\cal C}),\;\;a.s (2.10)

where Zμ:z2d(Xtμ(z),Vtμ(z))t[0,T]𝒞Z^{\mu}:z\in\mathbb{R}^{2d}\mapsto(X^{\mu}_{t}(z),V^{\mu}_{t}(z))_{t\in[0,T]}\in{\cal C} is the flow associated with the SDE

{dXtμ(z)=Vtμ(z)dt,dVtμ(z)=(LRCS[μt]+LRMT[μt]+SR[μt])(Ztμ(z))dt+KR[μt](Ztμ(z))dβt,Z0μ(z)=z.\left\{\begin{array}[]{l}dX^{\mu}_{t}(z)=V^{\mu}_{t}(z)dt,\\ dV^{\mu}_{t}(z)=\Big{(}L^{CS}_{R}[\mu_{t}]+L^{MT}_{R}[\mu_{t}]+S_{R}[\mu_{t}]\Big{)}(Z^{\mu}_{t}(z))dt+K_{R}[\mu_{t}](Z^{\mu}_{t}(z))d\beta_{t},\\ Z^{\mu}_{0}(z)=z.\end{array}\right. (2.11)
Proof.

For simplicity, let us consider the case p=2p=2. We start by noticing that, for fixed N1N\geq 1, the empirical measure μN\mu^{N} naturally satisfies the fixed-point-like equation (2.10). For all N,M1N,M\geq 1, introducing an optimal plan πΠ(μ0N,μ0M)\pi\in\Pi(\mu_{0}^{N},\mu_{0}^{M}) (one may refer to [20] for details) such that

W22[μ0N,μ0M]=(z1,z2)(2d)2|z1z2|2𝑑π(z1,z2),W_{2}^{2}[\mu_{0}^{N},\mu_{0}^{M}]=\int_{(z_{1},z_{2})\in(\mathbb{R}^{2d})^{2}}|z_{1}-z_{2}|^{2}d\pi(z_{1},z_{2}),

it follows that

W22[μN,μM](z1,z2)(2d)2ZμN(z1)ZμM(z2)2dπ(z1,z2)=:JTN,M\displaystyle W_{2}^{2}[\mu^{N},\mu^{M}]\leq\int_{(z_{1},z_{2})\in(\mathbb{R}^{2d})^{2}}\|Z^{\mu^{N}}(z_{1})-Z^{\mu^{M}}(z_{2})\|^{2}_{\infty}d\pi(z_{1},z_{2})=:J^{N,M}_{T} (2.12)

and we may simply re-write

JTN,M=1iN1jMsupt[0,T]|Zti,NZtj,M|2π({z0i,N,z0j,M})\displaystyle J^{N,M}_{T}=\sum_{1\leq i\leq N}\sum_{1\leq j\leq M}\sup_{t\in~[0,T]}\Big{|}Z^{i,N}_{t}-Z^{j,M}_{t}\Big{|}^{2}\pi(\{z_{0}^{i,N},z_{0}^{j,M}\}) (2.13)

where Zi,NZ^{i,N} is the solution of (2.5). Itô’s formula easily leads to

d|Zti,NZtj,M|2=(ζt1+ζt2+ζt3+ζt4)dt+ζt5dβt\displaystyle d\Big{|}Z^{i,N}_{t}-Z^{j,M}_{t}\Big{|}^{2}=\Big{(}\zeta^{1}_{t}+\zeta^{2}_{t}+\zeta^{3}_{t}+\zeta^{4}_{t}\Big{)}dt+\zeta^{5}_{t}d\beta_{t}

where

ζt1=2(Zti,NZtj,M)(LRCS[μtN](Zti,N)LRCS[μtM](Ztj,M)),\displaystyle\zeta^{1}_{t}=2\Big{(}Z^{i,N}_{t}-Z^{j,M}_{t}\Big{)}\cdot\Big{(}L^{CS}_{R}[\mu_{t}^{N}](Z^{i,N}_{t})-L^{CS}_{R}[\mu_{t}^{M}](Z^{j,M}_{t})\Big{)},
ζt2=2(Zti,NZtj,M)(LRMT[μtN](Zti,N)LRMT[μtM](Ztj,M)),\displaystyle\zeta^{2}_{t}=2\Big{(}Z^{i,N}_{t}-Z^{j,M}_{t}\Big{)}\cdot\Big{(}L^{MT}_{R}[\mu_{t}^{N}](Z^{i,N}_{t})-L^{MT}_{R}[\mu_{t}^{M}](Z^{j,M}_{t})\Big{)},
ζt3=2(Zti,NZtj,M)(SR[μtN](Zti,N)SR[μtM](Ztj,M)),\displaystyle\zeta^{3}_{t}=2\Big{(}Z^{i,N}_{t}-Z^{j,M}_{t}\Big{)}\cdot\Big{(}S_{R}[\mu_{t}^{N}](Z^{i,N}_{t})-S_{R}[\mu_{t}^{M}](Z^{j,M}_{t})\Big{)},
ζt4=|KR[μtN](Zti,N)KR[μtM](Ztj,M)|2,\displaystyle\zeta^{4}_{t}=\Big{|}K_{R}[\mu_{t}^{N}](Z^{i,N}_{t})-K_{R}[\mu_{t}^{M}](Z^{j,M}_{t})\Big{|}^{2},
ζt5=2(Zti,NZtj,M)(KR[μtN](Zti,N)KR[μtM](Ztj,M)).\displaystyle\zeta^{5}_{t}=2\Big{(}Z^{i,N}_{t}-Z^{j,M}_{t}\Big{)}\cdot\Big{(}K_{R}[\mu_{t}^{N}](Z^{i,N}_{t})-K_{R}[\mu_{t}^{M}](Z^{j,M}_{t})\Big{)}.

Using the Lipschitz estimate (2.3), we deduce (for fixed R>0R>0)

d|Zti,NZtj,M|2\displaystyle d\Big{|}Z^{i,N}_{t}-Z^{j,M}_{t}\Big{|}^{2} (|Zti,NZtj,M|2+W1[μtN,μtM]2)dt+ζt5dβt\displaystyle\lesssim\Big{(}\Big{|}Z^{i,N}_{t}-Z^{j,M}_{t}\Big{|}^{2}+W_{1}[\mu_{t}^{N},\mu_{t}^{M}]^{2}\Big{)}dt+\zeta^{5}_{t}d\beta_{t}
(|Zti,NZtj,M|2+JtN,M)dt+ζt5dβt.\displaystyle\lesssim\Big{(}\Big{|}Z^{i,N}_{t}-Z^{j,M}_{t}\Big{|}^{2}+J_{t}^{N,M}\Big{)}dt+\zeta^{5}_{t}d\beta_{t}. (2.14)

Taking the expectation in (2.14) and applying Gronwall’s lemma leads to

𝔼|Zti,NZtj,M|2(|z0i,Nz0j,M|2+𝔼[JtN,M]).\displaystyle\mathbb{E}\Big{|}Z^{i,N}_{t}-Z^{j,M}_{t}\Big{|}^{2}\lesssim\Big{(}|z_{0}^{i,N}-z_{0}^{j,M}|^{2}+\mathbb{E}[J_{t}^{N,M}]\Big{)}.

Coming back to (2.14), we may write

supσ[0,t]|Zσi,NZσj,M|2|z0i,Nz0j,M|2+0t(|Zsi,NZsj,M|2+JsN,M)𝑑s+supσ[0,t]0σζs5𝑑βs\displaystyle\sup_{\sigma\in[0,t]}\Big{|}Z^{i,N}_{\sigma}-Z^{j,M}_{\sigma}\Big{|}^{2}\lesssim|z_{0}^{i,N}-z_{0}^{j,M}|^{2}+\int_{0}^{t}\Big{(}\Big{|}Z^{i,N}_{s}-Z^{j,M}_{s}\Big{|}^{2}+J_{s}^{N,M}\Big{)}ds+\sup_{\sigma\in[0,t]}\int_{0}^{\sigma}\zeta^{5}_{s}d\beta_{s}

and therefore

𝔼[supσ[0,t]|Zσi,NZσj,M|2]|z0i,Nz0j,M|2+0t𝔼[JsN,M]𝑑s+𝔼[supσ[0,t]0σζs5𝑑βs].\displaystyle\mathbb{E}\Big{[}\sup_{\sigma\in[0,t]}\Big{|}Z^{i,N}_{\sigma}-Z^{j,M}_{\sigma}\Big{|}^{2}\Big{]}\lesssim|z_{0}^{i,N}-z_{0}^{j,M}|^{2}+\int_{0}^{t}\mathbb{E}[J_{s}^{N,M}]ds+\mathbb{E}\Big{[}\sup_{\sigma\in[0,t]}\int_{0}^{\sigma}\zeta^{5}_{s}d\beta_{s}\Big{]}. (2.15)

Burkholder-Davis-Gundy’s inequality gives

𝔼[supσ[0,t]0σζs5𝑑βs]𝔼[(0t|ζs5|2𝑑s)1/2]\displaystyle\mathbb{E}\Big{[}\sup_{\sigma\in[0,t]}\int_{0}^{\sigma}\zeta^{5}_{s}d\beta_{s}\Big{]}\lesssim\mathbb{E}\Big{[}\Big{(}\int_{0}^{t}|\zeta^{5}_{s}|^{2}ds\Big{)}^{1/2}\Big{]}
𝔼[sups[0,t]|Zsi,NZsj,M|](0t|KR[μsN](Zsi,N)KR[μsM](Zsj,M)|2ds)1/2]\displaystyle\hskip 42.67912pt\lesssim\mathbb{E}\Big{[}\sup_{s\in[0,t]}\Big{|}Z^{i,N}_{s}-Z^{j,M}_{s}\Big{|}\Big{]}\Big{(}\int_{0}^{t}\Big{|}K_{R}[\mu_{s}^{N}](Z^{i,N}_{s})-K_{R}[\mu_{s}^{M}](Z^{j,M}_{s})\Big{|}^{2}ds\Big{)}^{1/2}\Big{]}
𝔼[sups[0,t]|Zsi,NZsj,M|2]1/2(0t𝔼|KR[μsN](Zsi,N)KR[μsM](Zsj,M)|2𝑑s)1/2.\displaystyle\hskip 42.67912pt\lesssim\mathbb{E}\Big{[}\sup_{s\in[0,t]}\Big{|}Z^{i,N}_{s}-Z^{j,M}_{s}\Big{|}^{2}\Big{]}^{1/2}\Big{(}\int_{0}^{t}\mathbb{E}\Big{|}K_{R}[\mu_{s}^{N}](Z^{i,N}_{s})-K_{R}[\mu_{s}^{M}](Z^{j,M}_{s})\Big{|}^{2}ds\Big{)}^{1/2}.

Making use of (2.3) again, we get

𝔼[supσ[0,t]0σζs5𝑑βs]\displaystyle\mathbb{E}\Big{[}\sup_{\sigma\in[0,t]}\int_{0}^{\sigma}\zeta^{5}_{s}d\beta_{s}\Big{]} 𝔼[sups[0,t]|Zsi,NZsj,M|2]1/2(0t𝔼|Zsi,NZsj,M|2+E[JsN,M]ds)1/2\displaystyle\lesssim\mathbb{E}\Big{[}\sup_{s\in[0,t]}\Big{|}Z^{i,N}_{s}-Z^{j,M}_{s}\Big{|}^{2}\Big{]}^{1/2}\Big{(}\int_{0}^{t}\mathbb{E}\Big{|}Z^{i,N}_{s}-Z^{j,M}_{s}\Big{|}^{2}+E[J_{s}^{N,M}]ds\Big{)}^{1/2}

and we may come back (2.15) to obtain

𝔼[supσ[0,t]|Zσi,NZσj,M|2]|z0i,Nz0j,M|2+0t(𝔼|Zsi,NZsj,M|2+E[JsN,M])𝑑s.\displaystyle\mathbb{E}\Big{[}\sup_{\sigma\in[0,t]}\Big{|}Z^{i,N}_{\sigma}-Z^{j,M}_{\sigma}\Big{|}^{2}\Big{]}\lesssim|z_{0}^{i,N}-z_{0}^{j,M}|^{2}+\int_{0}^{t}\Big{(}\mathbb{E}\Big{|}Z^{i,N}_{s}-Z^{j,M}_{s}\Big{|}^{2}+E[J_{s}^{N,M}]\Big{)}ds.

Summing over ii and jj as in (2.13) finally leads to

𝔼[JtN,M]W22[μ0N,μ0M]+0t𝔼[JsN,M]𝑑s.\displaystyle\mathbb{E}\Big{[}J^{N,M}_{t}\Big{]}\lesssim W_{2}^{2}[\mu_{0}^{N},\mu_{0}^{M}]+\int_{0}^{t}\mathbb{E}[J^{N,M}_{s}]ds.

Gronwall’s lemma hence gives 𝔼[JTN,M]W22[μ0N,μ0M]\mathbb{E}[J_{T}^{N,M}]\lesssim W_{2}^{2}[\mu_{0}^{N},\mu_{0}^{M}] so that, coming back to (2.12), we get

𝔼[W22[μN,μM]]W22[μ0N,μ0M]0\displaystyle\mathbb{E}\Big{[}W_{2}^{2}[\mu^{N},\mu^{M}]\Big{]}\lesssim W_{2}^{2}[\mu_{0}^{N},\mu_{0}^{M}]\to 0

as N,MN,M are sent to infinity. We have shown that μN\mu^{N} converges to some μ\mu in the complete space L2(Ω;𝒫2(𝒞))L^{2}(\Omega;{\cal P}_{2}({\cal C})). Let us now prove that μ\mu satisfies the fixed-point identity (2.10). First, since 𝔼[z𝒞z2𝑑μ]<\mathbb{E}\Big{[}\int_{z\in{\cal C}}\|z\|_{\infty}^{2}d\mu\Big{]}~{}<~{}\infty, one could easily deduce from the sub-linearity of the coefficients that

𝔼[supt[0,T]|Ztμ(z)|2]<,\displaystyle\mathbb{E}[\sup_{t\in[0,T]}\Big{|}Z^{\mu}_{t}(z)\Big{|}^{2}\Big{]}<\infty,

thereby guaranteeing that the solution Ztμ(z)Z^{\mu}_{t}(z) of (2.11) is unique and global. Moreover, it is a well known fact (see e.g [13]) that the flow Zμ:z2d(Ztμ(z))t[0,T]Z^{\mu}:z\in\mathbb{R}^{2d}\mapsto(Z^{\mu}_{t}(z))_{t\in[0,T]} is almost-surely continuous, so that the push-forward measure involved in (2.10) is indeed well-defined. This could be seen in this case by establishing a Kolmogorov estimate

𝔼[supt[0,T]|Ztμ(z)Ztμ(z)|2]|zz|2,z,z2d.\mathbb{E}\Big{[}\sup_{t\in[0,T]}\Big{|}Z^{\mu}_{t}(z)-Z^{\mu}_{t}(z^{\prime})\Big{|}^{2}\Big{]}\lesssim|z-z^{\prime}|^{2},\;\;\forall z,z^{\prime}\in\mathbb{R}^{2d}.

Let us introduce the measure ν=(Zμ)μ0\nu=(Z^{\mu})^{*}\mu_{0}. Introducing an optimal plan πΠ(μ0N,μ0)\pi~{}\in~{}\Pi(\mu_{0}^{N},\mu_{0}) so that

W22[μ0N,μ0]=(z1,z2)(2d)2|z1z2|2𝑑π(z1,z2)W_{2}^{2}[\mu_{0}^{N},\mu_{0}]=\int_{(z_{1},z_{2})\in(\mathbb{R}^{2d})^{2}}|z_{1}-z_{2}|^{2}d\pi(z_{1},z_{2})

we have this time

W22[μN,ν](z1,z2)(2d)2supt[0,T]|ZtμN(z1)Ztμ(z2)|2dπ(z1,z2)=:JTN.W_{2}^{2}[\mu^{N},\nu]\leq\int_{(z_{1},z_{2})\in(\mathbb{R}^{2d})^{2}}\sup_{t\in[0,T]}\Big{|}Z_{t}^{\mu^{N}}(z_{1})-Z_{t}^{\mu}(z_{2})\Big{|}^{2}d\pi(z_{1},z_{2})=:J^{N}_{T}.

As in (2.14), Itô’s formula gives an expression of the form

d|ZtμN(z1)Ztμ(z2)|2\displaystyle d\Big{|}Z_{t}^{\mu^{N}}(z_{1})-Z_{t}^{\mu}(z_{2})\Big{|}^{2} (|ZtμN(z1)Ztμ(z2)|2+W12[μtN,μt])dt+ζtdβt\displaystyle\lesssim\Big{(}\Big{|}Z_{t}^{\mu^{N}}(z_{1})-Z_{t}^{\mu}(z_{2})\Big{|}^{2}+W_{1}^{2}[\mu_{t}^{N},\mu_{t}]\Big{)}dt+\zeta_{t}d\beta_{t}
(|ZtμN(z1)Ztμ(z2)|2+W22[μN,μ])dt+ζtdβt.\displaystyle\lesssim\Big{(}\Big{|}Z_{t}^{\mu^{N}}(z_{1})-Z_{t}^{\mu}(z_{2})\Big{|}^{2}+W_{2}^{2}[\mu^{N},\mu]\Big{)}dt+\zeta_{t}d\beta_{t}.

Proceeding as in the first part of the proof, we eventually obtain

𝔼[W22[μN,ν]]𝔼[W22[μN,μ]].\mathbb{E}\Big{[}W_{2}^{2}[\mu^{N},\nu]\Big{]}\lesssim\mathbb{E}\Big{[}W_{2}^{2}[\mu^{N},\mu]\Big{]}.

Letting NN go to infinity, we conclude that μ=ν\mu=\nu a.s, that is exactly (2.10).

We may once again use the same arguments to prove that the fixed-point-like equation (2.10) has a unique solution: considering μ\mu and ν\nu such that μ=(Zμ)μ0\mu=(Z^{\mu})^{*}\mu_{0} a.s and ν=(Zν)ν0\nu=(Z^{\nu})^{*}\nu_{0} a.s, we are led to

𝔼[W22[μ,ν]]W22[μ0,ν0]\mathbb{E}\Big{[}W_{2}^{2}[\mu,\nu]\Big{]}\lesssim W_{2}^{2}[\mu_{0},\nu_{0}]

so that μ0=ν0\mu_{0}=\nu_{0} implies μ=ν\mu=\nu a.s.

Finally, let us notice that any μ\mu satisfying (2.10) defines a solution of (2.8). Indeed, for any test function ΨCc(2d)\Psi\in C_{c}^{\infty}(\mathbb{R}^{2d}), Itô’s formula gives exactly

Ψ(Ztμ(z))=Ψ(z)+0tR[μs]Ψ(Zsμ(z))𝑑s+0tΨ(Zsμ(z))KR[μs](Zsμ(z))𝑑βs.\Psi(Z^{\mu}_{t}(z))=\Psi(z)+\int_{0}^{t}{\cal L}_{R}[\mu_{s}]\Psi(Z^{\mu}_{s}(z))ds+\int_{0}^{t}\nabla\Psi(Z^{\mu}_{s}(z))\cdot K_{R}[\mu_{s}](Z^{\mu}_{s}(z))d\beta_{s}.

where R[μ]Ψ{\cal L}_{R}[\mu]\Psi is given by (2.9). Since Ψ,μt=z2dΨ(Ztμ(z))𝑑μ0(z)\langle\Psi,\mu_{t}\rangle=\int_{z\in\mathbb{R}^{2d}}\Psi(Z^{\mu}_{t}(z))d\mu_{0}(z), integrating with respect to dμ0(z)d\mu_{0}(z) using a stochastic Fubini theorem gives the expected result. ∎

2.2 Flow of characteristics, regular solutions

For some fixed R>0R>0, let us consider the unique solution μ\mu of (2.8) constructed in Proposition 2.2. Although this measure, as well as the associated characteristics (Ztμ)t0(Z^{\mu}_{t})_{t\geq 0}, depend on R>0R>0, we shall hide this dependence in the following expressions to avoid cluttering notation. One may note from expressions (2.2) that the coefficients

LRCS[μt](z),LRMT[μt](z),SR[μt](z),KR[μt](z)L^{CS}_{R}[\mu_{t}](z),\;\;\;L^{MT}_{R}[\mu_{t}](z),\;\;\;S_{R}[\mu_{t}](z),\;\;\;K_{R}[\mu_{t}](z)

involved in the SDE (2.11) have the regularity C4(2d)C^{4}(\mathbb{R}^{2d}) in the zz variable. More precisely, assumption (1.11) guarantees that, for fixed R>0R>0, for 1|α|41\leq|\alpha|\leq 4,

|zαLRCS[μt]|+|zαLRMT[μt]|+|zαSR[μt]|+|zαKR[μt]|1\displaystyle\Big{|}\partial_{z}^{\alpha}L^{CS}_{R}[\mu_{t}]\Big{|}+\Big{|}\partial_{z}^{\alpha}L^{MT}_{R}[\mu_{t}]\Big{|}+\Big{|}\partial_{z}^{\alpha}S_{R}[\mu_{t}]\Big{|}+\Big{|}\partial_{z}^{\alpha}K_{R}[\mu_{t}]\Big{|}\lesssim 1

uniformly in t[0,T]t\in[0,T] and ωΩ\omega\in\Omega. In particular, the first, second and third order zz-derivatives of the coefficients are globally Lipschitz-continuous. As a result, Theorem 4.4 of [13], Chapter II, yields

[t[0,T],Ztμ:z2dZtμ(z)2d is a C3-diffeomorphism],a.s.\displaystyle\Big{[}\;\forall t\in[0,T],\;\;Z^{\mu}_{t}:z\in\mathbb{R}^{2d}\mapsto Z^{\mu}_{t}(z)\in\mathbb{R}^{2d}\text{ is a $C^{3}$-diffeomorphism}\;\Big{]},\;\;\mathbb{P}-a.s. (2.16)

More precisely, for 0st0\leq s\leq t, denoting by Zs,tμ(z)Z^{\mu}_{s,t}(z) the solution of the SDE

{Xs,tμ(z)=x+r=stVs,rμ(z)𝑑r,Vs,tμ(z)=v+r=st(LRCS[μr]+LRMT[μr]+SR[μr])(Zs,rμ(z))𝑑r+r=stKR[μr](Zs,rμ(z))𝑑βr,\left\{\begin{array}[]{l}X^{\mu}_{s,t}(z)=x+\displaystyle{\int_{r=s}^{t}V^{\mu}_{s,r}(z)dr},\\ \vspace{-3mm}\\ V^{\mu}_{s,t}(z)=v+\displaystyle{\int_{r=s}^{t}\Big{(}L^{CS}_{R}[\mu_{r}]+L^{MT}_{R}[\mu_{r}]+S_{R}[\mu_{r}]\Big{)}(Z^{\mu}_{s,r}(z))dr+\int_{r=s}^{t}K_{R}[\mu_{r}](Z^{\mu}_{s,r}(z))d\beta_{r}},\end{array}\right.

the inverse map (Zs,tμ)1(z)(Z^{\mu}_{s,t})^{-1}(z) satisfies the corresponding backward SDE

{(Xs,tμ)1(z)=xr=st(Vr,tμ)1(z)𝑑r,(Vs,tμ)1(z)=vr=st(LRCS[μr]+LRMT[μr]+SR[μr]S~R[μr])((Zr,tμ)1(z))𝑑rr=stKR[μr]((Zr,tμ)1(z))dβr^,\left\{\begin{array}[]{l l}(X^{\mu}_{s,t})^{-1}(z)&=x-\displaystyle{\int_{r=s}^{t}(V^{\mu}_{r,t})^{-1}(z)dr},\\ \vspace{-3mm}\\ (V^{\mu}_{s,t})^{-1}(z)&=v-\displaystyle{\int_{r=s}^{t}\Big{(}L^{CS}_{R}[\mu_{r}]+L^{MT}_{R}[\mu_{r}]+S_{R}[\mu_{r}]-\widetilde{S}_{R}[\mu_{r}]\Big{)}((Z^{\mu}_{r,t})^{-1}(z))dr}\\ &\hskip 142.26378pt-\displaystyle{\int_{r=s}^{t}K_{R}[\mu_{r}]((Z^{\mu}_{r,t})^{-1}(z))\widehat{d\beta_{r}}},\end{array}\right.

where S~R[μ]=vKR[μ]KR[μ]\widetilde{S}_{R}[\mu]=\nabla_{v}K_{R}[\mu]K_{R}[\mu] and stdβt^\int_{s}^{t}\cdot\;\widehat{d\beta_{t}} denotes the backward Stratonovich integral (see again [13] Theorem 7.3 for this result and p.194 for the definition of the backward integral). When s=0s=0, we simply denote

Ztμ(z)=Z0,tμ(z),(Ztμ)1(z)=(Z0,tμ)1(z).\displaystyle Z^{\mu}_{t}(z)=Z^{\mu}_{0,t}(z),\;\;\;(Z^{\mu}_{t})^{-1}(z)=(Z^{\mu}_{0,t})^{-1}(z).

In the particular case where the initial measure μ0\mu_{0} admits a density f0L1(2d)f_{0}~{}\in~{}L^{1}(\mathbb{R}^{2d}) with respect to the Lebesgue measure on 2d\mathbb{R}^{2d}, for any test function ΨCb(2d)\Psi\in C_{b}(\mathbb{R}^{2d}) we may write

z2dΨ(z)𝑑μt(z)=z2dΨ(Ztμ(z))f0(z)𝑑z=z2dΨ(z)|Jt(z)1|f0((Ztμ)1(z))𝑑z\displaystyle\int_{z\in\mathbb{R}^{2d}}\Psi(z)d\mu_{t}(z)=\int_{z\in\mathbb{R}^{2d}}\Psi(Z^{\mu}_{t}(z))f_{0}(z)dz=\int_{z\in\mathbb{R}^{2d}}\Psi(z)|J_{t}(z)^{-1}|f_{0}((Z^{\mu}_{t})^{-1}(z))dz

where Jt(z)1J_{t}(z)^{-1} denotes the jacobian determinant

Jt(z)1=det[Dz((Ztμ)1)(z)].\displaystyle J_{t}(z)^{-1}=\det\Big{[}D_{z}((Z^{\mu}_{t})^{-1})(z)\Big{]}. (2.17)

Considering a countable separating family of such test functions Ψ\Psi, from μ=(Zμ)f0\mu=(Z^{\mu})^{*}f_{0} in 𝒫(𝒞){\cal P}({\cal C}), we deduce that

[t[0,T],dμt(z)=f(t,z)dz],a.s,\Big{[}\forall t\in[0,T],\;\;d\mu_{t}(z)=f(t,z)dz\Big{]},\;\;\mathbb{P}-a.s,

where

f(t,z)=Jt(z)1f0((Ztμ)1(z)).\displaystyle f(t,z)=J_{t}(z)^{-1}f_{0}((Z^{\mu}_{t})^{-1}(z)). (2.18)

Note that we may drop the absolute value in (2.18) since Jt(z)10J_{t}(z)^{-1}\geq 0 a.s. Let us now give some estimates regarding the forward and backward characteristics.

Proposition 2.3.

For all p1p\geq 1, let f0L1(2d)f_{0}\in L^{1}(\mathbb{R}^{2d}) such that |z|pf0(z)𝑑s<\int|z^{\prime}|^{p}f_{0}(z)ds^{\prime}<\infty. Let μL2(Ω;𝒫2(𝒞))\mu\in L^{2}(\Omega;{\cal P}_{2}({\cal C})) such that μ=(Zμ)f0\mu=(Z^{\mu})^{*}f_{0} as in (2.10). Then for all z,z1,z22dz,z_{1},z_{2}\in\mathbb{R}^{2d}, t,t1,t2[0,T]t,t_{1},t_{2}\in[0,T] and 1k31\leq k\leq 3,

𝔼[supt[0,T]|Ztμ(z)|p]1+|z|p+|z|pf0(z)𝑑z,\displaystyle\mathbb{E}\Big{[}\sup_{t\in[0,T]}|Z_{t}^{\mu}(z)|^{p}\Big{]}\lesssim 1+|z|^{p}+\int|z^{\prime}|^{p}f_{0}(z)dz^{\prime}, (2.19)
𝔼[|Dzk(Ztμ)(z)|p]1+|z|pf0(z)𝑑z,\displaystyle\mathbb{E}\Big{[}|D^{k}_{z}(Z_{t}^{\mu})(z)|^{p}\Big{]}\lesssim 1+\int|z^{\prime}|^{p}f_{0}(z)dz^{\prime},
𝔼[supt[0,T]|Ztμ(z1)Ztμ(z2)|p]|z1z2|p,\displaystyle\mathbb{E}\Big{[}\sup_{t\in[0,T]}|Z_{t}^{\mu}(z_{1})-Z_{t}^{\mu}(z_{2})|^{p}\Big{]}\lesssim|z_{1}-z_{2}|^{p}, (2.20)
𝔼[|(Ztμ)1(z)|p]|z|p+|z|pf0(z)𝑑z,\displaystyle\mathbb{E}\Big{[}|(Z_{t}^{\mu})^{-1}(z)|^{p}\Big{]}\lesssim|z|^{p}+\int|z^{\prime}|^{p}f_{0}(z)dz^{\prime},
𝔼[|(Zt1μ)1)(z1)(Zt2μ)1)(z2)|p]|z1z2|p+(1+|z1|p+|z|pf0(z)dz)|t1t2|p/2,\displaystyle\mathbb{E}\Big{[}\Big{|}(Z^{\mu}_{t_{1}})^{-1})(z_{1})-(Z^{\mu}_{t_{2}})^{-1})(z_{2})\Big{|}^{p}\Big{]}\lesssim|z_{1}-z_{2}|^{p}+\Big{(}1+|z_{1}|^{p}+\int|z^{\prime}|^{p}f_{0}(z)dz^{\prime}\Big{)}|t_{1}-t_{2}|^{p/2}, (2.21)
𝔼[|Dzk((Ztμ)1)(z)|p]1+|z|pf0(z)𝑑z,\displaystyle\mathbb{E}\Big{[}|D^{k}_{z}((Z_{t}^{\mu})^{-1})(z)|^{p}\Big{]}\lesssim 1+\int|z^{\prime}|^{p}f_{0}(z)dz^{\prime}, (2.22)
𝔼[|Dzk((Zt1μ)1)(z)Dzk((Zt2μ)1)(z)|p]|zz|p+|t1t2|p/2.\displaystyle\mathbb{E}\Big{[}\Big{|}D^{k}_{z}((Z^{\mu}_{t_{1}})^{-1})(z)-D^{k}_{z}((Z^{\mu}_{t_{2}})^{-1})(z^{\prime})\Big{|}^{p}\Big{]}\lesssim|z-z^{\prime}|^{p}+|t_{1}-t_{2}|^{p/2}. (2.23)

The constants involved in \lesssim depend on p,k,T,Rp,k,T,R only.

These estimates are deduced in a classical manner from the sub-linearity and the global Lipschitz continuity of the coefficients of the equations satisfied by Ztμ(z)Z_{t}^{\mu}(z) and (Ztμ)1(z)(Z_{t}^{\mu})^{-1}(z). Applying Kolmogorov’s lemma to (2.21) and (2.23), we deduce that, for all 0k30\leq k\leq 3,

(t,z)[0,T]×2dDk((Ztμ)1)(z)\displaystyle(t,z)\in[0,T]\times\mathbb{R}^{2d}\mapsto D^{k}((Z^{\mu}_{t})^{-1})(z) (2.24)

is continuous a.s\mathbb{P}-a.s. Additionally, we may deduce from (2.19) and (2.20) the following estimate: given a compact set K2dK\subset\mathbb{R}^{2d} and z0Kz_{0}\in K, for any p1p\geq 1 and α(0,1)\alpha\in(0,1),

𝔼[supzKsupt[0,T]|Ztμ(z)|p]𝔼[supt[0,T]|Ztμ(z0)|p]+𝔼[ZμCzαLtp]diam(K)αpCK,T,p<.\displaystyle\mathbb{E}\Big{[}\sup_{z\in K}\sup_{t\in[0,T]}|Z_{t}^{\mu}(z)|^{p}\Big{]}\lesssim\mathbb{E}\Big{[}\sup_{t\in[0,T]}|Z_{t}^{\mu}(z_{0})|^{p}\Big{]}+\mathbb{E}\Big{[}\|Z^{\mu}\|_{C^{\alpha}_{z}L^{\infty}_{t}}^{p}\Big{]}\text{diam}(K)^{\alpha p}\leq C_{K,T,p}<\infty. (2.25)

In (2.25), diam(K)=sup{|z1z2|,z1,z2K}\text{diam}(K)=\sup\left\{|z_{1}-z_{2}|,\,z_{1},z_{2}\in K\right\} and ZμCzαLt\|Z^{\mu}\|_{C^{\alpha}_{z}L^{\infty}_{t}} denotes the α\alpha-Hölder semi-norm

ZμCzαLt=supz1z22dsupt[0,T]|Ztμ(z1)Ztμ(z2)||z1z2|α\|Z^{\mu}\|_{C^{\alpha}_{z}L^{\infty}_{t}}=\sup_{z_{1}\neq z_{2}\in\mathbb{R}^{2d}}\sup_{t\in[0,T]}\frac{|Z^{\mu}_{t}(z_{1})-Z^{\mu}_{t}(z_{2})|}{|z_{1}-z_{2}|^{\alpha}}

which satisfies indeed 𝔼[ZμCzαLtp]1\mathbb{E}\Big{[}\|Z^{\mu}\|_{C^{\alpha}_{z}L^{\infty}_{t}}^{p}\Big{]}\lesssim 1 by applying Kolmogorov’s lemma to (2.20). We may now establish the following result.

Proposition 2.4 (Regular solution).

Let f0Cc2(2d)f_{0}\in C_{c}^{2}(\mathbb{R}^{2d}). Let μL2(Ω;𝒫2(𝒞))\mu\in L^{2}(\Omega;{\cal P}_{2}({\cal C})) such that μ=(Zμ)f0\mu~{}=~{}(Z^{\mu})^{*}f_{0} as in (2.10), and let f(t,z)f(t,z) be defined as in (2.18), so that μ\mu may be represented as dμt(z)=f(t,z)dzd\mu_{t}(z)=f(t,z)dz. Then ff is a regular solution of (2.8) in the following sense:

  • \mathbb{P}-almost surely, f(t,)C2(2d)f(t,\cdot)\in C^{2}(\mathbb{R}^{2d}) for all t[0,T]t\in[0,T] and the maps

    (t,z)[0,T]×2dzαf(t,z),  0|α|2(t,z)\in[0,T]\times\mathbb{R}^{2d}\mapsto\partial_{z}^{\alpha}f(t,z),\;\;0\leq|\alpha|\leq 2

    are continuous.

  • Denoting the operator

    (R[f])g=vxgv((LRCS[f]+LRMT[f]+SR[f])g)+12vv(KR[f]KR[f]Tg),\displaystyle({\cal L}_{R}[f])^{*}g=-v\cdot\nabla_{x}g-\nabla_{v}\cdot\Big{(}(L_{R}^{CS}[f]+L_{R}^{MT}[f]+S_{R}[f])g\Big{)}+\frac{1}{2}\nabla_{v}\cdot\nabla_{v}\Big{(}K_{R}[f]K_{R}[f]^{T}g\Big{)}, (2.26)

    we have for all z2dz\in\mathbb{R}^{2d}, t[0,T]t\in[0,T], \mathbb{P}-almost surely,

    f(t,z)=f0(z)+0t(R[f(s)])f(s,z)𝑑s0tv(KR[f(s)]f(s,z))𝑑βs.\displaystyle f(t,z)=f_{0}(z)+\int_{0}^{t}({\cal L}_{R}[f(s)])^{*}f(s,z)ds-\int_{0}^{t}\nabla_{v}\cdot\Big{(}K_{R}[f(s)]f(s,z)\Big{)}d\beta_{s}. (2.27)
Proof.

As a consequence of (2.24), it is clear from expression (2.18) that the first condition is met. Furthermore, f(t,)f(t,\cdot) is almost surely compactly supported, uniformly in t[0,T]t\in[0,T], with

t[0,T],Supp(f(t,)){z2d,|z|supzKsupt[0,T]|Ztμ(z)|}.\displaystyle\forall t\in[0,T],\;\;\text{Supp}(f(t,\cdot))\subset\left\{z\in\mathbb{R}^{2d},\;\;|z|\leq\sup_{z^{\prime}\in K}\sup_{t\in[0,T]}|Z^{\mu}_{t}(z^{\prime})|\right\}.

Since ff is a solution of (2.8), for any ΨCc(2d)\Psi\in C_{c}^{\infty}(\mathbb{R}^{2d}), \mathbb{P}-almost surely,

f(t),Ψ=f0,Ψ+0t<(R[f(s)])f(s),Ψ>ds0tv(KR[f(s)]f(s)),Ψ𝑑βs.\langle f(t),\Psi\rangle=\langle f_{0},\Psi\rangle+\int_{0}^{t}\Bigl{<}({\cal L}_{R}[f(s)])^{*}f(s),\Psi\Bigr{>}ds-\int_{0}^{t}\langle\nabla_{v}\cdot\Big{(}K_{R}[f(s)]f(s)\Big{)},\Psi\rangle d\beta_{s}.

We can then interchange the integrals

0t,Ψds=0t,Ψds,0t,Ψdβs=0t,Ψdβs.\int_{0}^{t}\langle\cdot,\Psi\rangle ds=\langle\int_{0}^{t}\cdot,\Psi\rangle ds,\hskip 14.22636pt\int_{0}^{t}\langle\cdot,\Psi\rangle d\beta_{s}=\langle\int_{0}^{t}\cdot,\Psi\rangle d\beta_{s}.

Since all functions are compactly supported (for fixed ωΩ\omega\in\Omega), the integrals with respect to dsds cause no issue. As for the stochastic integral, we may use a stochastic Fubini theorem, as long as

0tz2d𝔼[|v(KR[f(s)]f(s))Ψ(z)|2]𝑑z𝑑s<.\int_{0}^{t}\int_{z\in\mathbb{R}^{2d}}\mathbb{E}\Big{[}\Big{|}\nabla_{v}\cdot\Big{(}K_{R}[f(s)]f(s)\Big{)}\Psi(z)\Big{|}^{2}\Big{]}dzds<\infty.

From expressions (2.2), we deduce (for fixed R>0R>0)

|KR[f(t)](z)|+|v(KR[f(t)])(z)|1,\displaystyle|K_{R}[f(t)](z)|+|\nabla_{v}\cdot(K_{R}[f(t)])(z)|\lesssim 1,

From expression (2.18), (and |det[A]||A|2d|\det[A]|\lesssim|A|^{2d}) it is clear that

|f(t,z)|2+|zf(t,z)|2(1+maxk=1,2|Dzk((Ztμ)1)(z)|q)(f0L2+zf0L2)|f(t,z)|^{2}+|\nabla_{z}f(t,z)|^{2}\lesssim\Big{(}1+\max_{k=1,2}\Big{|}D^{k}_{z}((Z^{\mu}_{t})^{-1})(z)\Big{|}^{q}\Big{)}\Big{(}\|f_{0}\|_{L^{\infty}}^{2}+\|\nabla_{z}f_{0}\|_{L^{\infty}}^{2}\Big{)}

for some q=q(d)1q=q(d)\geq 1. We may hence write

𝔼|v(KR[f(s)]f(s))(z)|2\displaystyle\mathbb{E}\Big{|}\nabla_{v}\cdot\Big{(}K_{R}[f(s)]f(s)\Big{)}(z)\Big{|}^{2} 1+𝔼[maxk=1,2|Dzk((Ztμ)1)(z)|q]1\displaystyle\lesssim 1+\mathbb{E}\Big{[}\max_{k=1,2}\Big{|}D^{k}_{z}((Z^{\mu}_{t})^{-1})(z)\Big{|}^{q}\Big{]}\lesssim 1 (2.28)

thanks to the bound (2.22). It follows that

0tz2d𝔼[|v(KR[f(s)]f(s))Ψ(z)|2]𝑑z𝑑sz2d|Ψ(z)|2𝑑z<.\int_{0}^{t}\int_{z\in\mathbb{R}^{2d}}\mathbb{E}\Big{[}\Big{|}\nabla_{v}\cdot\Big{(}K_{R}[f(s)]f(s)\Big{)}\Psi(z)\Big{|}^{2}\Big{]}dzds\lesssim\int_{z\in\mathbb{R}^{2d}}|\Psi(z)|^{2}dz<\infty.

Consequently, \mathbb{P}-a.s, (2.27) holds when integrated against any ΨCc(2d)\Psi\in C_{c}^{\infty}(\mathbb{R}^{2d}). We deduce that (2.27) holds for almost every z2dz\in\mathbb{R}^{2d}. Since both sides of (2.27) are continuous with respect to zz (thanks to (2.28) again), we conclude that the equality holds for every z2dz\in\mathbb{R}^{2d}.

3 Weak convergence of approximate solutions

From now on, let us fix some initial data f0f_{0} satisfying (at least) for some δ>1\delta>1,

f00,z2df0(z)𝑑z=1,z2d|z|δf0(z)𝑑z<.f_{0}\geq 0,\hskip 14.22636pt\int_{z\in\mathbb{R}^{2d}}f_{0}(z)dz=1,\hskip 14.22636pt\int_{z\in\mathbb{R}^{2d}}|z|^{\delta}f_{0}(z)dz<\infty.

For any R>0R>0, considering the particle system (2.5), with initial data satisfying

μ0N(dz)f0(z)dz in 𝒫δ(2d),\mu_{0}^{N}(dz)\to f_{0}(z)dz\text{ in }{\cal P}_{\delta}(\mathbb{R}^{2d}),

we may introduce the solution μR\mu_{R} of (2.8) constructed in Proposition 2.2. As previously discussed, we naturally identify μR\mu^{R} with its density fR=(fR(t,z))t[0,T],z2d{f^{R}}~{}=~{}({f^{R}}(t,z))_{t\in[0,T],z\in\mathbb{R}^{2d}} defined by (2.18).

3.1 Uniform estimates

In this section, we shall establish some estimates on fR{f^{R}} uniformly on the regularization parameter RR.

Proposition 3.1.

Let p1p\geq 1 and f0Lp(2d)f_{0}\in L^{p}(\mathbb{R}^{2d}). Then fRL([0,T];Lp(2d)){f^{R}}\in L^{\infty}([0,T];L^{p}(\mathbb{R}^{2d})) a.s, with the estimate

𝔼[supt[0,T]fR(t)Lpp]f0Lpp.\mathbb{E}\Big{[}\sup_{t\in[0,T]}\|{f^{R}}(t)\|_{L^{p}}^{p}\Big{]}\lesssim\|f_{0}\|_{L^{p}}^{p}.

The constant involved in \lesssim depends on pp and TT only.

Proof.

Let us start by considering f0Cc2(2d)f_{0}\in C^{2}_{c}(\mathbb{R}^{2d}) supported in some compact K2dK\subset\mathbb{R}^{2d}. Then fR{f^{R}} is a regular solution of (2.8) in the sense of Proposition 2.4. For any z2dz\in\mathbb{R}^{2d}, applying Itô’s formula to |fR(t,z)|p|{f^{R}}(t,z)|^{p} hence gives

|fR(t,z)|p\displaystyle|{f^{R}}(t,z)|^{p} =f0(z)pp0t|fR(s,z)|p1[vxfR(s,z)+v(BR[fR(s)]fR(s,z))]𝑑s\displaystyle=f_{0}(z)^{p}-p\int_{0}^{t}|{f^{R}}(s,z)|^{p-1}\Big{[}v\cdot\nabla_{x}{f^{R}}(s,z)+\nabla_{v}\cdot\Big{(}B_{R}[{f^{R}}(s)]{f^{R}}(s,z)\Big{)}\Big{]}ds
+p20t|fR(s,z)|p1vv(KR[fR(s)]KR[fR(s)]TfR(s,z))𝑑s\displaystyle+\frac{p}{2}\int_{0}^{t}|{f^{R}}(s,z)|^{p-1}\nabla_{v}\cdot\nabla_{v}\cdot\Big{(}K_{R}[{f^{R}}(s)]K_{R}[{f^{R}}(s)]^{T}{f^{R}}(s,z)\Big{)}ds
+p(p1)20t|fR(s,z)|p2|v(KR[fR(s)]fR(s,z))|2𝑑s\displaystyle+\frac{p(p-1)}{2}\int_{0}^{t}|{f^{R}}(s,z)|^{p-2}\Big{|}\nabla_{v}\cdot\Big{(}K_{R}[{f^{R}}(s)]{f^{R}}(s,z)\Big{)}\Big{|}^{2}ds
p0t|fR(s,z)|p1v(KR[fR(s)]fR(s,z))𝑑βs,\displaystyle-p\int_{0}^{t}|{f^{R}}(s,z)|^{p-1}\nabla_{v}\cdot\Big{(}K_{R}[{f^{R}}(s)]{f^{R}}(s,z)\Big{)}d\beta_{s}, (3.1)

where BR[f]B_{R}[f] denotes the drift coefficient LRCS[f]+LRMT[f]+SR[f]L^{CS}_{R}[f]+L^{MT}_{R}[f]+S_{R}[f]. We can now integrate (3.1) with respect to dzdz and interchange the integrals:

z2d0t𝑑s𝑑z=0tz2d𝑑z𝑑s,z2d0t𝑑βs𝑑z=0tz2d𝑑z𝑑βs,\int_{z\in\mathbb{R}^{2d}}\int_{0}^{t}\cdot\;\;dsdz=\int_{0}^{t}\int_{z\in\mathbb{R}^{2d}}\cdot\;\;dzds,\hskip 14.22636pt\int_{z\in\mathbb{R}^{2d}}\int_{0}^{t}\cdot\;\;d\beta_{s}dz=\int_{0}^{t}\int_{z\in\mathbb{R}^{2d}}\cdot\;\;dzd\beta_{s},

Since all the integrands in (3.1) are compactly supported in zz uniformly for t[0,T]t\in[0,T] (for fixed ωΩ\omega\in\Omega), the integrals with respect to dsds cause no issue. As for the stochastic integral, we may use a stochastic Fubnini theorem if we can justify that

:=𝔼[0Tz2d|fR(s,z)p1v(KR[fR(s)]fR(s,z))|2𝑑z𝑑s]<.\displaystyle{\cal E}:=\mathbb{E}\Big{[}\int_{0}^{T}\int_{z\in\mathbb{R}^{2d}}\Big{|}{f^{R}}(s,z)^{p-1}\nabla_{v}\cdot\Big{(}K_{R}[{f^{R}}(s)]{f^{R}}(s,z)\Big{)}\Big{|}^{2}dzds\Big{]}<\infty. (3.2)

As in the proof of Proposition 2.4, one can see that (with μ(dz)=fR(z)dz\mu(dz)=f^{R}(z)dz)

|fR(s,z)p1v(KR[fR(s)]fR(s,z))|21+maxk=1,2|Dzk((Ztμ)1)(z)|q\Big{|}{f^{R}}(s,z)^{p-1}\nabla_{v}\cdot\Big{(}K_{R}[{f^{R}}(s)]{f^{R}}(s,z)\Big{)}\Big{|}^{2}\lesssim 1+\max_{k=1,2}~\Big{|}D^{k}_{z}((Z^{\mu}_{t})^{-1})(z)\Big{|}^{q}

for some q=q(d,p)1q=q(d,p)\geq 1. Denoting S=supt[0,T]supzK|Ztμ(z)|S=\displaystyle{\sup_{t\in[0,T]}\sup_{z^{\prime}\in K}|Z_{t}^{\mu}(z^{\prime})|}, it follows that

z2d|fR(s,z)p1v(KR[fR(s)]fR(s,z))|2𝑑z|z|Smaxk=1,2|Dzk((Ztμ)1)(z)|qdz\displaystyle\int_{z\in\mathbb{R}^{2d}}\Big{|}{f^{R}}(s,z)^{p-1}\nabla_{v}\cdot\Big{(}K_{R}[{f^{R}}(s)]{f^{R}}(s,z)\Big{)}\Big{|}^{2}dz\lesssim\int_{|z|\leq S}\max_{k=1,2}\Big{|}D^{k}_{z}((Z^{\mu}_{t})^{-1})(z)\Big{|}^{q}dz

and, fixing some m>2dm>2d,

|z|Smaxk=1,2|Dzk((Ztμ)1)(z)|qdz=|z|S(1+|z|)m/2maxk=1,2|Dzk((Ztμ)1)(z)|q(1+|z|)m/2𝑑z\displaystyle\int_{|z|\leq S}\max_{k=1,2}\Big{|}D^{k}_{z}((Z^{\mu}_{t})^{-1})(z)\Big{|}^{q}dz=\int_{|z|\leq S}(1+|z|)^{m/2}\frac{\max_{k=1,2}\Big{|}D^{k}_{z}((Z^{\mu}_{t})^{-1})(z)\Big{|}^{q}}{(1+|z|)^{m/2}}~dz
|z|S(1+|z|)m𝑑z+z2dmaxk=1,2|Dzk((Ztμ)1)(z)|2qdz(1+|z|)m\displaystyle\hskip 28.45274pt\lesssim\int_{|z|\leq S}(1+|z|)^{m}dz+\int_{z\in\mathbb{R}^{2d}}~\max_{k=1,2}\Big{|}D^{k}_{z}((Z^{\mu}_{t})^{-1})(z)\Big{|}^{2q}\frac{dz}{(1+|z|)^{m}}
(1+Sm+1)+z2dmaxk=1,2|Dzk((Ztμ)1)(z)|2qdz(1+|z|)m.\displaystyle\hskip 28.45274pt\lesssim(1+S^{m+1})+\int_{z\in\mathbb{R}^{2d}}~\max_{k=1,2}\Big{|}D^{k}_{z}((Z^{\mu}_{t})^{-1})(z)\Big{|}^{2q}\frac{dz}{(1+|z|)^{m}}.

Using (2.25) and (2.22), we deduce

1+𝔼[Sm+1]+z2d𝔼[maxk=1,2|Dzk((Ztμ)1)(z)|2q]dz(1+|z|)m<{\cal E}\lesssim 1+\mathbb{E}\Big{[}S^{m+1}\Big{]}+\int_{z\in\mathbb{R}^{2d}}\mathbb{E}\Big{[}\max_{k=1,2}\Big{|}D^{k}_{z}((Z^{\mu}_{t})^{-1})(z)\Big{|}^{2q}\Big{]}\frac{dz}{(1+|z|)^{m}}<\infty

We may hence integrate (3.1) with respect to zz, which leads to

2d|fR(t,z)|p𝑑z\displaystyle\int_{\mathbb{R}^{2d}}|{f^{R}}(t,z)|^{p}dz =2df0(z)p𝑑z+I1(t)+I2(t)+I3(t)+I4(t),\displaystyle=\int_{\mathbb{R}^{2d}}f_{0}(z)^{p}dz+I_{1}(t)+I_{2}(t)+I_{3}(t)+I_{4}(t), (3.3)

where

I1(t)=p0t2d|fR(s,z)|p1[vxfR(s,z)+v(BR[fR(s)]fR(s,z))]𝑑z𝑑s,\displaystyle I_{1}(t)=-p\int_{0}^{t}\int_{\mathbb{R}^{2d}}|{f^{R}}(s,z)|^{p-1}\Big{[}v\cdot\nabla_{x}{f^{R}}(s,z)+\nabla_{v}\cdot\Big{(}B_{R}[{f^{R}}(s)]{f^{R}}(s,z)\Big{)}\Big{]}dzds,
I2(t)=p20t2d|fR(s,z)|p1v2(KR[fR(s)]KR[fR(s)]TfR(s,z))dzds,\displaystyle I_{2}(t)=\frac{p}{2}\int_{0}^{t}\int_{\mathbb{R}^{2d}}|{f^{R}}(s,z)|^{p-1}\nabla_{v}^{2}\Big{(}K_{R}[{f^{R}}(s)]K_{R}[{f^{R}}(s)]^{T}{f^{R}}(s,z)\Big{)}dzds,
I3(t)=p(p1)20t2d|fR(s,z)|p2|v(KR[fR(s)]fR(s,z))|2𝑑z𝑑s,\displaystyle I_{3}(t)=\frac{p(p-1)}{2}\int_{0}^{t}\int_{\mathbb{R}^{2d}}|{f^{R}}(s,z)|^{p-2}\Big{|}\nabla_{v}\cdot\Big{(}K_{R}[{f^{R}}(s)]{f^{R}}(s,z)\Big{)}\Big{|}^{2}dzds,
I4(t)=p0t2d|fR(s,z)|p1v(KR[fR(s)]fR(s,z))𝑑z𝑑βs.\displaystyle I_{4}(t)=-p\int_{0}^{t}\int_{\mathbb{R}^{2d}}|{f^{R}}(s,z)|^{p-1}\nabla_{v}\cdot\Big{(}K_{R}[{f^{R}}(s)]{f^{R}}(s,z)\Big{)}dzd\beta_{s}.

Simple calculations lead to the classical identity

|fR|p1v(BRfR)𝑑z=p1p|fR|p(vBR)𝑑z\displaystyle\int|f^{R}|^{p-1}\nabla_{v}\cdot(B_{R}f^{R})dz=\frac{p-1}{p}\int|f^{R}|^{p}(\nabla_{v}\cdot B_{R})dz (3.4)

so that

I1(t)\displaystyle I_{1}(t) =p1p0t2d|fR(s,z)|p(vBR[f(s)](z))𝑑z𝑑s\displaystyle=\frac{p-1}{p}\int_{0}^{t}\int_{\mathbb{R}^{2d}}{|f^{R}(s,z)|}^{p}(\nabla_{v}\cdot B_{R}[f(s)](z))dzds
vBR[fR]Lt,z([0,T]×2d)0t2d|fR(s,z)|p𝑑z𝑑s.\displaystyle\leq\Big{\|}\nabla_{v}\cdot B_{R}[{f^{R}}]\Big{\|}_{L_{t,z}^{\infty}([0,T]\times\mathbb{R}^{2d})}\int_{0}^{t}\int_{\mathbb{R}^{2d}}|f^{R}(s,z)|^{p}dzds.

From expressions (2.2), we derive

vLRCS[f](z)=d×dχR(xy)ψ(xy)v(θR(wv))f(y,w)𝑑y𝑑w,\displaystyle\nabla_{v}\cdot L^{CS}_{R}[f](z)=\int_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\chi_{R}(x-y)\psi(x-y)\nabla_{v}\cdot\Big{(}\theta_{R}(w-v)\Big{)}f(y,w)dydw,
vLRMT(z)=d,\displaystyle\nabla_{v}\cdot L^{MT}_{R}(z)=-d,
vKR[f](z)=d×dχR(xy)ψ~(xy)v(θR(wv))f(y,w)𝑑y𝑑w,\displaystyle\nabla_{v}\cdot K_{R}[f](z)=\int_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\chi_{R}(x-y)\widetilde{\psi}(x-y)\nabla_{v}\cdot\Big{(}\theta_{R}(w-v)\Big{)}f(y,w)dydw,
vSR[f](z)=12(d×dψ~(xy)f(y,w)𝑑y𝑑w)vKR[f](z)\displaystyle\nabla_{v}\cdot S_{R}[f](z)=-\frac{1}{2}\Big{(}\int_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\widetilde{\psi}(x-y)f(y,w)dydw\Big{)}\nabla_{v}\cdot K_{R}[f](z)

Assumptions (1.11), (2.1) guarantee that these terms are bounded uniformly in t[0,T]t\in[0,T], z2dz\in\mathbb{R}^{2d}, ωΩ\omega\in\Omega, R>0R>0, so that

I1(t)0t2d|fR(s,z)|p𝑑z𝑑s.I_{1}(t)\lesssim\int_{0}^{t}\int_{\mathbb{R}^{2d}}|f^{R}(s,z)|^{p}dzds.

Similarly, we have the following identity:

|fR|p1v2(KRKRTfR)dz+(p1)|fR|p2|v(KRfR)|2𝑑z\displaystyle\int|f^{R}|^{p-1}\nabla_{v}^{2}(K_{R}K_{R}^{T}f^{R})dz+(p-1)\int|f^{R}|^{p-2}|\nabla_{v}\cdot(K_{R}f^{R})|^{2}dz
=(p1)|fR|p2[|v(KRfR)|2vfRv(KRKRTfR)]𝑑z\displaystyle\;\;=(p-1)\int|f^{R}|^{p-2}\Big{[}|\nabla_{v}\cdot(K_{R}f^{R})|^{2}-\nabla_{v}f^{R}\cdot\nabla_{v}\cdot(K_{R}K_{R}^{T}f^{R})\Big{]}dz
=(p1)|fR|p|vKR|2\displaystyle\;\;=(p-1)\int|f^{R}|^{p}|\nabla_{v}\cdot K_{R}|^{2}

so that, again,

I2(t)+I3(t)=p(p1)20t2d|fR(s,z)|p|vKR[fR(s)]|2𝑑z𝑑s0t2d|fR(s,z)|p𝑑z𝑑s.I_{2}(t)+I_{3}(t)=\frac{p(p-1)}{2}\int_{0}^{t}\int_{\mathbb{R}^{2d}}|f^{R}(s,z)|^{p}\Big{|}\nabla_{v}\cdot K_{R}[f^{R}(s)]\Big{|}^{2}dzds\lesssim\int_{0}^{t}\int_{\mathbb{R}^{2d}}|f^{R}(s,z)|^{p}dzds.

Since (3.2) guarantees that I4(t)I_{4}(t) defines a (square integrable) martingale, we may take the expectation in (3.3) and apply Gronwall’s lemma to derive

𝔼[2d|fR(t,z)|p𝑑z]f0Lpp.\displaystyle\mathbb{E}\Big{[}\int_{\mathbb{R}^{2d}}|{f^{R}}(t,z)|^{p}dz\Big{]}\lesssim\|f_{0}\|_{L^{p}}^{p}.

It follows that

𝔼[supσ[0,t]2d|fR(σ,z)|p𝑑z]0t𝔼[2d|fR(s,z)|p𝑑z]𝑑s+𝔼[supσ[0,t]I4(σ)]\mathbb{E}\Big{[}\sup_{\sigma\in[0,t]}\int_{\mathbb{R}^{2d}}|{f^{R}}(\sigma,z)|^{p}dz\Big{]}\lesssim\int_{0}^{t}\mathbb{E}\Big{[}\int_{\mathbb{R}^{2d}}|{f^{R}}(s,z)|^{p}dz\Big{]}ds+\mathbb{E}\Big{[}\sup_{\sigma\in[0,t]}I_{4}(\sigma)\Big{]}

and Burkholder-Davis-Gundy’s inequality yields (making use of (3.4) again)

𝔼[supσ[0,t]I4(σ)]\displaystyle\mathbb{E}\Big{[}\sup_{\sigma\in[0,t]}I_{4}(\sigma)\Big{]} 𝔼[(0t|2d|fR|p1v(KR[fR]fR)𝑑z|2𝑑σ)1/2]\displaystyle\lesssim\mathbb{E}\Big{[}\Big{(}\int_{0}^{t}\Big{|}\int_{\mathbb{R}^{2d}}|{f^{R}}|^{p-1}\nabla_{v}\cdot(K_{R}[{f^{R}}]{f^{R}})dz\Big{|}^{2}d\sigma\Big{)}^{1/2}\Big{]}
𝔼[(0t|2d|fR(σ,z)|p𝑑z|2𝑑σ)1/2]\displaystyle\lesssim\mathbb{E}\Big{[}\Big{(}\int_{0}^{t}\Big{|}\int_{\mathbb{R}^{2d}}|{f^{R}}(\sigma,z)|^{p}dz\Big{|}^{2}d\sigma\Big{)}^{1/2}\Big{]}
12𝔼[supσ[0,t]2d|fR(σ,z)|p𝑑z]+C0t𝔼[2d|fR(σ,z)|p𝑑z]𝑑σ\displaystyle\leq\frac{1}{2}\mathbb{E}\Big{[}\sup_{\sigma\in[0,t]}\int_{\mathbb{R}^{2d}}|{f^{R}}(\sigma,z)|^{p}dz\Big{]}+C\int_{0}^{t}\mathbb{E}\Big{[}\int_{\mathbb{R}^{2d}}|{f^{R}}(\sigma,z)|^{p}dz\Big{]}d\sigma

which gives the expected result. We now extend the estimate to any f0Lp(2d)f_{0}\in L^{p}(\mathbb{R}^{2d}) satisfying |z|δf0(z)𝑑z<\int|z|^{\delta}f_{0}(z)dz<\infty by considering a sequence of densities f0kCc2(2d)f_{0}^{k}~{}\in~{}C_{c}^{2}(\mathbb{R}^{2d}) such that, as kk goes to infinity,

f0kf0 a.s and in Lp(2d),\displaystyle f_{0}^{k}\to f_{0}\text{ a.s and in }L^{p}(\mathbb{R}^{2d}),
supk1z2d|z|δfk(z)𝑑z<.\displaystyle\sup_{k\geq 1}\int_{z\in\mathbb{R}^{2d}}|z|^{\delta}f_{k}(z)dz<\infty.

It is easy to see that these assumptions imply in particular

f0kf0 in 𝒫r(2d)f_{0}^{k}\to f_{0}\text{ in }{\cal P}_{r}(\mathbb{R}^{2d})

for any 1<r<δ1<r<\delta. Denoting by fkf^{k} the solution of (2.8) with initial data f0kf_{0}^{k} constructed in Proposition 2.2, we may deduce (as in the proof of Proposition 2.2),

𝔼[supt[0,T]Wrr[ftk,ft]]𝔼[Wrr[fk,f]]Wrr[f0k,f0]0.\displaystyle\mathbb{E}\Big{[}\sup_{t\in[0,T]}W_{r}^{r}[f^{k}_{t},f_{t}]\Big{]}\leq\mathbb{E}\Big{[}W_{r}^{r}[f^{k},f]\Big{]}\lesssim W_{r}^{r}[f_{0}^{k},f_{0}]\to 0.

Up to a subsequence, we may hence assume that

supt[0,T]Wr[ftk,ft]0,a.s.\displaystyle\sup_{t\in[0,T]}W_{r}[f_{t}^{k},f_{t}]\to 0,\;\mathbb{P}-a.s. (3.5)

From the estimates

𝔼[supt[0,T]fk(t)Lpp]f0kLpp,k1,\mathbb{E}\Big{[}\sup_{t\in[0,T]}\|f^{k}(t)\|_{L^{p}}^{p}\Big{]}\lesssim\|f_{0}^{k}\|_{L^{p}}^{p},\;k\geq 1,

we derive that (fk)k1(f^{k})_{k\geq 1} is bounded in LωpLtLzpL^{p}_{\omega}L^{\infty}_{t}L^{p}_{z} and therefore, up to a subsequence

fkg weak  in Lp(Ω;L([0,T];Lp(2d)))\displaystyle f^{k}\rightharpoonup g\text{ weak $*$ in }L^{p}\Big{(}\Omega;L^{\infty}([0,T];L^{p}(\mathbb{R}^{2d}))\Big{)} (3.6)

where gg satisfies the bound

𝔼[supt[0,T]g(t)Lpp]lim supkf0kLpp=f0Lpp.\displaystyle\mathbb{E}\Big{[}\sup_{t\in[0,T]}\|g(t)\|_{L^{p}}^{p}\Big{]}\lesssim\limsup_{k}\|f_{0}^{k}\|_{L^{p}}^{p}=\|f_{0}\|_{L^{p}}^{p}. (3.7)

Let us consider ΨCc([0,T]×2d)\Psi\in C_{c}([0,T]\times\mathbb{R}^{2d}), ξL(Ω)\xi\in L^{\infty}(\Omega) and introduce Φ(ω,t,z)=ξ(ω)Ψ(t,z)\Phi(\omega,t,z)=\xi(\omega)\Psi(t,z). From (3.5) we deduce

[t[0,T],zfk(t,z)Ψ(t,z)𝑑zzf(t,z)Ψ(t,z)𝑑z],a.s.\displaystyle\Big{[}\forall t\in[0,T],\;\int_{z}f^{k}(t,z)\Psi(t,z)dz\to\int_{z}f(t,z)\Psi(t,z)dz\Big{]},\;\mathbb{P}-a.s.

and the bound |zfk(t,z)Ψ(t,z)𝑑z|ΨLt,z\Big{|}\int_{z}f^{k}(t,z)\Psi(t,z)dz\Big{|}\leq\|\Psi\|_{L^{\infty}_{t,z}} guarantees

t=0Tzfk(t,z)Ψ(t,z)𝑑z𝑑tt=0Tzf(t,z)Ψ(t,z)𝑑z𝑑t,a.s.\displaystyle\int_{t=0}^{T}\int_{z}f^{k}(t,z)\Psi(t,z)dzdt\to\int_{t=0}^{T}\int_{z}f(t,z)\Psi(t,z)dzdt,\;\mathbb{P}-a.s.

Finally, the bound |t=0Tzξ(ω)fk(ω,t,z)Ψ(t,z)𝑑z𝑑t|TξLωΨLt,z\Big{|}\int_{t=0}^{T}\int_{z}\xi(\omega)f^{k}(\omega,t,z)\Psi(t,z)dzdt\Big{|}\leq T\|\xi\|_{L^{\infty}_{\omega}}\|\Psi\|_{L^{\infty}_{t,z}} guarantees

𝔼[ξ(ω)t=0Tzfk(ω,t,z)Ψ(t,z)𝑑z𝑑t]\displaystyle\mathbb{E}\Big{[}\xi(\omega)\int_{t=0}^{T}\int_{z}f^{k}(\omega,t,z)\Psi(t,z)dzdt\Big{]}\to 𝔼[ξ(ω)t=0Tzf(ω,t,z)Ψ(t,z)𝑑z𝑑t]\displaystyle\mathbb{E}\Big{[}\xi(\omega)\int_{t=0}^{T}\int_{z}f(\omega,t,z)\Psi(t,z)dzdt\Big{]}

so that, according to (3.6),

𝔼[ξt=0Tzg(t,z)Ψ(t,z)𝑑z𝑑t]=𝔼[ξt=0Tzf(t,z)Ψ(t,z)𝑑z𝑑t].\displaystyle\mathbb{E}\Big{[}\xi\int_{t=0}^{T}\int_{z}g(t,z)\Psi(t,z)dzdt\Big{]}=\mathbb{E}\Big{[}\xi\int_{t=0}^{T}\int_{z}f(t,z)\Psi(t,z)dzdt\Big{]}.

We easily derive that f=gf=g in LωpLtLzpL^{p}_{\omega}L^{\infty}_{t}L^{p}_{z} and the bound (3.7) concludes the proof. ∎

Proposition 3.2.

For all f:2d+f:\mathbb{R}^{2d}\to\mathbb{R}^{+} and p1p\geq 1,

z2d|uR[f](x)|pf(z)𝑑zz2d|v|pf(z)𝑑z.\displaystyle\int_{z\in\mathbb{R}^{2d}}|u_{R}[f](x)|^{p}f(z)dz\lesssim\int_{z\in\mathbb{R}^{2d}}|v|^{p}f(z)dz.

The constant involved in \lesssim depends on ϕ\phi only.

Proof.

It is clear from the expression of uR[f]u_{R}[f] (2.2) and Jensen’s inequality that

z2d|uR[f](x)|pf(z)𝑑z\displaystyle\int_{z\in\mathbb{R}^{2d}}|u_{R}[f](x)|^{p}f(z)dz xvyw|w|pϕ(xy)f(y,w)𝑑y𝑑wywϕ(xy)f(y,w)𝑑y𝑑wf(x,v)𝑑x𝑑v\displaystyle\leq\int_{x}\int_{v}\frac{\int_{y}\int_{w}|w|^{p}\phi(x-y)f(y,w)dydw}{\int_{y}\int_{w}\phi(x-y)f(y,w)dydw}f(x,v)dxdv
=yw(xρ(x)ρ~(x)ϕ(xy)𝑑x)|w|pf(y,w)𝑑x𝑑y𝑑w\displaystyle=\int_{y}\int_{w}\Big{(}\int_{x}\frac{\rho(x)}{\widetilde{\rho}(x)}\phi(x-y)dx\Big{)}|w|^{p}f(y,w)dxdydw

where

ρ(x)=vdf(x,v),ρ~(x)=(ϕρ)(x)=ydwdϕ(xy)f(y,w)𝑑y𝑑w.\displaystyle\rho(x)=\int_{v\in\mathbb{R}^{d}}f(x,v),\hskip 14.22636pt\widetilde{\rho}(x)=(\phi*\rho)(x)=\int_{y\in\mathbb{R}^{d}}\int_{w\in\mathbb{R}^{d}}\phi(x-y)f(y,w)dydw.

The desired estimate hence follows from the inequality

yd,xdρ(x)ρ~(x)ϕ(xy)𝑑xC(ϕ)\displaystyle\forall y\in\mathbb{R}^{d},\;\;\int_{x\in\mathbb{R}^{d}}\frac{\rho(x)}{\widetilde{\rho}(x)}\phi(x-y)dx\leq C(\phi) (3.8)

where, with assumption (1.12) in mind, C(ϕ)C(\phi) is some constant proportional to

C(ϕ)supB(0,r2)ϕinfB(0,r1)ϕ(R/r)d.\displaystyle C(\phi)\propto\frac{\sup_{B(0,r_{2})}\phi}{\inf_{B(0,r_{1})}\phi}(R/r)^{d}. (3.9)

The proof of (3.8) is given in [12], Lemma 5.2.

Proposition 3.3.

Let k2k\geq 2, 1<δ21<\delta\leq 2 and f0f_{0} be a density satisfying z(|x|δ+|v|k)f0(z)𝑑z<\int_{z}(|x|^{\delta}+|v|^{k})f_{0}(z)dz<\infty. Then,

𝔼[supt[0,T]2d|v|kfR(t,z)𝑑z]1+2d|v|kf0(z)𝑑z.\displaystyle\mathbb{E}\Big{[}\sup_{t\in[0,T]}\int_{\mathbb{R}^{2d}}|v|^{k}{f^{R}}(t,z)dz\Big{]}\lesssim 1+\int_{\mathbb{R}^{2d}}|v|^{k}f_{0}(z)dz.
𝔼[supt[0,T]2d|x|δfR(t,z)𝑑z]1+2d(|x|δ+|v|2)f0(z)𝑑z.\displaystyle\mathbb{E}\Big{[}\sup_{t\in[0,T]}\int_{\mathbb{R}^{2d}}|x|^{\delta}{f^{R}}(t,z)dz\Big{]}\lesssim 1+\int_{\mathbb{R}^{2d}}(|x|^{\delta}+|v|^{2})f_{0}(z)dz.

The constants involved in \lesssim depends on kk, δ\delta, TT and ϕ\phi only.

Proof.

The first estimate should first be established with the stopping time

τM=inf{t0,z2d|v|kft(z)𝑑zM}T,\tau_{M}=\inf\left\{t\geq 0,\;\;\int_{z\in\mathbb{R}^{2d}}|v|^{k}f_{t}(z)dz\geq M\right\}\wedge T,

which should then be sent to TT, as MM goes to infinity. For the sake of simplicity, we omit this stopping time in the following. Let us denote f(t,z)=fR(t,z)f(t,z)={f^{R}}(t,z) and the associated characteristics Zt(z)=ZtμR(z)Z_{t}(z)~{}=~{}Z_{t}^{{\mu^{R}}}(z) (with μR(dz)=fR(z)dz\mu^{R}(dz)=f^{R}(z)dz), satisfying (2.11). We have

z2d|z|kf(t,z)𝑑z=z2d|Zt(z)|kf0(z)𝑑z\int_{z\in\mathbb{R}^{2d}}|z|^{k}f(t,z)dz=\int_{z\in\mathbb{R}^{2d}}\Big{|}Z_{t}(z)\Big{|}^{k}f_{0}(z)dz

and Itô’s formula gives, for fixed z2dz\in\mathbb{R}^{2d},

d[|Vt(z)|k]=\displaystyle d\Big{[}|V_{t}(z)|^{k}\Big{]}= k|Vt(z)|k2Vt(z)(LRCS[ft]+LRMT[ft]+SR[ft])(Zt(z))dt\displaystyle k|V_{t}(z)|^{k-2}V_{t}(z)\cdot\Big{(}L^{CS}_{R}[f_{t}]+L^{MT}_{R}[f_{t}]+S_{R}[f_{t}]\Big{)}(Z_{t}(z))dt
+k|Vt(z)|k2Vt(z)KR[ft](Zt(z))dβt\displaystyle+k|V_{t}(z)|^{k-2}V_{t}(z)\cdot K_{R}[f_{t}](Z_{t}(z))d\beta_{t}
+k(k/21)|Vt|k4|Vt(z)KR[ft](Zt(z))|2dt+k|Vt(z)|k2|KR[ft](Zt(z))|2dt.\displaystyle+k(k/2-1)|V_{t}|^{k-4}\Big{|}V_{t}(z)\cdot K_{R}[f_{t}](Z_{t}(z))\Big{|}^{2}dt+k|V_{t}(z)|^{k-2}\Big{|}K_{R}[f_{t}](Z_{t}(z))\Big{|}^{2}dt.

From the uniform sublinearity (2.4), we derive

|Vt(z)|k=|v|k+0t(as1(z)+as2(z))𝑑s+0tas3(z)𝑑βs,\displaystyle|V_{t}(z)|^{k}=|v|^{k}+\int_{0}^{t}\Big{(}a^{1}_{s}(z)+a^{2}_{s}(z)\Big{)}ds+\int_{0}^{t}a^{3}_{s}(z)d\beta_{s}, (3.10)

where

|at1(z)|+|at3(z)|(1+|Vt(z)|k+z|v|kft(z)𝑑z),\displaystyle|a^{1}_{t}(z)|+|a^{3}_{t}(z)|\lesssim\Big{(}1+|V_{t}(z)|^{k}+\int_{z^{\prime}}|v^{\prime}|^{k}f_{t}(z^{\prime})dz^{\prime}\Big{)},
at2(z)=k|Vt(z)|k2Vt(z)(uR[ft](Xt(z))Vt(z)).\displaystyle a^{2}_{t}(z)=k|V_{t}(z)|^{k-2}V_{t}(z)\cdot\Big{(}u_{R}[f_{t}](X_{t}(z))-V_{t}(z)\Big{)}.

We may then integrate (3.10) with respect to f0(z)dzf_{0}(z)dz using a stochastic Fubini theorem, which leads to

z2d|v|kf(t,z)𝑑z=z2d|v|kf0(z)𝑑z+0t(As1+As2)𝑑s+0tAs3𝑑βs,\displaystyle\int_{z\in\mathbb{R}^{2d}}|v|^{k}f(t,z)dz=\int_{z\in\mathbb{R}^{2d}}|v|^{k}f_{0}(z)dz+\int_{0}^{t}\Big{(}A_{s}^{1}+A_{s}^{2}\Big{)}ds+\int_{0}^{t}A_{s}^{3}d\beta_{s}, (3.11)

where

|At1|+|At3|(1+z|v|kft(z)𝑑z),\displaystyle|A^{1}_{t}|+|A^{3}_{t}|\lesssim\Big{(}1+\int_{z}|v|^{k}f_{t}(z)dz\Big{)}, (3.12)
At2=kz2d|v|k2v(uR[ft](x)v)ft(z)𝑑z.\displaystyle A^{2}_{t}=k\int_{z\in\mathbb{R}^{2d}}|v|^{k-2}v\cdot\Big{(}u_{R}[f_{t}](x)-v\Big{)}f_{t}(z)dz.

To deal with At2A^{2}_{t}, one may write

v(uv)=14(|v+(uv)|2|v(uv)|2)14|u|2v\cdot(u-v)=\frac{1}{4}\Big{(}|v+(u-v)|^{2}-|v-(u-v)|^{2}\Big{)}\leq\frac{1}{4}|u|^{2}

so that

z|v|k2v(u(x)v)f(z)𝑑z\displaystyle\int_{z}|v|^{k-2}v\cdot(u(x)-v)f(z)dz 14z|v|k2|u(x)|2f(z)𝑑z\displaystyle\leq\frac{1}{4}\int_{z}|v|^{k-2}|u(x)|^{2}f(z)dz
14z|v|kf(z)𝑑z+14z|u(x)|kf(z)𝑑z.\displaystyle\leq\frac{1}{4}\int_{z}|v|^{k}f(z)dz+\frac{1}{4}\int_{z}|u(x)|^{k}f(z)dz.

From Proposition 3.2, we hence deduce

|At2||v|kft(z)𝑑z.\displaystyle|A^{2}_{t}|\lesssim\int|v|^{k}f_{t}(z)dz. (3.13)

From SDE (3.11) with the sublinear terms (3.12) and (3.13), using Gronwall’s lemma and Burkholder-Davis-Gundy’s inequality, we classically obtain the first estimate. Moreover, from

|Xt(z)|δ=|x+0tVs(z)𝑑s|δ1+|x|δ+0t|Vs(z)|2𝑑s\displaystyle|X_{t}(z)|^{\delta}=\Big{|}x+\int_{0}^{t}V_{s}(z)ds\Big{|}^{\delta}\lesssim 1+|x|^{\delta}+\int_{0}^{t}|V_{s}(z)|^{2}ds

since δ2\delta\leq 2, we derive the second estimate, which concludes the proof. ∎

3.2 Stochastic averaging lemma

Proposition 3.4.

Let us assume that the initial data f0f_{0} satisfies, for some θ(0,1)\theta\in(0,1),

z2d|f0(z)|p𝑑z+z2d(1+|v|k)f0(z)𝑑z< with p=1+1θ,k>41θ.\displaystyle\int_{z\in\mathbb{R}^{2d}}|f_{0}(z)|^{p}dz+\int_{z\in\mathbb{R}^{2d}}(1+|v|^{k})f_{0}(z)dz<\infty\;\text{ with }\;p=1+\frac{1}{\theta},\hskip 11.38109ptk>\frac{4}{1-\theta}. (3.14)

For all, φCc(d)\varphi\in C_{c}^{\infty}(\mathbb{R}^{d}), letting η=1/6\eta=1/6, the averaged quantity

ρφR(t,x)=dφ(v)fR(t,x,v)𝑑x\displaystyle\rho_{\varphi}^{R}(t,x)=\int_{\mathbb{R}^{d}}\varphi(v)f^{R}(t,x,v)dx

lies in L2([0,T];Hη(d))L^{2}([0,T];H^{\eta}(\mathbb{R}^{d})) almost surely, with the estimate

𝔼[ρφRLt2Hxη2]1.\displaystyle\mathbb{E}\Big{[}\|{\rho_{\varphi}^{R}}\|_{L^{2}_{t}H^{\eta}_{x}}^{2}\Big{]}\lesssim 1. (3.15)

The constant involved in \lesssim in (3.15) depends on f0f_{0}, φ\varphi, TT and ϕ\phi only.

Proof.

The proof of this result is based on a classical L2L^{2} averaging lemma (see e.g [2] for the deterministic case), which we adapt here to the stochastic case, in a similar fashion to [9], Lemma 4.3. Let us first consider some initial data f0Cc2(2d)f_{0}\in C_{c}^{2}(\mathbb{R}^{2d}) so that f=fRf={f^{R}} is a regular solution of (2.8) in the sense of Proposition 2.4, which may be written as follows: \mathbb{P}-a.s, for all z2dz\in\mathbb{R}^{2d},

df(t,z)+vxf(t,z)dt=\displaystyle df(t,z)+v\cdot\nabla_{x}f(t,z)dt= (1idviGi(t,z)+1i,jdvivj2Gij(t,z))dt\displaystyle\Big{(}\sum_{1\leq i\leq d}\partial_{v_{i}}G^{i}(t,z)+\sum_{1\leq i,j\leq d}\partial^{2}_{v_{i}v_{j}}G^{ij}(t,z)\Big{)}dt
+(1idviHi(t,z))dβt\displaystyle+\Big{(}\sum_{1\leq i\leq d}\partial_{v_{i}}H^{i}(t,z)\Big{)}d\beta_{t} (3.16)

where we have introduced the coefficients

{Gi(t,z)=(LRCS[f(t)]+LRMT[f(t)]+SR[f(t)])(z)if(t,z),Gij(t,z)=KR[f(t)](z)iKR[f(t)](z)jf(t,z),Hi(t,z)=KR[f(t)]if(t,z).\left\{\begin{array}[]{l}G^{i}(t,z)=\Big{(}L^{CS}_{R}[f(t)]+L^{MT}_{R}[f(t)]+S_{R}[f(t)]\Big{)}(z)^{i}f(t,z),\vspace{1mm}\\ G^{ij}(t,z)=K_{R}[f(t)](z)^{i}K_{R}[f(t)](z)^{j}f(t,z),\vspace{1mm}\\ H^{i}(t,z)=K_{R}[f(t)]^{i}f(t,z).\end{array}\right. (3.17)

Let us fix ξd\xi\in\mathbb{R}^{d}. For simplicity, let us drop the summation signs in (3.16) and integrate it with respect to eiξxdxe^{-i\xi\cdot x}dx. This is possible (for every vdv\in\mathbb{R}^{d}) thanks to the bound (3.2) with p=1p=1 established previously. Denoting the xx-Fourier transform

f^(t,ξ,v)=xdeiξxf(t,x,v)𝑑x,\hat{f}(t,\xi,v)=\int_{x\in\mathbb{R}^{d}}e^{-i\xi\cdot x}f(t,x,v)dx,

we are led to

df^(t,ξ,v)+ivξf^(t,ξ,v)dt=+(viGi^+vivj2Gij^)(t,ξ,v)dt.+viHi^(t,ξ,v)dβt\displaystyle d\hat{f}(t,\xi,v)+iv\cdot\xi\hat{f}(t,\xi,v)dt=+\Big{(}\widehat{\partial_{v_{i}}G^{i}}+\widehat{\partial^{2}_{v_{i}v_{j}}G^{ij}}\Big{)}(t,\xi,v)dt.+\widehat{\partial_{v_{i}}H^{i}}(t,\xi,v)d\beta_{t}

Therefore, introducing some λ=λ(ξ)>0\lambda=\lambda(\xi)>0, we get

df^(t,ξ,v)+(λ+ivξ)f^(t,ξ,v)dt=λf^(t,ξ,v)dt\displaystyle d\hat{f}(t,\xi,v)+(\lambda+iv\cdot\xi)\hat{f}(t,\xi,v)dt=\lambda\hat{f}(t,\xi,v)dt +(viGi^+vivj2Gij^)(t,ξ,v)dt\displaystyle+\Big{(}\widehat{\partial_{v_{i}}G^{i}}+\widehat{\partial^{2}_{v_{i}v_{j}}G^{ij}}\Big{)}(t,\xi,v)dt
+viHi^(t,ξ,v)dβt,\displaystyle+\widehat{\partial_{v_{i}}H^{i}}(t,\xi,v)d\beta_{t},

from which we deduce the expression

f^(t,ξ,v)=\displaystyle\hat{f}(t,\xi,v)= e(λ+ivξ)tf0^(ξ,v)+λ0te(λ+ivξ)(ts)f^(s,ξ,v)𝑑s\displaystyle\;e^{-(\lambda+iv\cdot\xi)t}\hat{f_{0}}(\xi,v)+\lambda\int_{0}^{t}e^{-(\lambda+iv\cdot\xi)(t-s)}\hat{f}(s,\xi,v)ds
+0te(λ+ivξ)(ts)(viGi^+vivj2Gij^)(s,ξ,v)𝑑s\displaystyle+\int_{0}^{t}e^{-(\lambda+iv\cdot\xi)(t-s)}\Big{(}\widehat{\partial_{v_{i}}G^{i}}+\widehat{\partial^{2}_{v_{i}v_{j}}G^{ij}}\Big{)}(s,\xi,v)ds
+0te(λ+ivξ)(ts)viHi^(s,ξ,v)𝑑βs.\displaystyle+\int_{0}^{t}e^{-(\lambda+iv\cdot\xi)(t-s)}\widehat{\partial_{v_{i}}H^{i}}(s,\xi,v)d\beta_{s}. (3.18)

We now integrate (3.18) with respect to φ(v)dv\varphi(v)dv. This is possible since one could show that

𝔼0tvd|φ(v)e(λ+ivξ)(ts)viHi^(s,ξ,v)|2𝑑v𝑑s\displaystyle\mathbb{E}\int_{0}^{t}\int_{v\in\mathbb{R}^{d}}\Big{|}\varphi(v)e^{-(\lambda+iv\cdot\xi)(t-s)}\widehat{\partial_{v_{i}}H^{i}}(s,\xi,v)\Big{|}^{2}dvds
𝔼0Tvd(1+|v|2)|(xd|v(KR[f(s)]f(s))|dx)2dvds<\displaystyle\hskip 14.22636pt\lesssim\mathbb{E}\int_{0}^{T}\int_{v\in\mathbb{R}^{d}}(1+|v|^{2})\Big{|}(\int_{x\in\mathbb{R}^{d}}\Big{|}\nabla_{v}\cdot(K_{R}[f(s)]f(s))\Big{|}dx\Big{)}^{2}dvds<\infty

for fixed R>0R>0, by a method similar to the one employed to establish (3.2) in the proof Proposition 3.1. Introducing the (x,v)(x,v)-Fourier transform

(f)(t,ξ,ζ)=vdxdei(ξx+ζv)f(t,x,v)𝑑x𝑑v,({\cal F}f)(t,\xi,\zeta)=\int_{v\in\mathbb{R}^{d}}\int_{x\in\mathbb{R}^{d}}e^{-i(\xi\cdot x+\zeta\cdot v)}f(t,x,v)dxdv,

equation (3.18) leads to

ρφ^(t,ξ)=\displaystyle\widehat{\rho_{\varphi}}(t,\xi)= eλt(f0φ)(ξ,ξt)+λ0teλ(ts)(fφ)(s,ξ,ξ(ts))𝑑s\displaystyle e^{-\lambda t}{\cal F}(f_{0}\varphi)(\xi,\xi t)+\lambda\int_{0}^{t}e^{-\lambda(t-s)}{\cal F}(f\varphi)(s,\xi,\xi(t-s))ds
+0teλ(ts)(((viGi)φ)+((vivj2Gij)φ))(s,ξ,ξ(ts))𝑑s\displaystyle+\int_{0}^{t}e^{-\lambda(t-s)}\Big{(}{\cal F}((\partial_{v_{i}}G^{i})\varphi)+{\cal F}((\partial^{2}_{v_{i}v_{j}}G^{ij})\varphi)\Big{)}(s,\xi,\xi(t-s))ds
+0teλ(ts)((viHi)φ)(s,ξ,ξ(ts))𝑑βs.\displaystyle+\int_{0}^{t}e^{-\lambda(t-s)}{\cal F}((\partial_{v_{i}}H^{i})\varphi)(s,\xi,\xi(t-s))d\beta_{s}.

Note that since

(viGi)φ=vi(Giφ)Giviφ,\displaystyle(\partial_{v_{i}}G^{i})\varphi=\partial_{v_{i}}(G^{i}\varphi)-G^{i}\partial_{v_{i}}\varphi, (3.19)
(vivj2Gij)φ=vivj2(Gijφ)vi(Gijjφ)vj(Gijiφ)+Gijvivj2φ\displaystyle(\partial_{v_{i}v_{j}}^{2}G^{ij})\varphi=\partial_{v_{i}v_{j}}^{2}(G^{ij}\varphi)-\partial_{v_{i}}(G^{ij}\partial_{j}\varphi)-\partial_{v_{j}}(G^{ij}\partial_{i}\varphi)+G^{ij}\partial^{2}_{v_{i}v_{j}}\varphi (3.20)

forgetting the summation signs again, we may as well only consider terms of the form

ρφ^(t,ξ)=\displaystyle\widehat{\rho_{\varphi}}(t,\xi)= eλt(f0Ψ)(ξ,ξt)+λ0teλ(ts)(fΨ)(s,ξ,ξ(ts))𝑑s\displaystyle e^{-\lambda t}{\cal F}(f_{0}\Psi)(\xi,\xi t)+\lambda\int_{0}^{t}e^{-\lambda(t-s)}{\cal F}(f\Psi)(s,\xi,\xi(t-s))ds
+0teλ(ts)(vα(GβΨ))(s,ξ,ξ(ts))𝑑s\displaystyle+\int_{0}^{t}e^{-\lambda(t-s)}{\cal F}\Big{(}\partial_{v}^{\alpha}(G^{\beta}\Psi)\Big{)}(s,\xi,\xi(t-s))ds
+0teλ(ts)(vγ(HiΨ))(s,ξ,ξ(ts))𝑑βs\displaystyle+\int_{0}^{t}e^{-\lambda(t-s)}{\cal F}\Big{(}\partial_{v}^{\gamma}(H^{i}\Psi)\Big{)}(s,\xi,\xi(t-s))d\beta_{s} (3.21)

where ΨΨ(v)Cc(d)\Psi\equiv\Psi(v)\in C_{c}^{\infty}(\mathbb{R}^{d}) and α,β\alpha,\beta and γ\gamma are multi-indexes of order 0|α|20\leq|\alpha|\leq 2, 1|β|21\leq|\beta|\leq 2 and 0|γ|10\leq|\gamma|\leq 1. From (3.21) Itô’s isometry gives

𝔼0T|ρφ^(t,ξ)|2𝑑tJ1(ξ)+J2(ξ)+J3(ξ)+J4(ξ)\displaystyle\mathbb{E}\int_{0}^{T}|\widehat{\rho_{\varphi}}(t,\xi)|^{2}dt\lesssim J_{1}(\xi)+J_{2}(\xi)+J_{3}(\xi)+J_{4}(\xi) (3.22)

where

J1(ξ)=0Te2λt|(f0Ψ)(ξ,ξt)|2𝑑t,\displaystyle J_{1}(\xi)=\int_{0}^{T}e^{-2\lambda t}|{\cal F}(f_{0}\Psi)(\xi,\xi t)|^{2}dt,
J2(ξ)=λ2𝔼0T|0teλ(ts)(fΨ)(s,ξ,ξ(ts))𝑑s|2𝑑t,\displaystyle J_{2}(\xi)=\lambda^{2}\mathbb{E}\int_{0}^{T}\Big{|}\int_{0}^{t}e^{-\lambda(t-s)}{\cal F}(f\Psi)(s,\xi,\xi(t-s))ds\Big{|}^{2}dt,
J3(ξ)=𝔼0T|0teλ(ts)(vα(GβΨ))(s,ξ,ξ(ts))𝑑s|2𝑑t,\displaystyle J_{3}(\xi)=\mathbb{E}\int_{0}^{T}\Big{|}\int_{0}^{t}e^{-\lambda(t-s)}{\cal F}\Big{(}\partial_{v}^{\alpha}(G^{\beta}\Psi)\Big{)}(s,\xi,\xi(t-s))ds\Big{|}^{2}dt,
J4(ξ)=𝔼0T0te2λ(ts)|(vγ(HiΨ))(s,ξ,ξ(ts))|2𝑑s𝑑t.\displaystyle J_{4}(\xi)=\mathbb{E}\int_{0}^{T}\int_{0}^{t}e^{-2\lambda(t-s)}\Big{|}{\cal F}\Big{(}\partial_{v}^{\gamma}(H^{i}\Psi)\Big{)}(s,\xi,\xi(t-s))\Big{|}^{2}ds\,dt.

Using a trace lemma (see e.g [19], Theorem 2.7.2) we get, for some s>(d1)/2s>(d-1)/2,

J1(ξ)\displaystyle J_{1}(\xi) |(f0Ψ)(ξ,ξt)|2𝑑t=|ξ|1|(f0Ψ)(ξ,ξ|ξ|t)|2𝑑t\displaystyle\leq\int_{\mathbb{R}}\Big{|}{\cal F}(f_{0}\Psi)(\xi,\xi t)\Big{|}^{2}dt=|\xi|^{-1}\int_{\mathbb{R}}\Big{|}{\cal F}(f_{0}\Psi)(\xi,\frac{\xi}{|\xi|}t)\Big{|}^{2}dt
|ξ|1d|(IdΔw)s(f0Ψ)(ξ,w)|2𝑑w.\displaystyle\lesssim|\xi|^{-1}\int_{\mathbb{R}^{d}}\Big{|}(Id-\Delta_{w})^{s}{\cal F}(f_{0}\Psi)(\xi,w)\Big{|}^{2}dw.

It then follows by Plancherel’s identity, since Ψ\Psi is compactly supported, that

J1(ξ)|ξ|1d(1+|v|2)s|Ψ(v)|2|f^0(ξ,v)|2𝑑v|ξ|1d|f^0(ξ,v)|2𝑑v.\displaystyle J_{1}(\xi)\lesssim|\xi|^{-1}\int_{\mathbb{R}^{d}}(1+|v|^{2})^{s}|\Psi(v)|^{2}|\hat{f}_{0}(\xi,v)|^{2}dv\lesssim|\xi|^{-1}\int_{\mathbb{R}^{d}}|\hat{f}_{0}(\xi,v)|^{2}dv. (3.23)

For the second term J2(ξ)J_{2}(\xi), Jensen’s inequality gives

J2(ξ)\displaystyle J_{2}(\xi) λ𝔼0T0teλ(ts)|(fΨ)(s,ξ,ξ(ts))|2𝑑s𝑑t\displaystyle\leq\lambda\mathbb{E}\int_{0}^{T}\int_{0}^{t}e^{-\lambda(t-s)}\Big{|}{\cal F}(f\Psi)(s,\xi,\xi(t-s))\Big{|}^{2}dsdt
λ𝔼0TsT|(fΨ)(s,ξ,ξ(ts))|2𝑑t𝑑sλ𝔼0T(|(fΨ)(s,ξ,ξt)|2𝑑t)𝑑s\displaystyle\leq\lambda\mathbb{E}\int_{0}^{T}\int_{s}^{T}\Big{|}{\cal F}(f\Psi)(s,\xi,\xi(t-s))\Big{|}^{2}dtds\leq\lambda\mathbb{E}\int_{0}^{T}\Big{(}\int_{\mathbb{R}}\Big{|}{\cal F}(f\Psi)(s,\xi,\xi t)\Big{|}^{2}dt\Big{)}ds

and the same manipulation leads to

J2(ξ)λ|ξ|1𝔼0Td|f^(s,ξ,v)|2𝑑v𝑑s.\displaystyle J_{2}(\xi)\lesssim\lambda|\xi|^{-1}\mathbb{E}\int_{0}^{T}\int_{\mathbb{R}^{d}}|\hat{f}(s,\xi,v)|^{2}dvds. (3.24)

Similarly, for the third term J3(ξ)J_{3}(\xi), Jensen’s inequality gives

J3(ξ)\displaystyle J_{3}(\xi) λ1𝔼0T0teλ(ts)|(vα(GβΨ))(s,ξ,ξ(ts))|2𝑑s𝑑t\displaystyle\leq\lambda^{-1}\mathbb{E}\int_{0}^{T}\int_{0}^{t}e^{-\lambda(t-s)}\Big{|}{\cal F}(\partial^{\alpha}_{v}(G^{\beta}\Psi))(s,\xi,\xi(t-s))\Big{|}^{2}dsdt
λ1𝔼0T0teλ(ts)|ξ|4(ts)4|(GβΨ)(s,ξ,ξ(ts))|2𝑑s𝑑t\displaystyle\leq\lambda^{-1}\mathbb{E}\int_{0}^{T}\int_{0}^{t}e^{-\lambda(t-s)}|\xi|^{4}(t-s)^{4}\Big{|}{\cal F}(G^{\beta}\Psi)(s,\xi,\xi(t-s))\Big{|}^{2}dsdt
λ5|ξ|4𝔼0T(|(GβΨ)(s,ξ,ξt)|2𝑑t)𝑑s\displaystyle\lesssim\lambda^{-5}|\xi|^{4}\mathbb{E}\int_{0}^{T}\Big{(}\int_{\mathbb{R}}\Big{|}{\cal F}(G^{\beta}\Psi)(s,\xi,\xi t)\Big{|}^{2}dt\Big{)}ds

where we have used eλ(ts)(ts)4λ4e^{-\lambda(t-s)}(t-s)^{4}\lesssim\lambda^{-4}. The same manipulation hence leads to

J3(ξ)λ5|ξ|3𝔼0Td|Gβ^(s,ξ,v)|2𝑑v𝑑s.\displaystyle J_{3}(\xi)\lesssim\lambda^{-5}|\xi|^{3}\mathbb{E}\int_{0}^{T}\int_{\mathbb{R}^{d}}|\widehat{G^{\beta}}(s,\xi,v)|^{2}dvds. (3.25)

Finally, for the fourth term J4(ξ)J_{4}(\xi), we get

J4(ξ)\displaystyle J_{4}(\xi) 𝔼0T0te2λ(ts)|ξ|2(ts)2|(HiΨ)(s,ξ,ξ(ts))|2𝑑s𝑑t\displaystyle\leq\mathbb{E}\int_{0}^{T}\int_{0}^{t}e^{-2\lambda(t-s)}|\xi|^{2}(t-s)^{2}\Big{|}{\cal F}(H^{i}\Psi)(s,\xi,\xi(t-s))\Big{|}^{2}dsdt
λ2|ξ|2𝔼0T(|(HiΨ)(s,ξ,ξ(ts))|2𝑑t)𝑑s\displaystyle\lesssim\lambda^{-2}|\xi|^{2}\mathbb{E}\int_{0}^{T}\Big{(}\int_{\mathbb{R}}\Big{|}{\cal F}(H^{i}\Psi)(s,\xi,\xi(t-s))\Big{|}^{2}dt\Big{)}ds

and the same manipulation leads to

J4(ξ)λ2|ξ|𝔼0Td|Hi^(s,ξ,v)|2𝑑v𝑑s.\displaystyle J_{4}(\xi)\lesssim\lambda^{-2}|\xi|\mathbb{E}\int_{0}^{T}\int_{\mathbb{R}^{d}}|\widehat{H^{i}}(s,\xi,v)|^{2}dvds. (3.26)

Now fixing some η>0\eta>0, let us consider

𝔼0Td(1+|ξ|2η)|ρφ^(t,ξ)|2𝑑ξ𝑑t=I1+I2\displaystyle\mathbb{E}\int_{0}^{T}\int_{\mathbb{R}^{d}}(1+|\xi|^{2\eta})|\widehat{\rho_{\varphi}}(t,\xi)|^{2}d\xi dt=I_{1}+I_{2}

where

I1=𝔼0Td|ρφ^(t,ξ)|2𝑑ξ𝑑t+𝔼0T|ξ|1|ξ|2η|ρφ^(t,ξ)|2𝑑ξ𝑑t,\displaystyle I_{1}=\mathbb{E}\int_{0}^{T}\int_{\mathbb{R}^{d}}|\widehat{\rho_{\varphi}}(t,\xi)|^{2}d\xi dt+\mathbb{E}\int_{0}^{T}\int_{|\xi|\leq 1}|\xi|^{2\eta}|\widehat{\rho_{\varphi}}(t,\xi)|^{2}d\xi dt, (3.27)
I2=𝔼0T|ξ|>1|ξ|2η|ρφ^(t,ξ)|2𝑑ξ𝑑t.\displaystyle I_{2}=\mathbb{E}\int_{0}^{T}\int_{|\xi|>1}|\xi|^{2\eta}|\widehat{\rho_{\varphi}}(t,\xi)|^{2}d\xi dt. (3.28)

On one hand, making use of Plancherel’s identity, and since φ\varphi is compactly supported,

I12𝔼0Td|ρφ^(t,ξ)|2𝑑ξ𝑑t\displaystyle I_{1}\leq 2\mathbb{E}\int_{0}^{T}\int_{\mathbb{R}^{d}}|\widehat{\rho_{\varphi}}(t,\xi)|^{2}d\xi dt 𝔼0Td|ρφ(t,x)|2𝑑x𝑑t\displaystyle\lesssim\mathbb{E}\int_{0}^{T}\int_{\mathbb{R}^{d}}|\rho_{\varphi}(t,x)|^{2}dxdt
𝔼0Tdd|f(t,x,v)|2𝑑v𝑑x𝑑t1\displaystyle\lesssim\mathbb{E}\int_{0}^{T}\int_{\mathbb{R}^{d}}\int_{\mathbb{R}^{d}}|f(t,x,v)|^{2}dvdxdt\lesssim 1

thanks to Proposition 3.1, and the initial bound (3.14). On the other hand, from the bounds (3.23), (3.24), (3.25) and (3.26) we deduce (using Plancherel’s identity once again)

I2\displaystyle I_{2}\lesssim sup|ξ>1[|ξ|2η1+λ(ξ)|ξ|2η1+λ(ξ)5|ξ|2η+3+λ(ξ)2|ξ|2η+1]\displaystyle\sup_{|\xi>1}\Big{[}|\xi|^{2\eta-1}+\lambda(\xi)|\xi|^{2\eta-1}+\lambda(\xi)^{-5}|\xi|^{2\eta+3}+\lambda(\xi)^{-2}|\xi|^{2\eta+1}\Big{]}
×[z2d|f0|2𝑑z+𝔼0Tz2d(|f|2+|Gβ|2+|Hi|2)𝑑z𝑑t].\displaystyle\times\Big{[}\int_{z\in\mathbb{R}^{2d}}|f_{0}|^{2}dz+\mathbb{E}\int_{0}^{T}\int_{z\in\mathbb{R}^{2d}}\Big{(}|f|^{2}+|G^{\beta}|^{2}+|H^{i}|^{2}\Big{)}dzdt\Big{]}. (3.29)

Expressions (3.17) and the sublinearity estimate (2.4) immediately give

|Gi(t,z)|2(1+|v|2+|v|2f(t,z)𝑑z+|uR[f(t)](x)|2)|f(t,z)|2,\displaystyle|G^{i}(t,z)|^{2}\lesssim\Big{(}1+|v|^{2}+\int|v^{\prime}|^{2}f(t,z^{\prime})dz^{\prime}+\Big{|}u_{R}[f(t)](x)\Big{|}^{2}\Big{)}|f(t,z)|^{2}, (3.30)
|Gij(t,z)|2(1+|v|4+|v|4f(t,z)𝑑z)|f(t,z)|2,\displaystyle|G^{ij}(t,z)|^{2}\lesssim\Big{(}1+|v|^{4}+\int|v^{\prime}|^{4}f(t,z^{\prime})dz^{\prime}\Big{)}|f(t,z)|^{2}, (3.31)
|Hi(t,z)|2(1+|v|2+|v|2f(t,z)𝑑z)|f(t,z)|2,\displaystyle|H^{i}(t,z)|^{2}\lesssim\Big{(}1+|v|^{2}+\int|v^{\prime}|^{2}f(t,z^{\prime})dz^{\prime}\Big{)}|f(t,z)|^{2}, (3.32)

from which we easily deduce (using the same method as in the proof of Proposition 3.3 for the term involving uR[f(t)]u_{R}[f(t)]), for θ(0,1)\theta\in(0,1),

z2d|Gi|2𝑑zz2d(1+|v|21θ)f(t)𝑑z+z2d|f(t)|1+1θ𝑑z,\displaystyle\int_{z\in\mathbb{R}^{2d}}|G^{i}|^{2}dz\lesssim\int_{z\in\mathbb{R}^{2d}}(1+|v|^{\frac{2}{1-\theta}})f(t)dz+\int_{z\in\mathbb{R}^{2d}}|f(t)|^{1+\frac{1}{\theta}}dz,
z2d|Gij|2𝑑zz2d(1+|v|41θ)f(t)𝑑z+z2d|f(t)|1+1θ𝑑z,\displaystyle\int_{z\in\mathbb{R}^{2d}}|G^{ij}|^{2}dz\lesssim\int_{z\in\mathbb{R}^{2d}}(1+|v|^{\frac{4}{1-\theta}})f(t)dz+\int_{z\in\mathbb{R}^{2d}}|f(t)|^{1+\frac{1}{\theta}}dz,
z2d|Hi|2𝑑zz2d(1+|v|21θ)f(t)𝑑z+z2d|f(t)|1+1θ𝑑z,\displaystyle\int_{z\in\mathbb{R}^{2d}}|H^{i}|^{2}dz\lesssim\int_{z\in\mathbb{R}^{2d}}(1+|v|^{\frac{2}{1-\theta}})f(t)dz+\int_{z\in\mathbb{R}^{2d}}|f(t)|^{1+\frac{1}{\theta}}dz,

so that, thanks again to Proposition 3.1, Proposition 3.3 and the initial bound (3.14), (3.29) yields

I2sup|ξ>1[|ξ|2η1+λ(ξ)|ξ|2η1+λ(ξ)5|ξ|2η+3+λ(ξ)2|ξ|2η+1].\displaystyle I_{2}\lesssim\sup_{|\xi>1}\Big{[}|\xi|^{2\eta-1}+\lambda(\xi)|\xi|^{2\eta-1}+\lambda(\xi)^{-5}|\xi|^{2\eta+3}+\lambda(\xi)^{-2}|\xi|^{2\eta+1}\Big{]}.

Considering λ(ξ)=|ξ|r\lambda(\xi)=|\xi|^{r}, this supremum is bounded under the requirements

2η10,2η1+r0,2η+35r0,2η+12r0,\displaystyle 2\eta-1\leq 0,\hskip 14.22636pt2\eta-1+r\leq 0,\hskip 14.22636pt2\eta+3-5r\leq 0,\hskip 14.22636pt2\eta+1-2r\leq 0,

which can be met as soon as η1/6\eta\leq 1/6. For such η\eta, we have shown that

𝔼[ρφLt2Hxη2]𝔼0Td(1+|ξ|2η)|ρφ^(t,ξ)|2𝑑ξ𝑑t1.\displaystyle\mathbb{E}\Big{[}\|\rho_{\varphi}\|_{L^{2}_{t}H^{\eta}_{x}}^{2}\Big{]}\lesssim\mathbb{E}\int_{0}^{T}\int_{\mathbb{R}^{d}}(1+|\xi|^{2\eta})|\widehat{\rho_{\varphi}}(t,\xi)|^{2}d\xi dt\lesssim 1.

which concludes the proof in the case of a regular initial data f0Cc2(2d)f_{0}\in C_{c}^{2}(\mathbb{R}^{2d}). We may extend this last inequality to general initial data similarly to the proof of Proposition 3.1. ∎

3.3 Tightness

Given an increasing weight function, say

W(x,v)=1+|x|+|v|,\displaystyle W(x,v)=1+|x|+|v|,

let us introduce the weighted Sobolev space

HW2(2d)={ΨD(2d),ΨHW22:=max|β|2|zβΨ|2(z)W(z)𝑑z<}\displaystyle H^{2}_{W}(\mathbb{R}^{2d})=\left\{\Psi\in D^{\prime}(\mathbb{R}^{2d}),\;\|\Psi\|_{H^{2}_{W}}^{2}:=\max_{|\beta|\leq 2}\int|\partial_{z}^{\beta}\Psi|^{2}(z)W(z)dz<\infty\right\} (3.33)

and the dual space

HW12(2d)=(HW12(2d)) with hHW12=sup{h,Ψ,ΨHW2=1}.\displaystyle H^{-2}_{W^{-1}}(\mathbb{R}^{2d})=\Big{(}H^{2}_{W^{-1}}(\mathbb{R}^{2d})\Big{)}^{\prime}\text{~with }\|h\|_{H^{-2}_{W^{-1}}}=\sup\left\{\langle h,\Psi\rangle,\;\|\Psi\|_{H^{2}_{W}}=1\right\}. (3.34)

We may also define the intermediate Sobolev spaces HWσ(2d)H^{\sigma}_{W}(\mathbb{R}^{2d}), HWσ(2d)H^{-\sigma}_{W}(\mathbb{R}^{2d}) for non-integer 0<σ<20<\sigma<2. Note that ΨH2ΨHW2\|\Psi\|_{H^{2}}\leq\|\Psi\|_{H^{2}_{W}} so that hHW12hH2\|h\|_{H^{-2}_{W^{-1}}}\leq\|h\|_{H^{-2}}.

Proposition 3.5.

Let us assume that the initial data f0f_{0} satisfies, for some δ>1\delta>1 and θ(0,1)\theta\in(0,1),

z2d|f0(z)|p𝑑z+z2d(|x|δ+|v|k)f0(z)𝑑z< with p=1+1θ,k>41θ.\displaystyle\int_{z\in\mathbb{R}^{2d}}|f_{0}(z)|^{p}dz+\int_{z\in\mathbb{R}^{2d}}(|x|^{\delta}+|v|^{k})f_{0}(z)dz<\infty\;\text{ with }\;p=1+\frac{1}{\theta},\hskip 11.38109ptk>\frac{4}{1-\theta}. (3.35)

Then for all σ>0\sigma>0, the family of random variables (fR)R>0(f^{R})_{R>0} is tight in C([0,T];HW1σ(2d))C([0,T];H^{-\sigma}_{W^{-1}}(\mathbb{R}^{2d})) .

Proof.

Without loss of generality, let σ(0,2)\sigma\in(0,2). For some α(0,1/2)\alpha\in(0,1/2) and M>0M>0, let us introduce the set

KM={fC([0,T];HW1σ(2d))|fLtLz2M,fCtαHW12M},K_{M}=\left\{f\in C([0,T];H^{-\sigma}_{W^{-1}}(\mathbb{R}^{2d}))\;\Big{|}\;\|f\|_{L^{\infty}_{t}L^{2}_{z}}\leq M,\;\|f\|_{C^{\alpha}_{t}H_{W^{-1}}^{-2}}\leq M\right\},

where CtαHW2\|\cdot\|_{C^{\alpha}_{t}H^{-2}_{W}} denotes the α\alpha-Hölder semi-norm

fCtαHW12:=suptsft)f(s)HW12|ts|αfCtαH2.\displaystyle\|f\|_{C^{\alpha}_{t}H^{-2}_{W^{-1}}}:=\sup_{t\neq s}\frac{\|ft)-f(s)\|_{H^{-2}_{W^{-1}}}}{|t-s|^{\alpha}}\leq\|f\|_{C^{\alpha}_{t}H^{-2}}.

Since the embedding HWσ(2d)L2(2d)H^{\sigma}_{W}(\mathbb{R}^{2d})\subset L^{2}(\mathbb{R}^{2d}) is compact, the dual embedding L2(2d)HW1σ(2d)L^{2}(\mathbb{R}^{2d})\subset H^{-\sigma}_{W^{-1}}(\mathbb{R}^{2d}) is compact. Additionally, for fKMf\in K_{M}, an interpolation inequality (in weighted Sobolev spaces) yields, for some τ(0,1)\tau\in(0,1),

fCtαHW1σfLtLW12τfCtαHW121τM2.\|f\|_{C^{\alpha}_{t}H^{-\sigma}_{W^{-1}}}\lesssim\|f\|_{L^{\infty}_{t}L^{2}_{W^{-1}}}^{\tau}\|f\|_{C^{\alpha}_{t}H^{-2}_{W^{-1}}}^{1-\tau}\lesssim M^{2}.

Consequently, Arzelà-Ascoli’s theorem guarantees that KMK_{M} is a relatively compact subset of the separable, complete space C([0,T];HW1σ(2d))C([0,T];H^{-\sigma}_{W^{-1}}(\mathbb{R}^{2d})). Markov’s inequality gives

[fRKM]M2𝔼[supt[0,T]fR(t)L22]+Mγ𝔼[fRCtαH2γ]).\mathbb{P}\Big{[}f^{R}\notin K_{M}\Big{]}\leq M^{-2}\mathbb{E}\Big{[}\sup_{t\in[0,T]}\|f^{R}(t)\|^{2}_{L^{2}}\Big{]}+M^{-\gamma}\mathbb{E}\Big{[}\|f^{R}\|^{\gamma}_{C^{\alpha}_{t}H^{-2}}\Big{]}\Big{)}.

for γ>0\gamma>0. The first term is bounded uniformly in RR thanks to Proposition 3.1. The bound on the second term results directly from Kolmogorov’s continuity theorem and the following lemma.

Lemma 3.1.

For some q>1q>1, for all t,s[0,T]t,s\in[0,T],

𝔼[fR(t)fR(s)H22q]|ts|q.\displaystyle\mathbb{E}\Big{[}\|f^{R}(t)-f^{R}(s)\|_{H^{-2}}^{2q}\Big{]}\lesssim|t-s|^{q}.

The constant involved in \lesssim depends on qq, f0f_{0} and TT only.

We now prove this lemma: let 0stT0\leq s\leq t\leq T. Let us simply denote f=fRf=f^{R} and note that

𝔼[f(t)f(s)H22q]=𝔼[|dd(1+|ξ|4+|ζ|4)1|f(t,ξ,ζ)f(s,ξ,ζ)|2𝑑ξ𝑑ζ|q].\displaystyle\mathbb{E}\Big{[}\|f(t)-f(s)\|_{H^{-2}}^{2q}\Big{]}=\mathbb{E}\Big{[}\Big{|}\int_{\mathbb{R}^{d}}\int_{\mathbb{R}^{d}}(1+|\xi|^{4}+|\zeta|^{4})^{-1}\Big{|}{\cal F}f(t,\xi,\zeta)-{\cal F}f(s,\xi,\zeta)\Big{|}^{2}d\xi d\zeta\Big{|}^{q}\Big{]}.

As in the proof of Proposition (3.4), we may apply the (x,v)(x,v)-Fourier transform f(t,ξ,ζ){\cal F}f(t,\xi,\zeta) to equation (3.16) to get (forgetting the summation signs)

f(t)f(s)=st(iξ(vf)+iζi(Gi)ζiζj(Gij))𝑑σ+istζi(Hi)𝑑βσ.\displaystyle{\cal F}f(t)-{\cal F}f(s)=\int_{s}^{t}\Big{(}-i\xi\cdot{\cal F}(vf)+i\zeta^{i}{\cal F}(G^{i})-\zeta^{i}\zeta^{j}{\cal F}(G^{ij})\Big{)}d\sigma+i\int_{s}^{t}\zeta^{i}{\cal F}(H^{i})d\beta_{\sigma}.

Itô’s formula results in

|f(t)f(s)|2=\displaystyle\Big{|}{\cal F}f(t)-{\cal F}f(s)\Big{|}^{2}= 2stf(σ)t(s)¯(iξ(vf)+iζi(Gi)ζiζj(Gij))𝑑σ\displaystyle 2\int_{s}^{t}\overline{{\cal F}f(\sigma)-{\cal F}t(s)}\Big{(}-i\xi\cdot{\cal F}(vf)+i\zeta^{i}{\cal F}(G^{i})-\zeta^{i}\zeta^{j}{\cal F}(G^{ij})\Big{)}d\sigma
+st|ζi|2|(Hi)|2𝑑σ+2istf(σ)t(s)¯ζi(Hi)𝑑βσ\displaystyle+\int_{s}^{t}|\zeta^{i}|^{2}|{\cal F}(H^{i})|^{2}d\sigma+2i\int_{s}^{t}\overline{{\cal F}f(\sigma)-{\cal F}t(s)}\zeta^{i}{\cal F}(H^{i})d\beta_{\sigma}

so that, integrating against (1+|ξ|4+|ζ|4)1dξdζ(1+|\xi|^{4}+|\zeta|^{4})^{-1}d\xi d\zeta, we get

f(t)f(s)H22\displaystyle\|f(t)-f(s)\|^{2}_{H^{-2}}\lesssim stf(σ)f(s)H22𝑑σ+Dt+Mt\displaystyle\int_{s}^{t}\|f(\sigma)-f(s)\|^{2}_{H^{-2}}d\sigma+D_{t}+M_{t} (3.36)

where

Dt=st2d(|(vf)|2+|(Gi)|2+|(Gij)|2+|(Hi)|2)𝑑ξ𝑑ζ𝑑σ,\displaystyle D_{t}=\int_{s}^{t}\int_{\mathbb{R}^{2d}}\Big{(}|{\cal F}(vf)|^{2}+|{\cal F}(G^{i})|^{2}+|{\cal F}(G^{ij})|^{2}+|{\cal F}(H^{i})|^{2}\Big{)}d\xi d\zeta d\sigma,
Mt=2ist(2d(1+|ξ|4+|ζ|4)1f(σ)t(s)¯ζi(Hi)𝑑ξ𝑑ζ)𝑑βσ.\displaystyle M_{t}=2i\int_{s}^{t}\Big{(}\int_{\mathbb{R}^{2d}}(1+|\xi|^{4}+|\zeta|^{4})^{-1}\overline{{\cal F}f(\sigma)-{\cal F}t(s)}\zeta^{i}{\cal F}(H^{i})d\xi d\zeta\Big{)}d\beta_{\sigma}.

From (3.36) we derive

𝔼f(t)f(s)H22qst𝔼f(σ)f(s)H22q𝑑σ+𝔼[|Dt|q]+𝔼[|Mt|q].\displaystyle\mathbb{E}\|f(t)-f(s)\|^{2q}_{H^{-2}}\lesssim\int_{s}^{t}\mathbb{E}\|f(\sigma)-f(s)\|^{2q}_{H^{-2}}d\sigma+\mathbb{E}\Big{[}|D_{t}|^{q}\Big{]}+\mathbb{E}\Big{[}|M_{t}|^{q}\Big{]}. (3.37)

Plancherel’s identity gives

|Dt|q\displaystyle|D_{t}|^{q} =|st(vfL22+GiL22+GijL22+HiL22)𝑑σ|q\displaystyle=\Big{|}\int_{s}^{t}\Big{(}\|vf\|_{L^{2}}^{2}+\|G^{i}\|^{2}_{L^{2}}+\|G^{ij}\|^{2}_{L^{2}}+\|H^{i}\|^{2}_{L^{2}}\Big{)}d\sigma\Big{|}^{q}
|ts|q1st(vfL22q+GiL22q+GijL22q+HiL22q)𝑑σ\displaystyle\lesssim|t-s|^{q-1}\int_{s}^{t}\Big{(}\|vf\|_{L^{2}}^{2q}+\|G^{i}\|^{2q}_{L^{2}}+\|G^{ij}\|^{2q}_{L^{2}}+\|H^{i}\|^{2q}_{L^{2}}\Big{)}d\sigma

so that

𝔼[|Dt|q]|ts|q𝔼[supt[0,T](vf(t)L22q+Gi(t)L22q+Gij(t)L22q+Hi(t)L22q)]\displaystyle\mathbb{E}\Big{[}|D_{t}|^{q}\Big{]}\lesssim|t-s|^{q}\mathbb{E}\Big{[}\sup_{t\in[0,T]}\Big{(}\|vf(t)\|_{L^{2}}^{2q}+\|G^{i}(t)\|^{2q}_{L^{2}}+\|G^{ij}(t)\|^{2q}_{L^{2}}+\|H^{i}(t)\|^{2q}_{L^{2}}\Big{)}\Big{]}

Considering (3.30), (3.31), (3.32), we see that we essentially need a bound (uniform in RR) on

𝔼[supt[0,T]|2d(1+|v|4)|f(t,z)|2𝑑z|q].\displaystyle\mathbb{E}\Big{[}\sup_{t\in[0,T]}\Big{|}\int_{\mathbb{R}^{2d}}(1+|v|^{4})|f(t,z)|^{2}dz\Big{|}^{q}\Big{]}. (3.38)

This is possible since, for any m>2dm>2d,

|2d(1+|v|4)|f(t,z)|2𝑑z|q\displaystyle\Big{|}\int_{\mathbb{R}^{2d}}(1+|v|^{4})|f(t,z)|^{2}dz\Big{|}^{q} 2d(1+|v|4q)(1+|z|)m(q1)|f|2q𝑑z\displaystyle\lesssim\int_{\mathbb{R}^{2d}}(1+|v|^{4q})(1+|z|)^{m(q-1)}|f|^{2q}dz
2d(1+|v|4q+m(q1)+|x|m(q1))|f|2q𝑑z\displaystyle\lesssim\int_{\mathbb{R}^{2d}}(1+|v|^{4q+m(q-1)}+|x|^{m(q-1)})|f|^{2q}dz

so that, for τ(0,1)\tau\in(0,1),

|2d(1+|v|4)|f(t,z)|2𝑑z|q\displaystyle\Big{|}\int_{\mathbb{R}^{2d}}(1+|v|^{4})|f(t,z)|^{2}dz\Big{|}^{q} 2d(1+|x|m(q1)1τ+|v|4q+m(q1)1τ)f𝑑z+2d|f|1+2q1τ𝑑z\displaystyle\lesssim\int_{\mathbb{R}^{2d}}(1+|x|^{\frac{m(q-1)}{1-\tau}}+|v|^{\frac{4q+m(q-1)}{1-\tau}})fdz+\int_{\mathbb{R}^{2d}}|f|^{1+\frac{2q-1}{\tau}}dz
2d(1+|x|δ+|v|k)f𝑑z+2d|f|p𝑑z\displaystyle\lesssim\int_{\mathbb{R}^{2d}}(1+|x|^{\delta}+|v|^{k})fdz+\int_{\mathbb{R}^{2d}}|f|^{p}dz

whenever, recalling (3.35), for some γ>0\gamma>0,

2q1τ=1θ,m(q1)1τδ,4q+m(q1)1τk:=4+γ1θ.\displaystyle\frac{2q-1}{\tau}=\frac{1}{\theta},\hskip 28.45274pt\frac{m(q-1)}{1-\tau}\leq\delta,\hskip 28.45274pt\frac{4q+m(q-1)}{1-\tau}\leq k:=\frac{4+\gamma}{1-\theta}.

These requirements can be met for some q=q(γ)>1q=q(\gamma)>1 close enough to 11 (and τ\tau close to θ\theta). As for the martingale term, Burkholder-Davis-Gundy’s inequality gives

𝔼[|Mt|q]\displaystyle\mathbb{E}\Big{[}|M_{t}|^{q}\Big{]} 𝔼[|st(2d(1+|ξ|4+|ζ|4)1|f(σ)t(s)||ζ||(Hi)|𝑑ξ𝑑ζ)2𝑑σ|q/2]\displaystyle\lesssim\mathbb{E}\Big{[}\Big{|}\int_{s}^{t}\Big{(}\int_{\mathbb{R}^{2d}}(1+|\xi|^{4}+|\zeta|^{4})^{-1}|{\cal F}f(\sigma)-{\cal F}t(s)||\zeta||{\cal F}(H^{i})|d\xi d\zeta\Big{)}^{2}d\sigma\Big{|}^{q/2}\Big{]}
𝔼[|stf(σ)f(s)H22HiL22𝑑σ|q/2]\displaystyle\lesssim\mathbb{E}\Big{[}\Big{|}\int_{s}^{t}\|f(\sigma)-f(s)\|_{H^{-2}}^{2}\|H^{i}\|_{L^{2}}^{2}d\sigma\Big{|}^{q/2}\Big{]}
𝔼[|supt[0,T]Hi(t)L22|q/2|stf(σ)f(s)H22𝑑σ|q/2]\displaystyle\lesssim\mathbb{E}\Big{[}\Big{|}\sup_{t\in[0,T]}\|H^{i}(t)\|^{2}_{L^{2}}\Big{|}^{q/2}\Big{|}\int_{s}^{t}\|f(\sigma)-f(s)\|_{H^{-2}}^{2}d\sigma\Big{|}^{q/2}\Big{]}
𝔼[supt[0,T]Hi(t)L22q]+st𝔼f(σ)f(s)H22q𝑑σ.\displaystyle\lesssim\mathbb{E}\Big{[}\sup_{t\in[0,T]}\|H^{i}(t)\|^{2q}_{L^{2}}\Big{]}+\int_{s}^{t}\mathbb{E}\|f(\sigma)-f(s)\|_{H^{-2}}^{2q}d\sigma. (3.39)

Considering (3.32), the first term in (3.39) is again controlled by the bound on (3.38). We may then come back to (3.37) and use Grönwall’s lemma to conclude.

Proposition 3.6.

Let us assume that the initial data f0f_{0} satisfies (3.35).
Then, for all φCc(d)\varphi\in C_{c}^{\infty}(\mathbb{R}^{d}), the family of random variables (ρφR)R>0(\rho^{R}_{\varphi})_{R>0} is tight in L2([0,T];L2(d))L^{2}([0,T];L^{2}(\mathbb{R}^{d})).

Proof.

Let us introduce the weight function

V(x)=1+|x|V(x)=1+|x|

and the associated weighted spaces HV2(d)H^{2}_{V}(\mathbb{R}^{d}) and HV12(d)H^{-2}_{V^{-1}}(\mathbb{R}^{d}) as in (3.33) and (3.34). One could prove the following lemma as previously.

Lemma 3.2.

For some q>1q>1, for all t,s[0,T]t,s\in[0,T],

𝔼[ρφR(t)ρφR(s)H22q]|ts|q.\displaystyle\mathbb{E}\Big{[}\|\rho^{R}_{\varphi}(t)-\rho^{R}_{\varphi}(s)\|_{H^{-2}}^{2q}\Big{]}\lesssim|t-s|^{q}.

The constant involved in \lesssim depends on qq, f0f_{0}, φ\varphi and TT only.

We may now fix some M>0M>0 and naturally introduce the set

KM={ρφ:=φ(v)f𝑑v|fM},\displaystyle K_{M}=\left\{\rho_{\varphi}:=\int\varphi(v)fdv\;|\;f\in{\cal F}_{M}\right\},

where M{\cal F}_{M} denotes the set of functions ff(t,x,v)f\equiv f(t,x,v) satisfying, for η=1/6\eta=1/6 and some α(0,1/2)\alpha\in(0,1/2),

supt[0,T]f(t)LzppM,\displaystyle\sup_{t\in[0,T]}\|f(t)\|_{L^{p}_{z}}^{p}\leq M, (3.40)
supt[0,T](1+|x|δ+|v|k)f(t)𝑑zM,\displaystyle\sup_{t\in[0,T]}\int(1+|x|^{\delta}+|v|^{k})f(t)dz\leq M, (3.41)
ρφLt2Hxη2M,\displaystyle\|\rho_{\varphi}\|_{L^{2}_{t}H^{\eta}_{x}}^{2}\leq M, (3.42)
ρφCtαH2M.\displaystyle\|\rho_{\varphi}\|_{C^{\alpha}_{t}H^{-2}}\leq M.

Markov’s inequality, gives, for some γ>0\gamma>0,

[ρφRKM]\displaystyle\mathbb{P}\Big{[}\rho^{R}_{\varphi}\notin K_{M}\Big{]} M1𝔼[supt[0,T]fR(t)Lzpp]+M1𝔼[supt[0,T](1+|x|δ+|v|k)fR(t)𝑑z]\displaystyle\leq M^{-1}\mathbb{E}\Big{[}\sup_{t\in[0,T]}\|f^{R}(t)\|_{L^{p}_{z}}^{p}\Big{]}+M^{-1}\mathbb{E}\Big{[}\sup_{t\in[0,T]}\int(1+|x|^{\delta}+|v|^{k})f^{R}(t)dz\Big{]}
+M1𝔼[ρφRLt2Hxη2]+Mγ𝔼[ρφRCtαH2γ]\displaystyle\;\;+M^{-1}\mathbb{E}\Big{[}\|{\rho^{R}_{\varphi}}\|_{L^{2}_{t}H^{\eta}_{x}}^{2}\Big{]}+M^{-\gamma}\mathbb{E}\Big{[}\|{\rho^{R}_{\varphi}}\|_{C^{\alpha}_{t}H^{-2}}^{\gamma}\Big{]}

which tends to zero uniformly in R>0R>0 as MM goes to infinity, thanks to Proposition 3.1, 3.3, 3.4 and Lemma 3.2. It only remains to prove that KMK_{M} is a relatively compact subset of L2([0,T];L2(d))L^{2}([0,T];L^{2}(\mathbb{R}^{d})). Let us introduce a sequence (ρφn)n1(\rho_{\varphi}^{n})_{n\geq 1} in KMK_{M}.

First, let us show that (ρφn)n(\rho_{\varphi}^{n})_{n} is compact locally in space, that is in L2([0,T];L2(B(0,r)))L^{2}\Big{(}[0,T];L^{2}(B(0,r))\Big{)} for any r>0r>0. Since ρφnLtLx1φL\|\rho^{n}_{\varphi}\|_{L^{\infty}_{t}L^{1}_{x}}\leq\|\varphi\|_{L^{\infty}}, we deduce from (3.40) that

ρφnLtLx2M.\|\rho^{n}_{\varphi}\|_{L^{\infty}_{t}L^{2}_{x}}\lesssim M.

Similarly to the proof of Proposition 3.5, we may then use Arzelà-Ascoli’s theorem to deduce that (ρφn)n(\rho_{\varphi}^{n})_{n} converges in C([0,T];HV12(d))C([0,T];H^{-2}_{V^{-1}}(\mathbb{R}^{d})) up to some subsequence (which we omit for clarity). An interpolation inequality (in weighted Sobolev spaces: HV1ηLV12HV12H^{\eta}_{V^{-1}}\subset L^{2}_{V^{-1}}\subset H^{-2}_{V^{-1}}) yields

ρL2(B(0,r))ρLV12ρHV12τρHV1η1τρHV12τρHη1τ\displaystyle\|\rho\|_{L^{2}(B(0,r))}\lesssim\|\rho\|_{L^{2}_{V^{-1}}}\lesssim\|\rho\|_{H^{-2}_{V^{-1}}}^{\tau}\|\rho\|_{H^{\eta}_{V^{-1}}}^{1-\tau}\lesssim\|\rho\|_{H^{-2}_{V^{-1}}}^{\tau}\|\rho\|_{H^{\eta}}^{1-\tau}

where τ=ηη+2(0,1)\tau=\frac{\eta}{\eta+2}\in(0,1). It follows that

0Tρφn(t)ρφm(t)L2(B(0,r))2𝑑t\displaystyle\int_{0}^{T}\|\rho_{\varphi}^{n}(t)-\rho_{\varphi}^{m}(t)\|_{L^{2}(B(0,r))}^{2}dt 0Tρφn(t)ρφm(t)HV122τ×ρφn(t)ρφm(t)Hη2(1τ)𝑑t\displaystyle\leq\int_{0}^{T}\|\rho_{\varphi}^{n}(t)-\rho_{\varphi}^{m}(t)\|_{H^{-2}_{V^{-1}}}^{2\tau}\times\|\rho_{\varphi}^{n}(t)-\rho_{\varphi}^{m}(t)\|_{H^{\eta}}^{2(1-\tau)}dt
(0Tρφn(t)ρφm(t)HV122𝑑t)τ(0Tρφn(t)ρφm(t)Hη2𝑑t)1τ\displaystyle\leq\Big{(}\int_{0}^{T}\|\rho_{\varphi}^{n}(t)-\rho_{\varphi}^{m}(t)\|_{H^{-2}_{V^{-1}}}^{2}dt\Big{)}^{\tau}\Big{(}\int_{0}^{T}\|\rho_{\varphi}^{n}(t)-\rho_{\varphi}^{m}(t)\|_{H^{\eta}}^{2}dt\Big{)}^{1-\tau}
TτρφnρφmC([0,T];HV12)2λ×ρφnρφmLt2Hη2(1τ).\displaystyle\leq T^{\tau}\|\rho_{\varphi}^{n}-\rho_{\varphi}^{m}\|_{C([0,T];H^{-2}_{V^{-1}})}^{2\lambda}\times\|\rho_{\varphi}^{n}-\rho_{\varphi}^{m}\|^{2(1-\tau)}_{L^{2}_{t}H^{\eta}}.

Thanks to (3.42), we deduce that

ρφnρφmL2([0,T];L2(B(0,r))0 as n,m.\displaystyle\|\rho_{\varphi}^{n}-\rho_{\varphi}^{m}\|_{L^{2}([0,T];L^{2}(B(0,r))}\to 0\;\text{ as }n,m\to\infty.

To derive compactness globally in space, that is in L2([0,T];L2(d))L^{2}([0,T];L^{2}(\mathbb{R}^{d})), it is enough to establish a uniform integrability estimate of the form

supn1[0T|x|r|ρφn(t,x)|2𝑑x𝑑t]0 as r.\sup_{n\geq 1}\Big{[}\int_{0}^{T}\int_{|x|\geq r}|\rho_{\varphi}^{n}(t,x)|^{2}dxdt\Big{]}\to 0\;\text{ as }r\to\infty.

To this intent, since φ\varphi is compactly supported, we may simply write

|x|r|ρφn(x)|2𝑑xd|x|r|fn(x,v)|2𝑑x𝑑v\displaystyle\int_{|x|\geq r}|\rho^{n}_{\varphi}(x)|^{2}dx\lesssim\int_{\mathbb{R}^{d}}\int_{|x|\geq r}|f^{n}(x,v)|^{2}dxdv rγdd(1+|x|γ)|fn(x,v)|2𝑑x𝑑v\displaystyle\leq r^{-\gamma}\int_{\mathbb{R}^{d}}\int_{\mathbb{R}^{d}}(1+|x|^{\gamma})|f^{n}(x,v)|^{2}dxdv
rγ(2d(1+|x|δ)fn(z)𝑑z+2d|fn(z)|p𝑑z)\displaystyle\lesssim r^{-\gamma}\Big{(}\int_{\mathbb{R}^{2d}}(1+|x|^{\delta})f^{n}(z)dz+\int_{\mathbb{R}^{2d}}|f^{n}(z)|^{p}dz\Big{)}

where γ=δ(1θ)>0\gamma=\delta(1-\theta)>0, according to (3.35), and then use the bounds (3.40) and (3.41) to conclude. ∎

3.4 Convergence of the martingale problem

Let us introduce a sequence RnR_{n}\to\infty and a countable subset 𝒟{\cal D} of Cc(d)C^{\infty}_{c}(\mathbb{R}^{d}), which we assume to contain the truncation functions

{θRn,n1}𝒟Cc(d).\displaystyle\left\{\theta_{R_{n}},n\geq 1\right\}\subset{\cal D}\subset C_{c}^{\infty}(\mathbb{R}^{d}). (3.43)

Recall that the function θR(v)\theta_{R}(v) has been introduced in (2.1). Since 𝒟{\cal D} is countable, it follows from Proposition 3.5, Proposition 3.6 and Tykhonov’s theorem that the family of random variables (fRn,(ρφRn)φ𝒟)n1(f^{R_{n}},(\rho^{R_{n}}_{\varphi})_{\varphi\in{\cal D}})_{n\geq 1} is tight in the space

C([0,T];HW1σ(2d))×(L2([0,T];L2(d))𝒟C([0,T];H^{-\sigma}_{W^{-1}}(\mathbb{R}^{2d}))\times\Big{(}L^{2}([0,T];L^{2}(\mathbb{R}^{d})\Big{)}^{\cal D}

for σ>0\sigma>0. Using Skorokhod’s representation theorem, up to a subsequence of (Rn)n(R_{n})_{n} which we omit for simplicity, we may introduce random variables f¯n,ρ¯φn,f¯,ρ¯φ\overline{f}^{n},\overline{\rho}_{\varphi}^{n},\overline{f},\overline{\rho}_{\varphi} defined on some other probability space (Ω¯,¯,¯)(\overline{\Omega},\overline{\cal F},\overline{\mathbb{P}}) such that, for all n1n\geq 1,

(f¯n,(ρ¯φn)φ𝒟)(fRn,(ρφRn)φ𝒟) in law, in C([0,T];HW1σ(2d))×(L2([0,T];L2(d)))𝒟,\displaystyle(\overline{f}^{n},(\overline{\rho}^{n}_{\varphi})_{\varphi\in{\cal D}})\sim(f^{R_{n}},(\rho^{R_{n}}_{\varphi})_{\varphi\in{\cal D}})\text{ in law, in }C([0,T];H^{-\sigma}_{W^{-1}}(\mathbb{R}^{2d}))\times\Big{(}L^{2}([0,T];L^{2}(\mathbb{R}^{d}))\Big{)}^{\cal D},

and the following convergences hold ¯\overline{\mathbb{P}}-almost surely:

f¯nf¯ in C([0,T];HW1σ(2d)) a.s,\displaystyle\overline{f}^{n}\to\overline{f}\text{ in }C([0,T];H^{-\sigma}_{W^{-1}}(\mathbb{R}^{2d}))\text{ a.s},
φ𝒟,ρ¯φnρ¯φ in L2([0,T];L2(d)) a.s.\displaystyle\forall\varphi\in{\cal D},\;\;\overline{\rho}^{n}_{\varphi}\to\overline{\rho}_{\varphi}\text{ in }L^{2}([0,T];L^{2}(\mathbb{R}^{d}))\text{~a.s}.

More precisely, since Proposition 3.1 and Proposition 3.4 provide the bounds

𝔼¯[f¯nLtLx22]=𝔼[fnLtLx22]1,𝔼¯[ρ¯φnLt2Hxη2]=𝔼[ρφnLt2Hxη2]1\displaystyle\overline{\mathbb{E}}\Big{[}\|\overline{f}^{n}\|_{L^{\infty}_{t}L^{2}_{x}}^{2}\Big{]}=\mathbb{E}\Big{[}\|f^{n}\|_{L^{\infty}_{t}L^{2}_{x}}^{2}\Big{]}\lesssim 1,\hskip 14.22636pt\overline{\mathbb{E}}\Big{[}\|\overline{\rho}^{n}_{\varphi}\|^{2}_{L^{2}_{t}H^{\eta}_{x}}\Big{]}=\mathbb{E}\Big{[}\|\rho^{n}_{\varphi}\|^{2}_{L^{2}_{t}H^{\eta}_{x}}\Big{]}\lesssim 1 (3.44)

uniformly in n1n\geq 1, we derive in particular that the families of random variables (f¯n)n(\overline{f}^{n})_{n} and (ρ¯φn)n(\overline{\rho}^{n}_{\varphi})_{n} are uniformly integrable, and therefore

f¯nf¯ a.s and in L1(Ω¯;C([0,T];HW1σ(2d))).\displaystyle\overline{f}^{n}\to\overline{f}\text{ a.s and in }L^{1}\Big{(}\overline{\Omega};C([0,T];H^{-\sigma}_{W^{-1}}(\mathbb{R}^{2d}))\Big{)}. (3.45)
φ𝒟,ρ¯φnρ¯φ a.s and in L1(Ω¯;L2([0,T];L2(d))).\displaystyle\forall\varphi\in{\cal D},\;\;\overline{\rho}^{n}_{\varphi}\to\overline{\rho}_{\varphi}\text{ a.s and in }L^{1}\Big{(}\overline{\Omega};L^{2}([0,T];L^{2}(\mathbb{R}^{d}))\Big{)}. (3.46)

Consequently, up to a subsequence, we may also assume that

φ𝒟,[ρ¯φn(t,x)ρ¯φ(t,x),dtdx-a.e],¯-a.s.\displaystyle\forall\varphi\in{\cal D},\;\;\Big{[}\overline{\rho}^{n}_{\varphi}(t,x)\to\overline{\rho}_{\varphi}(t,x),\;dt\otimes dx\text{-a.e}\Big{]},\;\;\overline{\mathbb{P}}\text{-a.s}. (3.47)
Remark 3.1.

For all φ𝒟\varphi\in{\cal D}, from the equality ρφn=φ(v)fn𝑑v\rho^{n}_{\varphi}=\int\varphi(v)f^{n}dv, \mathbb{P}-a.s, it is clear that ρ¯φn=dφ(v)f¯n𝑑v\overline{\rho}^{n}_{\varphi}=\int_{\mathbb{R}^{d}}\varphi(v)\overline{f}^{n}dv, ¯\overline{\mathbb{P}}-a.s. The convergence (3.45) then guarantees that, for all φ𝒟\varphi\in{\cal D}, ρ¯φ\overline{\rho}_{\varphi} is indeed given by

ρ¯φ=dφ(v)f¯𝑑v, ¯-a.s.\overline{\rho}_{\varphi}=\int_{\mathbb{R}^{d}}\varphi(v)\overline{f}dv,\;\text{ $\overline{\mathbb{P}}$-a.s}.

Let us now introduce the averaged quantities

ρ=df𝑑v,j=dvf𝑑v.\displaystyle\rho=\int_{\mathbb{R}^{d}}fdv,\hskip 28.45274ptj=\int_{\mathbb{R}^{d}}vfdv.

By requiring some greater moments for the initial data, we may extend the convergence (3.46) to ρ\rho and jj.

Lemma 3.3.

Assume that the initial data f0f_{0} satisfies, for some θ(0,1)\theta\in(0,1),

z2d|f0(z)|p𝑑z+z2d(1+|v|k)f0(z)𝑑z< with p=1+1θ,k>d+21θ.\displaystyle\int_{z\in\mathbb{R}^{2d}}|f_{0}(z)|^{p}dz+\int_{z\in\mathbb{R}^{2d}}(1+|v|^{k})f_{0}(z)dz<\infty\;\text{ with }\;p=1+\frac{1}{\theta},\hskip 11.38109ptk>\frac{d+2}{1-\theta}. (3.48)

Then the following convergences hold in L1(Ω¯;L2([0,T];L2(d)))L^{1}\Big{(}\overline{\Omega};L^{2}([0,T];L^{2}(\mathbb{R}^{d}))\Big{)}:

ρ¯nρ¯,j¯nj¯,ρ¯θRnnj¯,\displaystyle\overline{\rho}^{n}\to\overline{\rho},\hskip 28.45274pt\overline{j}^{n}\to\overline{j},\hskip 28.45274pt\overline{\rho}^{n}_{\theta_{R_{n}}}\to\overline{j},
ϕρ¯nϕρ¯,ϕj¯nϕj¯,ϕρ¯θRnnϕj¯\displaystyle\phi*\overline{\rho}^{n}\to\phi*\overline{\rho},\hskip 28.45274pt\phi*\overline{j}^{n}\to\phi*\overline{j},\hskip 28.45274pt\phi*\overline{\rho}^{n}_{\theta_{R_{n}}}\to\phi*\overline{j}

Consequently, up to a subsequence, we may also assume that these convergences hold dtdxdt\otimes dx almost everywhere, ¯\overline{\mathbb{P}} almost surely.

Proof.

Let us, for instance, prove the convergence of j¯n\overline{j}^{n}. For fixed N1N\geq 1, denoting φ=θRN\varphi=\theta_{R_{N}},

j¯nj¯Lt2Lx2j¯nρ¯φnLt2Lx2+ρ¯φnρ¯φLt2Lx2+ρ¯φj¯Lt2Lx2\displaystyle\|\overline{j}^{n}-\overline{j}\|_{L^{2}_{t}L^{2}_{x}}\leq\|\overline{j}^{n}-\overline{\rho}^{n}_{\varphi}\|_{L^{2}_{t}L^{2}_{x}}+\|\overline{\rho}^{n}_{\varphi}-\overline{\rho}_{\varphi}\|_{L^{2}_{t}L^{2}_{x}}+\|\overline{\rho}_{\varphi}-\overline{j}\|_{L^{2}_{t}L^{2}_{x}} (3.49)

and we note that, for any γ>0\gamma>0 and m>dm>d,

|j¯nρ¯φn|2\displaystyle|\overline{j}^{n}-\overline{\rho}^{n}_{\varphi}|^{2} =||vθRN(v)|f¯n𝑑v|2||v|RN|v|f¯n𝑑v|2RN2γ|d|v|1+γf¯n𝑑v|2\displaystyle=\Big{|}\int|v-\theta_{R_{N}}(v)|\overline{f}^{n}dv\Big{|}^{2}\lesssim\Big{|}\int_{|v|\geq R_{N}}|v|\overline{f}^{n}dv\Big{|}^{2}\leq R_{N}^{-2\gamma}\Big{|}\int_{\mathbb{R}^{d}}|v|^{1+\gamma}\overline{f}^{n}dv\Big{|}^{2}
RN2γd(1+|v|)2+2γ+m|f¯n|2𝑑v\displaystyle\lesssim R_{N}^{-2\gamma}\int_{\mathbb{R}^{d}}(1+|v|)^{2+2\gamma+m}|\overline{f}^{n}|^{2}dv
RN2γ(d(1+|v|)2+2γ+m1θf¯n𝑑v+d|f¯n|1+1θ𝑑v)\displaystyle\lesssim R_{N}^{-2\gamma}\Big{(}\int_{\mathbb{R}^{d}}(1+|v|)^{\frac{2+2\gamma+m}{1-\theta}}\overline{f}^{n}dv+\int_{\mathbb{R}^{d}}|\overline{f}^{n}|^{1+\frac{1}{\theta}}dv\Big{)}
RN2γ(d(1+|v|)kf¯n𝑑v+d|f¯n|p𝑑v)\displaystyle\lesssim R_{N}^{-2\gamma}\Big{(}\int_{\mathbb{R}^{d}}(1+|v|)^{k}\overline{f}^{n}dv+\int_{\mathbb{R}^{d}}|\overline{f}^{n}|^{p}dv\Big{)}

for (γ,m)(\gamma,m) close enough to (0,d)(0,d). As a result,

supn1𝔼¯[j¯nρ¯φnLt2Lx22]RN2γ0 as N.\displaystyle\sup_{n\geq 1}\overline{\mathbb{E}}\Big{[}\|\overline{j}^{n}-\overline{\rho}^{n}_{\varphi}\|_{L^{2}_{t}L^{2}_{x}}^{2}\Big{]}\lesssim R_{N}^{-2\gamma}\to 0\text{ as }N\to\infty.

Since, from (3.45), we classically derive

𝔼¯[0T2d((1+|v|)kf¯+|f¯|p)𝑑z𝑑t]supn1𝔼¯[0T2d((1+|v|)kf¯n+|f¯n|p)𝑑z𝑑t]1\overline{\mathbb{E}}\Big{[}\int_{0}^{T}\int_{\mathbb{R}^{2d}}\Big{(}(1+|v|)^{k}\overline{f}+|\overline{f}|^{p}\Big{)}dzdt\Big{]}\leq\sup_{n\geq 1}\overline{\mathbb{E}}\Big{[}\int_{0}^{T}\int_{\mathbb{R}^{2d}}\Big{(}(1+|v|)^{k}\overline{f}^{n}+|\overline{f}^{n}|^{p}\Big{)}dzdt\Big{]}\lesssim 1

we deduce similarly that 𝔼¯[j¯ρ¯φLt2Lx22]0\overline{\mathbb{E}}\Big{[}\|\overline{j}-\overline{\rho}_{\varphi}\|_{L^{2}_{t}L^{2}_{x}}^{2}\Big{]}\to 0 as NN goes to infinity. We may hence come back to (3.49) and conclude. The convergence of the convoluted functions is easily deduced.

For fixed n1n\geq 1, fRnf^{R_{n}} defines a (strong) solution of (2.8) on (Ω,,)(\Omega,{\cal F},\mathbb{P}). In particular, it satisfies the associated martingale problem: recalling the operator R[f]{\cal L}_{R}[f] defined in (2.9), for all ΨCc(2d)\Psi\in C_{c}^{\infty}(\mathbb{R}^{2d}), the process

MΨn(t)=Ψ,fRn(t)Ψ,f00t<Rn[fRn(s)]Ψ,fRn(s)>ds,t[0,T]\displaystyle M^{n}_{\Psi}(t)=\langle\Psi,f^{R_{n}}(t)\rangle-\langle\Psi,f_{0}\rangle-\int_{0}^{t}\Bigl{<}{\cal L}_{R_{n}}[f^{R_{n}}(s)]\Psi,f^{R_{n}}(s)\Bigr{>}ds,\hskip 14.22636ptt\in[0,T] (3.50)

defines a continuous, real valued L2L^{2} martingale on (Ω,,)(\Omega,{\cal F},\mathbb{P}) with respect to the filtration

tn=σ(fRn(s)HW1σ(2d),s[0,t]),t[0,T].{\cal F}^{n}_{t}=\sigma\Big{(}f^{R_{n}}(s)\in H^{-\sigma}_{W^{-1}}(\mathbb{R}^{2d}),\;\;s\in[0,t]\Big{)},\;\;\;t\in[0,T].

Its quadratic variation is given by

[MΨn](t)=VΨn(t):=0t|<KRn[fRn(s)]vΨ,fRn(s)>|2ds.\displaystyle\Big{[}M^{n}_{\Psi}\Big{]}(t)=V^{n}_{\Psi}(t):=\int_{0}^{t}\Big{|}\Bigl{<}K_{R_{n}}[f^{R_{n}}(s)]\cdot\nabla_{v}\Psi,f^{R_{n}}(s)\Bigr{>}\Big{|}^{2}ds. (3.51)

We are now ready to state the following result.

Proposition 3.7.

Let us introduce, on (Ω¯,¯,¯)(\overline{\Omega},\overline{\cal F},\overline{\mathbb{P}}), the filtration

¯t=σ(f¯(s)HW1σ(2d),s[0,t]),t[0,T].\overline{\cal F}_{t}=\sigma\Big{(}\overline{f}(s)\in H^{-\sigma}_{W^{-1}}(\mathbb{R}^{2d}),\;\;s\in[0,t]\Big{)},\;\;\;t\in[0,T].

Recalling the operator [f]{\cal L}[f] defined in (1.15), for all test functions Ψ\Psi of the form

Ψ(x,v)=Ψ1(x)Ψ2(v),Ψ1,Ψ2Cc(d) with vΨ2𝒟,\displaystyle\Psi(x,v)=\Psi_{1}(x)\Psi_{2}(v),\hskip 14.22636pt\Psi_{1},\Psi_{2}\in C_{c}^{\infty}(\mathbb{R}^{d})\text{~with }\nabla_{v}\Psi_{2}\in{\cal D}, (3.52)

the process

M¯Ψ(t)=Ψ,f¯(t)Ψ,f00t<[f¯(s)]Ψ,f¯(s)>ds,t[0,T]\displaystyle\overline{M}_{\Psi}(t)=\langle\Psi,\overline{f}(t)\rangle-\langle\Psi,f_{0}\rangle-\int_{0}^{t}\Bigl{<}{\cal L}[\overline{f}(s)]\Psi,\overline{f}(s)\Bigr{>}ds,\hskip 14.22636ptt\in[0,T]

defines a continuous, real-valued L2L^{2} martingale with respect to (¯t)t0(\overline{\cal F}_{t})_{t\geq 0}, with quadratic variation

[M¯Ψ](t)=V¯Ψ(t):=0t|<K[f¯(s)]vΨ,f¯(s)>|2ds.\displaystyle\Big{[}\overline{M}_{\Psi}\Big{]}(t)=\overline{V}_{\Psi}(t):=\int_{0}^{t}\Big{|}\Bigl{<}K[\overline{f}(s)]\cdot\nabla_{v}\Psi,\overline{f}(s)\Bigr{>}\Big{|}^{2}ds.
Remark 3.2.

The assumption vΨ2𝒟\nabla_{v}\Psi_{2}\in{\cal D} in (3.52) is only technical: for any given countable family F=(Ψ2)Ψ2FF~{}=~{}(\Psi_{2})_{\Psi_{2}\in F} of test functions in Cc(d)C_{c}^{\infty}(\mathbb{R}^{d}), on can initially choose the countable subset 𝒟{\cal D} such that {vΨ2,Ψ2F}𝒟\{\nabla_{v}\Psi_{2},\;\Psi_{2}\in F\}\subset{\cal D}, so that the conclusion of Proposition 3.7 holds true for all Ψ(x,v)=Ψ1(x)Ψ2(v)\Psi(x,v)=\Psi_{1}(x)\Psi_{2}(v) with Ψ1Cc(d)\Psi_{1}\in C_{c}^{\infty}(\mathbb{R}^{d}) and Ψ2F\Psi_{2}\in F.

Proof of Proposition 3.7.

The martingale problem set on Ω\Omega may be expressed as

𝔼[(MΨn(t)MΨn(s))Hn]=0,\displaystyle\mathbb{E}\Big{[}\Big{(}M^{n}_{\Psi}(t)-M^{n}_{\Psi}(s)\Big{)}H^{n}\Big{]}=0,
𝔼[|MΨn(t)MΨn(s)|2Hn]=𝔼[(VΨn(t)VΨn(s))Hn]\displaystyle\mathbb{E}\Big{[}\Big{|}M^{n}_{\Psi}(t)-M^{n}_{\Psi}(s)\Big{|}^{2}H^{n}\Big{]}=\mathbb{E}\Big{[}\Big{(}V^{n}_{\Psi}(t)-V^{n}_{\Psi}(s)\Big{)}H^{n}\Big{]}

for all Hn=h(fRn(ti),1im)H^{n}=h\Big{(}f^{R_{n}}(t_{i}),1\leq i\leq m\Big{)}, where 0t1,,tmst0\leq t_{1},\ldots,t_{m}\leq s\leq t, and h:(HW1σ(2d))mh:(H^{-\sigma}_{W^{-1}}(\mathbb{R}^{2d}))^{m}\to\mathbb{R} is continuous and bounded, Since the laws of fnf^{n} and f¯n\overline{f}^{n} coincide, it follows that, on Ω¯\overline{\Omega},

𝔼¯[(M¯Ψn(t)M¯Ψn(s))H¯n]=0,\displaystyle\overline{\mathbb{E}}\Big{[}\Big{(}\overline{M}^{n}_{\Psi}(t)-\overline{M}^{n}_{\Psi}(s)\Big{)}\overline{H}^{n}\Big{]}=0, (3.53)
𝔼¯[|M¯Ψn(t)M¯Ψn(s)|2H¯n]=𝔼¯[(V¯Ψn(t)V¯Ψn(s))H¯n]\displaystyle\overline{\mathbb{E}}\Big{[}\Big{|}\overline{M}^{n}_{\Psi}(t)-\overline{M}^{n}_{\Psi}(s)\Big{|}^{2}\overline{H}^{n}\Big{]}=\overline{\mathbb{E}}\Big{[}\Big{(}\overline{V}_{\Psi}^{n}(t)-\overline{V}_{\Psi}^{n}(s)\Big{)}\overline{H}^{n}\Big{]} (3.54)

where H¯n=h(f¯n(ti),1im)\overline{H}^{n}=h\Big{(}\overline{f}^{n}(t_{i}),1\leq i\leq m\Big{)} and M¯Ψn\overline{M}^{n}_{\Psi} and V¯Ψn(t)\overline{V}^{n}_{\Psi}(t) are naturally defined on Ω¯\overline{\Omega}, as M¯Ψ\overline{M}_{\Psi} and V¯Ψ(t)\overline{V}_{\Psi}(t) . We may decompose these into

M¯Ψn(t)=Ψ,f¯n(t)Ψ,f00t(<vxΨ,f¯n(s)>+i=14ni(f¯n(s)))ds,\displaystyle\overline{M}_{\Psi}^{n}(t)=\langle\Psi,\overline{f}^{n}(t)\rangle-\langle\Psi,f_{0}\rangle-\int_{0}^{t}\Big{(}\Bigl{<}v\cdot\nabla_{x}\Psi,\overline{f}^{n}(s)\Bigr{>}+\sum_{i=1}^{4}{\cal M}^{i}_{n}(\overline{f}^{n}(s))\Big{)}ds, (3.55)
V¯Ψn(t)=0t|5(f¯n(s))|2𝑑s,\displaystyle\overline{V}^{n}_{\Psi}(t)=\int_{0}^{t}\Big{|}{\cal M}^{5}(\overline{f}^{n}(s))\Big{|}^{2}ds,

where, recalling expressions (2.2),

n1(f)\displaystyle{\cal M}^{1}_{n}(f) =xvLRnCS[f]vΨfdxdv=xv(Φn1f)vΨfdxdv,\displaystyle=\int_{x}\int_{v}L^{CS}_{R_{n}}[f]\cdot\nabla_{v}\Psi fdxdv=\int_{x}\int_{v}(\Phi^{1}_{n}*f)\cdot\nabla_{v}\Psi fdxdv,
n2(f)\displaystyle{\cal M}^{2}_{n}(f) =xvSRn[f]vΨfdxdv\displaystyle=\int_{x}\int_{v}S_{R_{n}}[f]\cdot\nabla_{v}\Psi fdxdv
=12xv(ψ~(Φn2+Φn3f)ψ~L1(Φn2+Φn3f))vΨfdxdv,\displaystyle=\frac{1}{2}\int_{x}\int_{v}\Big{(}\widetilde{\psi}*(\Phi^{2}_{n}+\Phi^{3}_{n}*f)-\|\widetilde{\psi}\|_{L^{1}}(\Phi^{2}_{n}+\Phi^{3}_{n}*f)\Big{)}\cdot\nabla_{v}\Psi fdxdv,
n3(f)\displaystyle{\cal M}^{3}_{n}(f) =121i,jdxvKRni[f]KRnj[f]vivj2Ψfdxdv\displaystyle=\frac{1}{2}\sum_{1\leq i,j\leq d}\int_{x}\int_{v}K_{R_{n}}^{i}[f]K_{R_{n}}^{j}[f]\partial^{2}_{v_{i}v_{j}}\Psi fdxdv
=121i,jdxv(Φn2+Φn3f)i(Φn2+Φn3f)jvivj2Ψfdxdv,\displaystyle=\frac{1}{2}\sum_{1\leq i,j\leq d}\int_{x}\int_{v}\Big{(}\Phi^{2}_{n}+\Phi^{3}_{n}*f\Big{)}^{i}\Big{(}\Phi^{2}_{n}+\Phi^{3}_{n}*f\Big{)}^{j}\partial^{2}_{v_{i}v_{j}}\Psi fdxdv,
n4(f)\displaystyle{\cal M}^{4}_{n}(f) =xvLRMT[f]vΨfdxdv=xv(uR[f]v)vΨfdxdv\displaystyle=\int_{x}\int_{v}L^{MT}_{R}[f]\cdot\nabla_{v}\Psi fdxdv=\int_{x}\int_{v}(u_{R}[f]-v)\cdot\nabla_{v}\Psi fdxdv
=xϕρθRnR1+ϕρρΨ2Ψ1𝑑xxvvvΨfdxdv\displaystyle=\int_{x}\frac{\phi*\rho_{\theta_{R_{n}}}}{R^{-1}+\phi*\rho}\cdot\rho_{\nabla\Psi_{2}}\Psi_{1}dx-\int_{x}\int_{v}v\cdot\nabla_{v}\Psi fdxdv
n5(f)\displaystyle{\cal M}^{5}_{n}(f) =xvKRn[f]vΨfdxdv=xv(Φn2+Φn3f)vΨfdxdv\displaystyle=\int_{x}\int_{v}K_{R_{n}}[f]\cdot\nabla_{v}\Psi fdxdv=\int_{x}\int_{v}(\Phi^{2}_{n}+\Phi^{3}_{n}*f)\cdot\nabla_{v}\Psi fdxdv

with

Φn1(x,v)=χRn(x)ψ(x)θRn(v),\displaystyle\Phi^{1}_{n}(x,v)=\chi_{R_{n}}(x)\psi(x)\theta_{R_{n}}(-v),
Φn2(x)=χRn(x)F(x),\displaystyle\Phi^{2}_{n}(x)=\chi_{R_{n}}(x)F(x),
Φn3(x,v)=χRn(x)ψ~(x)θRn(v),\displaystyle\Phi^{3}_{n}(x,v)=\chi_{R_{n}}(x)\widetilde{\psi}(x)\theta_{R_{n}}(-v),

We wish to send nn to infinity in (3.53) and (3.54). Thanks to the convergence (3.45), the first three linear terms in (3.55) cause no issue ; let us hence focus on the remaining terms. Let us consider the term involving n1(f){\cal M}^{1}_{n}(f): defining the natural limiting term

1(f)=xv(Φ1f)vΨfdxdv{\cal M}^{1}(f)=\int_{x}\int_{v}(\Phi^{1}*f)\cdot\nabla_{v}\Psi fdxdv

with Φ1(x,v)=ψ(x)(v)\Phi^{1}(x,v)=\psi(x)(-v), we have

|stn1(f¯n(σ))𝑑σst1(f¯(σ))𝑑σ|𝒥n1+𝒥n2\displaystyle\Big{|}\int_{s}^{t}{\cal M}^{1}_{n}(\overline{f}^{n}(\sigma))d\sigma-\int_{s}^{t}{\cal M}^{1}(\overline{f}(\sigma))d\sigma\Big{|}\leq{\cal J}^{1}_{n}+{\cal J}^{2}_{n} (3.56)

where

𝒥n1=stxv|(Φn1f¯n)(Φ1f¯)||vΨ||f¯|𝑑x𝑑v𝑑σ,\displaystyle{\cal J}^{1}_{n}=\int_{s}^{t}\int_{x}\int_{v}\Big{|}(\Phi^{1}_{n}*\overline{f}^{n})-(\Phi^{1}*\overline{f})\Big{|}|\nabla_{v}\Psi||\overline{f}|dxdvd\sigma,
𝒥n2=|stxv(Φn1f¯n)vΨ(f¯nf¯)𝑑x𝑑v𝑑σ|.\displaystyle{\cal J}^{2}_{n}=\Big{|}\int_{s}^{t}\int_{x}\int_{v}(\Phi^{1}_{n}*\overline{f}^{n})\cdot\nabla_{v}\Psi(\overline{f}^{n}-\overline{f})dxdvd\sigma\Big{|}.

First, for 𝒥n1{\cal J}^{1}_{n} we have, thanks to (3.44),

𝔼[𝒥1]\displaystyle\mathbb{E}\Big{[}{\cal J}^{1}\Big{]} 𝔼[stzSupp(Ψ)|(Φn1f¯n)(Φ1f¯)|2𝑑z𝑑σ]1/2𝔼[stz|f¯|2𝑑z𝑑σ]1/2\displaystyle\lesssim\mathbb{E}\Big{[}\int_{s}^{t}\int_{z\in Supp(\Psi)}\Big{|}(\Phi^{1}_{n}*\overline{f}^{n})-(\Phi^{1}*\overline{f})\Big{|}^{2}dzd\sigma\Big{]}^{1/2}\mathbb{E}\Big{[}\int_{s}^{t}\int_{z}|\overline{f}|^{2}dzd\sigma\Big{]}^{1/2}
𝔼[stzSupp(Ψ)|(Φn1f¯n)(Φ1f¯)|2𝑑z𝑑σ]1/2.\displaystyle\lesssim\mathbb{E}\Big{[}\int_{s}^{t}\int_{z\in Supp(\Psi)}\Big{|}(\Phi^{1}_{n}*\overline{f}^{n})-(\Phi^{1}*\overline{f})\Big{|}^{2}dzd\sigma\Big{]}^{1/2}. (3.57)

For all nm1n\geq m\geq 1, for fixed (x,v)Supp(Ψ)(x,v)\in Supp(\Psi), we may write

𝔼¯[|Φn1f¯nΦ1f¯|2]\displaystyle\overline{\mathbb{E}}\Big{[}\Big{|}\Phi^{1}_{n}*\overline{f}^{n}-\Phi^{1}*\overline{f}\Big{|}^{2}\Big{]} 𝔼¯[|Φm1f¯nΦm1f¯|2]+𝔼¯[|(Φn1Φm1)f¯n|2]\displaystyle\lesssim\overline{\mathbb{E}}\Big{[}\Big{|}\Phi^{1}_{m}*\overline{f}^{n}-\Phi^{1}_{m}*\overline{f}\Big{|}^{2}\Big{]}+\overline{\mathbb{E}}\Big{[}\Big{|}(\Phi^{1}_{n}-\Phi^{1}_{m})*\overline{f}^{n}\Big{|}^{2}\Big{]}
+𝔼¯[|(Φ1Φm1)f¯|2]\displaystyle\hskip 14.22636pt+\overline{\mathbb{E}}\Big{[}\Big{|}(\Phi^{1}-\Phi^{1}_{m})*\overline{f}\Big{|}^{2}\Big{]}

which converges to 0 as nn goes to infinity thanks to the convergence (3.45), and the bounds

𝔼¯[|(Φn1Φm1)f¯n|2]\displaystyle\overline{\mathbb{E}}\Big{[}\Big{|}(\Phi^{1}_{n}-\Phi^{1}_{m})*\overline{f}^{n}\Big{|}^{2}\Big{]} 𝔼¯[||vw|Rm|vw|fn(y,w)dydw|2]\displaystyle\lesssim\overline{\mathbb{E}}\Big{[}\Big{|}\int\int_{|v-w|\geq R_{m}}|v-w|f^{n}(y,w)dydw\Big{|}^{2}\Big{]}
Rm2(|v|4+𝔼[|w|4fn𝑑y𝑑w])Rm2(1+|v|4)\displaystyle\lesssim R_{m}^{-2}\Big{(}|v|^{4}+\mathbb{E}\Big{[}\int|w|^{4}f^{n}dydw\Big{]}\Big{)}\lesssim R_{m}^{-2}(1+|v|^{4})

and, for some γ>0\gamma>0, recalling that (x,v)Supp(Ψ)(x,v)\in Supp(\Psi),

𝔼¯[|Φm1f¯n|2+γ(x,v)]1+𝔼¯[|w|2+γf¯n𝑑yfw]1\displaystyle\overline{\mathbb{E}}\Big{[}\Big{|}\Phi_{m}^{1}*\overline{f}^{n}\Big{|}^{2+\gamma}(x,v)\Big{]}\lesssim 1+\overline{\mathbb{E}}\Big{[}\int|w|^{2+\gamma}\overline{f}^{n}dyfw\Big{]}\lesssim 1

for all nmn\geq m. Note that this last bound also guarantees the uniform integrability in (ω,σ,x,v)(\omega,\sigma,x,v) of the integrand in (3.57), so that 𝔼¯[𝒥n1]0\overline{\mathbb{E}}\Big{[}{\cal J}^{1}_{n}\Big{]}\to 0. Additionally, for γ>0\gamma>0 small enough, the bound

𝔼¯[stzSupp(Ψ)|Φn1f¯n|2+γ|f¯|2+γ𝑑z𝑑σ]\displaystyle\overline{\mathbb{E}}\Big{[}\int_{s}^{t}\int_{z\in Supp(\Psi)}\Big{|}\Phi^{1}_{n}*\overline{f}^{n}\Big{|}^{2+\gamma}|\overline{f}|^{2+\gamma}dzd\sigma\Big{]}
𝔼¯[stzSupp(Ψ)(1+|w|2+γf¯n𝑑y𝑑w)|f¯|2+γ𝑑z𝑑σ]\displaystyle\hskip 85.35826pt\lesssim\overline{\mathbb{E}}\Big{[}\int_{s}^{t}\int_{z\in Supp(\Psi)}\Big{(}1+\int|w|^{2+\gamma}\overline{f}^{n}dydw\Big{)}|\overline{f}|^{2+\gamma}dzd\sigma\Big{]}
𝔼[st2d|v|kf¯n+|f¯|pdzdσ]1\displaystyle\hskip 142.26378pt\lesssim\mathbb{E}\Big{[}\int_{s}^{t}\int_{\mathbb{R}^{2d}}|v|^{k}\overline{f}^{n}+|\overline{f}|^{p}dzd\sigma\Big{]}\lesssim 1

guarantees that 𝔼[|𝒥n1|2+γ]1.\mathbb{E}\Big{[}\Big{|}{\cal J}^{1}_{n}\Big{|}^{2+\gamma}\Big{]}\lesssim 1. Similarly, for 𝒥n2{\cal J}^{2}_{n}, we write, for all nm1n\geq m\geq 1,

𝒥n2\displaystyle{\cal J}^{2}_{n}\leq |stxv(Φm1f¯m)vΨ(f¯nf¯)𝑑x𝑑v𝑑σ|\displaystyle\Big{|}\int_{s}^{t}\int_{x}\int_{v}(\Phi^{1}_{m}*\overline{f}^{m})\cdot\nabla_{v}\Psi(\overline{f}^{n}-\overline{f})dxdvd\sigma\Big{|}
+stxv|(Φm1f¯m)(Φn1f¯n)||vΨ|(|f¯n|+|f¯|)𝑑x𝑑v𝑑σ\displaystyle\;\;+\int_{s}^{t}\int_{x}\int_{v}\Big{|}(\Phi^{1}_{m}*\overline{f}^{m})-(\Phi^{1}_{n}*\overline{f}^{n})\Big{|}|\nabla_{v}\Psi|(|\overline{f}^{n}|+|\overline{f}|)dxdvd\sigma

so that

𝔼[𝒥n2]\displaystyle\mathbb{E}\Big{[}{\cal J}_{n}^{2}\Big{]}\lesssim 𝔼[|stxv((Φm1f¯m)vΨ)(f¯nf¯)𝑑x𝑑v𝑑σ|]\displaystyle\;\mathbb{E}\Big{[}\Big{|}\int_{s}^{t}\int_{x}\int_{v}\Big{(}(\Phi^{1}_{m}*\overline{f}^{m})\cdot\nabla_{v}\Psi\Big{)}(\overline{f}^{n}-\overline{f})dxdvd\sigma\Big{|}\Big{]}
+𝔼[stzSupp(Ψ)|(Φm1f¯m)(Φn1f¯n)|2𝑑z𝑑σ].\displaystyle+\mathbb{E}\Big{[}\int_{s}^{t}\int_{z\in Supp(\Psi)}\Big{|}(\Phi^{1}_{m}*\overline{f}^{m})-(\Phi^{1}_{n}*\overline{f}^{n})\Big{|}^{2}dzd\sigma\Big{]}.

Since, for fixed m1m\geq 1, (Φm1f¯m)vΨCc(2d)(\Phi^{1}_{m}*\overline{f}^{m})\cdot\nabla_{v}\Psi\in C_{c}^{\infty}(\mathbb{R}^{2d}), we conclude in a similar fashion that 𝔼[𝒥n2]0\mathbb{E}\Big{[}{\cal J}^{2}_{n}\Big{]}\to 0 and 𝔼[|𝒥n2|2+γ]1\mathbb{E}\Big{[}\Big{|}{\cal J}^{2}_{n}\Big{|}^{2+\gamma}\Big{]}\lesssim 1. Coming back to (3.56), we have shown that

stn1(f¯n(σ))𝑑σst1(f¯(σ))𝑑σ in probability,\displaystyle\int_{s}^{t}{\cal M}^{1}_{n}(\overline{f}^{n}(\sigma))d\sigma\to\int_{s}^{t}{\cal M}^{1}(\overline{f}(\sigma))d\sigma\text{ in probability},
𝔼¯[|stn1(f¯n(σ))𝑑σ|2+γ]1,\displaystyle\overline{\mathbb{E}}\Big{[}\Big{|}\int_{s}^{t}{\cal M}^{1}_{n}(\overline{f}^{n}(\sigma))d\sigma\Big{|}^{2+\gamma}\Big{]}\lesssim 1,

which is sufficient to pass to the limit in the corresponding term of (3.53) and the left-hand side of (3.54).

The terms involving ni(f){\cal M}^{i}_{n}(f) for i=2,3,5i=2,3,5 can be treated with similar arguments. Let us now handle the more delicate term, involving n4(f){\cal M}^{4}_{n}(f): we wish to prove that

stxϕρ¯θRnnRn1+ϕρ¯nρ¯Ψ2nΨ1𝑑x𝑑σ=stxuR[f¯n]ρ¯Ψ2nΨ1𝑑x𝑑σ\displaystyle\int_{s}^{t}\int_{x}\frac{\phi*\overline{\rho}^{n}_{\theta_{R_{n}}}}{R_{n}^{-1}+\phi*\overline{\rho}^{n}}\cdot\overline{\rho}^{n}_{\nabla\Psi_{2}}\Psi_{1}dxd\sigma=\int_{s}^{t}\int_{x}u_{R}[\overline{f}^{n}]\cdot\overline{\rho}^{n}_{\nabla\Psi_{2}}\Psi_{1}dxd\sigma (3.58)

converges to the expected limiting term. Let us introduce

J¯n:=uRn[f¯n]ρ¯Ψ2n=uRn[f¯n]vvΨ2(v)f¯n𝑑v.\displaystyle\overline{J}^{n}:=u_{R_{n}}[\overline{f}^{n}]\cdot\overline{\rho}^{n}_{\nabla\Psi_{2}}=u_{R_{n}}[\overline{f}^{n}]\cdot\int_{v}\nabla_{v}\Psi_{2}(v)\overline{f}^{n}dv. (3.59)

For some q>2q>2 and τ(0,1)\tau\in(0,1), we have

|J¯n|q𝑑xxv|uRn[f¯n]|q|f¯n|q𝑑x𝑑v\displaystyle\int|\overline{J}^{n}|^{q}dx\lesssim\int_{x}\int_{v}|u_{R_{n}}[\overline{f}^{n}]|^{q}|\overline{f}^{n}|^{q}dxdv xv|uRn[f¯n]|q1τf¯n𝑑x𝑑v+xv|f¯n|1+q1τ𝑑x𝑑v\displaystyle\lesssim\int_{x}\int_{v}|u_{R_{n}}[\overline{f}^{n}]|^{\frac{q}{1-\tau}}\overline{f}^{n}dxdv+\int_{x}\int_{v}|\overline{f}^{n}|^{1+\frac{q-1}{\tau}}dxdv
xv(1+|uRn[f¯n]|k)f¯n𝑑x𝑑v+xv|f¯n|p𝑑x𝑑v\displaystyle\lesssim\int_{x}\int_{v}(1+|u_{R_{n}}[\overline{f}^{n}]|^{k})\overline{f}^{n}dxdv+\int_{x}\int_{v}|\overline{f}^{n}|^{p}dxdv

as soon as τ=θ(q1)\tau=\theta(q-1) and q1τ41θ\frac{q}{1-\tau}\leq\frac{4}{1-\theta}, which can be met for qq close enough to 22. Hence, thanks to Proposition 3.2, we derive

𝔼¯[supt[0,T]x|J¯n(t,x)|q𝑑x]𝔼[supt[0,T](z(1+|v|)kfn(t,z)𝑑z+z|fn(t,z)|p𝑑z)]1.\displaystyle\overline{\mathbb{E}}\Big{[}\sup_{t\in[0,T]}\int_{x}|\overline{J}^{n}(t,x)|^{q}dx\Big{]}\lesssim\mathbb{E}\Big{[}\sup_{t\in[0,T]}\Big{(}\int_{z}(1+|v|)^{k}f^{n}(t,z)dz+\int_{z}|f^{n}(t,z)|^{p}dz\Big{)}\Big{]}\lesssim 1. (3.60)

As a consequence, up to some subsequence which we omit for simplicity,

J¯nJ¯ weak  inLq(Ω¯;L([0,T];Lq(d))).\displaystyle\overline{J}^{n}\rightharpoonup\overline{J}\;\;\text{ weak $*$ in}~L^{q}\Big{(}\overline{\Omega};L^{\infty}([0,T];L^{q}(\mathbb{R}^{d}))\Big{)}. (3.61)

It is clear that this weak convergence is enough for the term (3.58) to pass to the limit in (3.53). Therefore, it only remains to identify the limit as

J¯=u¯ρ¯Ψ2, where u¯(x,t):=u[f¯](x,t)={(ϕj¯)(t,x)(ϕρ¯)(t,x) if (ϕρ¯)(t,x)00 if (ϕρ¯)(t,x)=0.\displaystyle\overline{J}=\overline{u}\cdot\overline{\rho}_{\nabla\Psi_{2}},\;\text{~where }\;\overline{u}(x,t):=u[\overline{f}](x,t)=\left\{\begin{array}[]{l l}\displaystyle{\frac{(\phi*\overline{j})(t,x)}{(\phi*\overline{\rho})(t,x)}}&\text{ if }(\phi*\overline{\rho})(t,x)\neq 0\\ 0&\text{~if }(\phi*~\overline{\rho})(t,x)=0.\end{array}\right. (3.64)

First, considering the set, for r>0r>0, ωΩ¯\omega\in\overline{\Omega},

Ar(ω)={(t,x)[0,T]×B(0,r)|(ϕρ¯)(ω,t,x)=0},A_{r}(\omega)=\{(t,x)\in[0,T]\times B(0,r)\;|\;(\phi*\overline{\rho})(\omega,t,x)=0\},

we have

Ar|J¯n|𝑑x𝑑t\displaystyle\int_{A_{r}}|\overline{J}^{n}|dxdt ((t,x)ArvSupp(Ψ2)|uRn[f¯n]|2f¯n𝑑x𝑑t)1/2(Arρ¯n𝑑x𝑑t)1/2\displaystyle\lesssim\Big{(}\int_{(t,x)\in A_{r}}\int_{v\in Supp(\Psi_{2})}|u_{R_{n}}[\overline{f}^{n}]|^{2}\overline{f}^{n}dxdt\Big{)}^{1/2}\Big{(}\int_{A_{r}}\overline{\rho}^{n}dxdt\Big{)}^{1/2} (3.65)

so that, with Proposition 3.2,

𝔼¯[Ar(ω)|J¯n|dxdt]𝔼[0T2d|v]|2fRndzdt]1/2𝔼¯[Ar(ω)ρ¯ndxdt]1/2𝔼¯[Ar(ω)ρ¯ndxdt]1/2.\displaystyle\overline{\mathbb{E}}\Big{[}\int_{A_{r}(\omega)}|\overline{J}^{n}|dxdt\Big{]}\lesssim\mathbb{E}\Big{[}\int_{0}^{T}\int_{\mathbb{R}^{2d}}|v]|^{2}f^{R_{n}}dzdt\Big{]}^{1/2}\overline{\mathbb{E}}\Big{[}\int_{A_{r}(\omega)}\overline{\rho}^{n}dxdt\Big{]}^{1/2}\lesssim\overline{\mathbb{E}}\Big{[}\int_{A_{r}(\omega)}\overline{\rho}^{n}dxdt\Big{]}^{1/2}.

From (3.61) and Lemma 3.3, we deduce

𝔼¯[0TAr(ω)|J¯|𝑑x𝑑t]𝔼¯[0T{xB(0,r)|ϕρ¯=0}ρ¯𝑑x𝑑t]1/2=0.\displaystyle\overline{\mathbb{E}}\Big{[}\int_{0}^{T}\int_{A_{r}(\omega)}|\overline{J}|dxdt\Big{]}\lesssim\overline{\mathbb{E}}\Big{[}\int_{0}^{T}\int_{\{x\in B(0,r)\;|\;\phi*\overline{\rho}=0\}}\overline{\rho}dxdt\Big{]}^{1/2}=0.

Indeed, it is easy to see that, when ρC(d)\rho\in C(\mathbb{R}^{d}),

{xB(0,r)|ϕρ¯=0}ρ(x)𝑑x=0\int_{\{x\in B(0,r)\;|\;\phi*\overline{\rho}=0\}}\rho(x)dx=0

and the same equality is deduced for any ρL2(d)\rho\in L^{2}(\mathbb{R}^{d}) by density. Hence, ¯\overline{\mathbb{P}} almost surely, J¯=0\overline{J}=0 a.e on ArA_{r}. Since this holds for any r>0r>0, we deduce that, ¯\overline{\mathbb{P}} almost surely, J¯=0\overline{J}=0 holds whenever ϕρ¯=0\phi*\overline{\rho}=0, so that we only have to check equality (3.64) whenever (ϕρ¯)(t,x)0(\phi*\overline{\rho})(t,x)\neq 0.

Recalling (3.47) and Lemma 3.3, we have

J¯n=ϕρ¯θRnnRn1+ϕρ¯nρ¯Ψ2nϕj¯ϕρ¯ρ¯Ψ2=u¯ρΨ2\displaystyle\overline{J}^{n}=\frac{\phi*\overline{\rho}^{n}_{\theta_{R_{n}}}}{R_{n}^{-1}+\phi*\overline{\rho}^{n}}\cdot\overline{\rho}^{n}_{\nabla\Psi_{2}}\to\frac{\phi*\overline{j}}{\phi*\overline{\rho}}\cdot\overline{\rho}_{\nabla\Psi_{2}}=\overline{u}\cdot\rho_{\nabla\Psi_{2}}

almost everywhere on the set {(t,x)[0,T]×d|ϕρ¯0}\{(t,x)\in[0,T]\times\mathbb{R}^{d}\;|\;\phi*\overline{\rho}\neq 0\}, ¯\overline{\mathbb{P}} almost surely. It follows that equality (3.64) holds.

Finally, it remains to pass (3.58) to the limit in the quadratic equality (3.54). To this intent, we simply notice that

|stxϕρ¯θRnnRn1+ϕρ¯nρ¯Ψ2nΨ1𝑑x𝑑σ|2=σ1σ2xyJ¯n(σ1,x)J¯n(σ2,y)Ψ1(x)Ψ2(y)𝑑x𝑑y𝑑σ1𝑑σ2\displaystyle\Big{|}\int_{s}^{t}\int_{x}\frac{\phi*\overline{\rho}^{n}_{\theta_{R_{n}}}}{R_{n}^{-1}+\phi*\overline{\rho}^{n}}\cdot\overline{\rho}^{n}_{\nabla\Psi_{2}}\Psi_{1}dxd\sigma\Big{|}^{2}=\int_{\sigma_{1}}\int_{\sigma_{2}}\int_{x}\int_{y}\overline{J}^{n}(\sigma_{1},x)\overline{J}^{n}(\sigma_{2},y)\Psi_{1}(x)\Psi_{2}(y)dxdyd\sigma_{1}d\sigma_{2}

where J¯n(σ,x)\overline{J}^{n}(\sigma,x) is defined as in (3.59). We have the uniform bound

𝔼¯[|supσ1[0,T]supσ2[0,T]xy|J¯n(σ1,x)J¯n(σ2,y)|qdxdy|1/2]=𝔼¯[|supσ[0,T]x|J¯n(σ,x)|qdx]1\overline{\mathbb{E}}\Big{[}\Big{|}\sup_{\sigma_{1}\in[0,T]}\sup_{\sigma_{2}\in[0,T]}\int_{x}\int_{y}\Big{|}\overline{J}^{n}(\sigma_{1},x)\overline{J}^{n}(\sigma_{2},y)\Big{|}^{q}dxdy\Big{|}^{1/2}\Big{]}=\overline{\mathbb{E}}\Big{[}\Big{|}\sup_{\sigma\in[0,T]}\int_{x}\Big{|}\overline{J}^{n}(\sigma,x)\Big{|}^{q}dx\Big{]}\lesssim 1

with q/2>1q/2>1 so that, up to some subsequence

J¯n(σ1,x)J¯n(σ2,y)J¯(σ1,σ2,x,y) weak  in Lq/2(Ω¯;L([0,T]2;Lq((d)2))).\displaystyle\overline{J}^{n}(\sigma_{1},x)\overline{J}^{n}(\sigma_{2},y)\rightharpoonup\overline{J}(\sigma_{1},\sigma_{2},x,y)\text{ weak $*$ in }L^{q/2}\Big{(}\overline{\Omega};L^{\infty}([0,T]^{2};L^{q}((\mathbb{R}^{d})^{2}))\Big{)}.

This time, thanks to (3.60)

𝔼¯[(x,σ1)Ar(y,σ2)B(0,r)×[0,T]|J¯n(σ1,x)J¯n(σ2,y)|𝑑x𝑑σ1𝑑y𝑑σ2]\displaystyle\overline{\mathbb{E}}\Big{[}\int_{(x,\sigma_{1})\in A_{r}}\int_{(y,\sigma_{2})\in B(0,r)\times[0,T]}\Big{|}\overline{J}^{n}(\sigma_{1},x)\overline{J}^{n}(\sigma_{2},y)\Big{|}dxd\sigma_{1}dyd\sigma_{2}\Big{]}
𝔼¯[(Ar|J¯n|𝑑x𝑑t)2]1/2𝔼¯[0TB(0,r)|J¯n|2𝑑x𝑑t]1/2𝔼¯[(Ar|J¯n|𝑑x𝑑t)2]1/2.\displaystyle\hskip 56.9055pt\lesssim\overline{\mathbb{E}}\Big{[}\Big{(}\int_{A_{r}}\Big{|}\overline{J}^{n}\Big{|}dxdt\Big{)}^{2}\Big{]}^{1/2}\overline{\mathbb{E}}\Big{[}\int_{0}^{T}\int_{B(0,r)}\Big{|}\overline{J}^{n}\Big{|}^{2}dxdt\Big{]}^{1/2}\lesssim\overline{\mathbb{E}}\Big{[}\Big{(}\int_{A_{r}}\Big{|}\overline{J}^{n}\Big{|}dxdt\Big{)}^{2}\Big{]}^{1/2}.

Using (3.65), it follows that

𝔼¯[(Ar|J¯n|𝑑x𝑑t)2]\displaystyle\overline{\mathbb{E}}\Big{[}\Big{(}\int_{A_{r}}\Big{|}\overline{J}^{n}\Big{|}dxdt\Big{)}^{2}\Big{]} 𝔼[0T2d|uRn[fRn]|4|fRn|2𝑑z𝑑t]1/2𝔼¯[Ar|ρ¯n|2𝑑x𝑑t]1/2\displaystyle\lesssim\mathbb{E}\Big{[}\int_{0}^{T}\int_{\mathbb{R}^{2d}}|u_{R_{n}}[f^{R_{n}}]|^{4}|f^{R_{n}}|^{2}dzdt\Big{]}^{1/2}\overline{\mathbb{E}}\Big{[}\int_{A_{r}}|\overline{\rho}^{n}|^{2}dxdt\Big{]}^{1/2}
𝔼[0T2d|v|kfRn+|fRn|pdzdt]1/2𝔼¯[Ar|ρ¯n|2𝑑x𝑑t]1/2\displaystyle\lesssim\mathbb{E}\Big{[}\int_{0}^{T}\int_{\mathbb{R}^{2d}}|v|^{k}f^{R_{n}}+|f^{R_{n}}|^{p}dzdt\Big{]}^{1/2}\overline{\mathbb{E}}\Big{[}\int_{A_{r}}|\overline{\rho}^{n}|^{2}dxdt\Big{]}^{1/2}
𝔼¯[Ar|ρ¯n|2𝑑x𝑑t]1/2\displaystyle\lesssim\overline{\mathbb{E}}\Big{[}\int_{A_{r}}|\overline{\rho}^{n}|^{2}dxdt\Big{]}^{1/2}

Using again Lemma 3.3, we deduce that, ¯\overline{\mathbb{P}} almost surely, J¯=0\overline{J}=0 almost everywhere on

{(σ1,σ2,x,y)[0,T]2×(d)2|(ϕρ¯)(σ1,x)=0 or (ϕρ¯)(σ2,y)=0}.\left\{(\sigma_{1},\sigma_{2},x,y)\in[0,T]^{2}\times(\mathbb{R}^{d})^{2}\;|\;(\phi*\overline{\rho})(\sigma_{1},x)=0\text{ or }(\phi*\overline{\rho})(\sigma_{2},y)=0\right\}.

The limit is then determined to be J¯(σ1,σ2,x,y)=(u¯(x)ρΨ2(σ1,x))(u¯(y)ρΨ2(σ2,y))\overline{J}(\sigma_{1},\sigma_{2},x,y)=\Big{(}\overline{u}(x)\cdot\rho_{\nabla\Psi_{2}}(\sigma_{1},x)\Big{)}\Big{(}\overline{u}(y)\cdot\rho_{\nabla\Psi_{2}}(\sigma_{2},y)\Big{)} on the complementary set using pointwise convergence, as done previously.

From the martingale problem of Proposition 3.7, we classically construct a martingale solution of (1.7), in the sense of Definition 1.1, using a martingale representation theorem. First, note that estimates (1.13) and (1.14) are easily derived from the convergence (3.45) using Fatou’s Lemma. Introducing the process

M¯(t)=f¯(t)f00t([f¯(s)])f¯(s)𝑑s,t[0,T]\displaystyle\overline{M}(t)=\overline{f}(t)-f_{0}-\int_{0}^{t}({\cal L}[\overline{f}(s)])^{*}\overline{f}(s)ds,\;\;\;t\in[0,T] (3.66)

we see that, for all test function ΨCc(2d)\Psi\in C^{\infty}_{c}(\mathbb{R}^{2d}) of the form (3.52),

<M¯(t),Ψ>=M¯Ψ(t),t[0,T]\Bigl{<}\overline{M}(t),\Psi\Bigr{>}=\overline{M}_{\Psi}(t),\;\;\;t\in[0,T]

which is a continuous L2L^{2} martingale with respect to the filtration (¯t)t[0,T](\overline{\cal F}_{t})_{t\in[0,T]}, with quadratic variation V¯Ψ\overline{V}_{\Psi}. With Remark 3.2 in mind, by density, we may carefully extend this statement to any test function Ψ\Psi in some separable Hilbert space {\cal H}. One may convince oneself for instance that the weighted Sobolev space

={ΨD(2d),max|β|22de|z||zβΨ(z)|2𝑑z}{\cal H}=\left\{\Psi\in D^{\prime}(\mathbb{R}^{2d}),\;\;\max_{|\beta|\leq 2}\int_{\mathbb{R}^{2d}}e^{|z|}|\partial^{\beta}_{z}\Psi(z)|^{2}dz\right\}

is suitable. Using a polarization formula, we deduce that for Ψ1,Ψ2\Psi_{1},\Psi_{2}\in{\cal H},

<M¯(t),Ψ1><M¯(t),Ψ2><V¯(t)Ψ1,Ψ2>,t[0,T]\Bigl{<}\overline{M}(t),\Psi_{1}\Bigr{>}\Bigl{<}\overline{M}(t),\Psi_{2}\Bigr{>}-\Bigl{<}\overline{V}(t)\Psi_{1},\Psi_{2}\Bigr{>},\;\;\;t\in[0,T]

defines a continuous (¯t)t[0,T](\overline{\cal F}_{t})_{t\in[0,T]}-martingale, where the operator V¯(t)\overline{V}(t) is defined through

<V¯(t)Ψ1,Ψ2>=0t<K[f¯(s)]vΨ1,f¯(s)><K[f¯(s)]vΨ2,f¯(s)>ds.\Bigl{<}\overline{V}(t)\Psi_{1},\Psi_{2}\Bigr{>}=\int_{0}^{t}\Bigl{<}K[\overline{f}(s)]\cdot\nabla_{v}\Psi_{1},\overline{f}(s)\Bigr{>}\Bigl{<}K[\overline{f}(s)]\cdot\nabla_{v}\Psi_{2},\overline{f}(s)\Bigr{>}ds.

The martingale representation theorem from [8] (p222, Theorem 9.2) holds for the {\cal H}^{\prime}-valued process (3.66), giving another probability space (Ω^,^,^)(\hat{\Omega},\hat{\cal F},\hat{\mathbb{P}}) equipped with a filtration (^t)t[0,T](\hat{\cal F}_{t})_{t\in[0,T]}, and a (¯t×^t)(\overline{\cal F}_{t}\times\hat{\cal F}_{t})-brownian motion (Wt)t[0,T](W_{t})_{t\in[0,T]} on Ω¯×Ω^\overline{\Omega}\times\hat{\Omega} such that

M¯(t)(ω¯,ω^):=M¯(t)(ω¯)=0tv(K[f¯(s)]f¯(s))𝑑Ws(ω¯,ω^).\overline{M}(t)(\overline{\omega},\hat{\omega}):=\overline{M}(t)(\overline{\omega})=-\int_{0}^{t}\nabla_{v}\cdot\Big{(}K[\overline{f}(s)]\overline{f}(s)\Big{)}dW_{s}(\overline{\omega},\hat{\omega}).

It follows that (ω¯,ω^)Ω¯×Ω^(f¯(t)(ω¯))t[0,T](\overline{\omega},\hat{\omega})\in\overline{\Omega}\times\hat{\Omega}\mapsto(\overline{f}(t)(\overline{\omega}))_{t\in[0,T]} defines a solution of (1.7) on Ω¯×Ω^\overline{\Omega}\times\hat{\Omega}. This concludes the proof of Theorem 1.

3.5 Strong local alignment

The proof of Theorem 2 can be established using the same arguments. Indeed, the only role played by the weight function ϕ\phi in the estimates of sections 3.1 and 3.2 is through the constant C(ϕ)C(\phi) in Proposition 3.2, given by (3.9). Introducing ϕr\phi_{r} of the form ϕr(x)=rdϕ1(x/r)\phi_{r}(x)=r^{d}\phi_{1}(x/r) we notice that

C(ϕr)rdsupB(0,r2)ϕ1infB(0,r1)ϕ1(r2r1×r)d=supB(0,r2)ϕ1infB(0,r1)ϕ1(r2r1)dC(ϕ1)C(\phi_{r})\propto\frac{r^{d}\sup_{B(0,r_{2})}\phi_{1}}{\inf_{B(0,r_{1})}\phi_{1}}\Big{(}\frac{r_{2}}{r_{1}\times r}\Big{)}^{d}=\frac{\sup_{B(0,r_{2})}\phi_{1}}{\inf_{B(0,r_{1})}\phi_{1}}\Big{(}\frac{r_{2}}{r_{1}}\Big{)}^{d}\propto C(\phi_{1})

uniformly in r>0r>0. As a result, introducing frf^{r} a weak solution of (1.7) with ϕ=ϕr\phi=\phi^{r} constructed (in law) as previously, the following estimates hold uniformly in r>0r>0: for any φCc(d)\varphi\in C_{c}^{\infty}(\mathbb{R}^{d}),

𝔼[supt[0,T]2d|fr(t,z)|p+(1+|x|2+|v|k)fr(t,z)dz]+𝔼[ρφrLt2Hxη2]1\displaystyle\mathbb{E}\Big{[}\sup_{t\in[0,T]}\int_{\mathbb{R}^{2d}}|f^{r}(t,z)|^{p}+(1+|x|^{2}+|v|^{k})f^{r}(t,z)dz\Big{]}+\mathbb{E}\Big{[}\|\rho^{r}_{\varphi}\|^{2}_{L^{2}_{t}H^{\eta}_{x}}\Big{]}\lesssim 1

and for some q>1q>1,

E[fr(t)fr(s)H22q]+𝔼[ρr(t)ρr(s)H22q]|ts|q.\displaystyle E\Big{[}\|f^{r}(t)-f^{r}(s)\|^{2q}_{H^{-2}}\Big{]}+\mathbb{E}\Big{[}\|\rho^{r}(t)-\rho^{r}(s)\|^{2q}_{H^{-2}}\Big{]}\leq|t-s|^{q}.

From here, we may use the same arguments as previously to establish the tightness of (fr)r>0(f^{r})_{r>0} (resp. (ρφr)r>0(\rho^{r}_{\varphi})_{r>0}) in C([0,T];HW1σ(2d))C([0,T];H^{-\sigma}_{W^{-1}}(\mathbb{R}^{2d})) for σ>0\sigma>0 (resp. in L2([0,T];L2(d))L^{2}([0,T];L^{2}(\mathbb{R}^{d}))) and then pass to the limit in the martingale problem satisfied by frnf^{r_{n}} as nn goes to infinity.

References

  • [1] S. Ahn and S. Ha. Stochastic flocking dynamics of the cucker–smale model with multiplicative white noises. Journal of Mathematical Physics, 51(10):103301, 2010.
  • [2] F. Bouchut and L. Desvillettes. Averaging lemmas without time fourier transform and application to discretized kinetic equations. Proceedings of the Royal Society of Edinburgh: Section A Mathematics, 129(1):19–36, 1999.
  • [3] D. Burkholder, B. Davis, and R. Gundy. Integral inequalities for convex functions of operators on martingales. In Proceedings of the Sixth Berkeley Symposium on Mathematical Statistics and Probability, Volume 2: Probability Theory, pages 223–240, Berkeley, Calif., 1972. University of California Press.
  • [4] Y. Choi and S. Salem. Cucker-smale flocking particles with multiplicative noises: Stochastic mean-field limit and phase transition. Kinetic & Related Models, 12, 11 2017.
  • [5] M. Coghi and F. Flandoli. Propagation of chaos for interacting particles subject to environmental noise. The Annals of Applied Probability, 26, 03 2014.
  • [6] F. Cucker and S. Smale. Emergent behavior in flocks. IEEE Transactions on Automatic Control, 52(5):852–862, 2007.
  • [7] F. Cucker and S. Smale. On the mathematics of emergence. Japanese Journal of Mathematics, 2(1):197–227, 2007.
  • [8] G. Da Prato and J. Zabczyk. Stochastic Equations in Infinite Dimensions. Cambridge University Press, Cambridge, 2014.
  • [9] Arnaud Debussche, Sylvain De Moor, and Julien Vovelle. Diffusion limit for the radiative transfer equation perturbed by a Wiener process. Kinetic and Related Models , 8(3):467–492, 2015. 27 pages.
  • [10] S. Ha, J. Jung, and M. Röckner. Collective stochastic dynamics of the cucker-smale ensemble under uncertain communications. https://arxiv.org/abs/1903.07771, 03 2019.
  • [11] T. Karper, A. Mellet, and K. Trivisa. On strong local alignment in the kinetic cucker-smale model. Springer Proceedings in Mathematics & Statistics, 02 2012.
  • [12] T. Karper, A. Mellet, and K. Trivisa. Existence of weak solutions to kinetic flocking models. SIAM Journal on Mathematical Analysis, 45(1):215–243, 2013.
  • [13] H. Kunita. Stochastic differential equations and stochastic flows of diffeomorphisms. In P. L. Hennequin, editor, École d’Été de Probabilités de Saint-Flour XII - 1982, pages 143–303, Berlin, Heidelberg, 1984. Springer Berlin Heidelberg.
  • [14] H. Kunita. Stochastic Flows and Stochastic Differential Equations. Cambridge University Press, Cambridge, 1990.
  • [15] Y. Liu and X. Xue. The motsch-tadmor model with multiplicative white noises in flocks. Dynamic Systems and Applications, 24:503–522, 12 2015.
  • [16] S. Motsch and E. Tadmor. A new model for self-organized dynamics and its flocking behavior. Journal of Statistical Physics, 144(5):923–947, Aug 2011.
  • [17] Sebastien Motsch and Eitan Tadmor. Heterophilious dynamics enhances consensus. SIAM Review, 56(4):577–621, 2014.
  • [18] A. Rosello. Weak and strong mean-field limits for stochastic cucker-smale particle systems. https://arxiv.org/abs/1905.02499, 2019.
  • [19] H. Triebel. Theory of Function Spaces. Modern Birkhäuser Classics. Birkhäuser Basel, 1983.
  • [20] C. Villani. Optimal Transport: Old and New. Springer, 2009 edition, 2008.