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

Sample-path large deviation principle for a 2-D stochastic interacting vortex dynamics with singular kernel

Chenyang Chenlabel=e1][email protected] [    Hao Gelabel=e3 [    mark][email protected] Beijing International Center for Mathematical Research, Peking University
Abstract

We consider a stochastic interacting vortex system of NN particles, approximating the vorticity formulation of 2-D Navier-Stokes equation on torus. The singular interaction kernel is given by the Biot-Savart law. We only require the initial state to have finite energy, and obtain a sample-path large deviation principle for the empirical measure when the number of vortices goes to infinity. The rate function is characterized by an explicit formula supporting on sample paths with finite energy and finite integral of L2L^{2} norms over time. The proof utilizes a symmetrization technique for the representation of singular kernel, together with a detailed regularity analysis of the sample path with finite rate function. The key step is to prove that the singular term after symmetrization can be bounded by the integral of L2L^{2} norms along sample paths.

60K35,
60F10,
76D05,
60H10,
60B10,
2D Navier-Stokes equation,
Stochastic vortex dynamics,
Large deviation principle,
Energy dissipation structure,
keywords:
[class=MSC]
keywords:
\startlocaldefs\endlocaldefs

,

footnotetext:

1 Introduction

The subject of this paper is the sample-path large deviation principle (LDP) for a stochastic interacting vortex dynamics of 2D Navier-Stokes equation starting from initial data with finite energy.

Define

r(x,y):=infk2|xyk|,x,y2.r(x,y):=\inf_{k\in\mathbb{Z}^{2}}|x-y-k|,\quad\forall x,y\in\mathbb{R}^{2}.

Let 𝕋2=2\2\mathbb{T}^{2}=\mathbb{R}^{2}\backslash\mathbb{Z}^{2} be the two dimensional torus, i.e. the unit cube [12,12)2[-\frac{1}{2},\frac{1}{2})^{2} with metric rr, and 𝒩(x)\mathcal{N}(x) be the Green function on 𝕋2\mathbb{T}^{2}, i.e.

Δ𝒩(x)=δ0(x)1,𝕋2𝒩(x)𝑑x=0,-\Delta\mathcal{N}(x)=\delta_{0}(x)-1,\quad\int_{\mathbb{T}^{2}}\mathcal{N}(x)dx=0, (1.1)

in which δ0\delta_{0} means Dirac measure with singularity at 0. We define a nn-particle stochastic vortex dynamics by stochastic differential equations on 2D torus for any given ν>0\nu>0:

dXi(t)=1nji𝒦(Xi(t)Xj(t))dt+2νdBi(t),i=1,2,,n,dX_{i}(t)=\frac{1}{n}\sum_{j\neq i}\mathcal{K}(X_{i}(t)-X_{j}(t))dt+\sqrt{2\nu}dB_{i}(t),\;i=1,2,\cdots,n,\\ (1.2)

with random initial data {Xi(0)}i=1,,n\{X_{i}(0)\}_{i=1,\cdots,n}, independent of the family (Bi(t))i=1,,n(B_{i}(t))_{i=1,\cdots,n} of i.i.d.i.i.d. 2D-Brownian motions, where the Biot-Savart kernel 𝒦\mathcal{K} is given by 𝒦=𝒩=(2𝒩,1𝒩)\mathcal{K}=-\nabla^{\perp}\mathcal{N}=(-\partial_{2}\mathcal{N},\partial_{1}\mathcal{N}). Supposing the initial energy, i.e.

12n2ij𝒩(Xi(0)Xj(0)),\frac{1}{2n^{2}}\sum_{i\neq j}\mathcal{N}(X_{i}(0)-X_{j}(0)),

is finite, we are interested in the sample-path LDP of the empirical measure process ρn:=ρn(t)=1ni=1nδXi(t)\rho_{n}:=\rho_{n}(t)=\frac{1}{n}\sum_{i=1}^{n}\delta_{X_{i}(t)} when nn goes to ++\infty.

The model without noise, i.e. ν=0\nu=0, was originally introduced by Helmholtz [27] and then studied by Kirchhoff [34], describing the motion of vortices. Schochet [46] proved its convergence to the periodic weak solutions of Euler equation. (1.2) with ν=0\nu=0 is exactly a Hamiltonian system with the singular potential 𝒩\mathcal{N} and is closely related to Onsager’s theory [41] for two dimensional fluids, which was latter developed by Joyce and Montgomery [30]. The associated thermodynamic limit of microcanonical/canonical distributions was made rigorous by Caglioti, Lions, Marchioro and Pulvirenti [9, 10] as well as Eyink and Spohn [19] in 1990s. The idea behind these theories is exactly LDP for a sequence of distributions. Since then, from the perspective of equilibrium statistical physics, the large deviation principle for more general Gibbs measures with nonsingular and singular Hamiltonians have be extensively studied [2, 26, 4, 36, 17].

It’s well known that the empirical measure process {ρn(t)}\{\rho_{n}(t)\} of the corresponding model in 2\mathbb{R}^{2} converges to the vorticity form of two dimensional Navier-Stokes equations as nn\rightarrow\infty.

Consider the incompressible Navier-Stokes equations on 2\mathbb{R}^{2}:

tu+(u)u+p=νΔu,divu=0,\displaystyle\partial_{t}u+(u\cdot\nabla)u+\nabla p=\nu\Delta u,\quad{\rm div}\,u=0, (1.3)
u(0,x)=u0,divu0=0,\displaystyle u(0,x)=u_{0},\quad{\rm div}\,u_{0}=0,

Its vorticity ρ(t):=curlu(t)=2u11u2\rho(t):={\rm curl}\,u(t)=\partial_{2}u_{1}-\partial_{1}u_{2} satisfies a continuity equation

tρ+div(ρu)νΔρ=0,\partial_{t}\rho+{\rm div}\,(\rho u)-\nu\Delta\rho=0, (1.4)

with a velocity field driven by the vorticity itself u(t):=𝒦ρ(t)u(t):=\mathcal{K}*\rho(t), in which 𝒦(x)=12π|x|2(x2,x1)\mathcal{K}(x)=\frac{1}{2\pi|x|^{2}}(-x_{2},x_{1}) for the case of 2\mathbb{R}^{2}. (1.4) is called the voricity form of two-dimensional Navier-Stokes equations.

Osada [42, 43] investigated the well posedness of (1.2) in 2\mathbb{R}^{2} and obtained a propagation of chaos for the equation (1.4) for large ν\nu. For general positive ν\nu, using the cutoff kernels 𝒦n\mathcal{K}_{n} converging to the original one with singularity, Marchioro and Pulvirenti [37] and Méléard [39] also proved propagation of chaos. Fournier, Hauray and Mischler [24] proved a stronger ”entropic chaos” result not requiring ν\nu large and without cutoff when the distributions of initial data are regular enough.

Back to 𝕋2\mathbb{T}^{2}, the well posedness for this SDEs with the singular kernel 𝒦\mathcal{K} is firstly studied in [15]. In this case, as a classical result, 𝒩\mathcal{N} are smooth even function on 𝕋2\(0,0)\mathbb{T}^{2}\backslash(0,0) and

𝒩(x)=12πlog(|x|)+𝒩0(x),\mathcal{N}(x)=-\frac{1}{2\pi}\log(|x|)+\mathcal{N}_{0}(x),

where 𝒩0\mathcal{N}_{0} is a bounded correction to periodize 𝒩\mathcal{N} on 𝕋2\mathbb{T}^{2}. In our paper, we need a refined result that

𝒩(x)=ψ(x)2πlog|x|+σ1(x),\displaystyle\mathcal{N}(x)=-\frac{\psi(x)}{2\pi}\log|x|+\sigma_{1}(x), (1.5)
𝒦(x)=ψ(x)2π|x|2(x2,x1)+(σ2(x),σ3(x)),\displaystyle\mathcal{K}(x)=\frac{\psi(x)}{2\pi|x|^{2}}(-x_{2},x_{1})+(\sigma_{2}(x),\sigma_{3}(x)),

for certain smooth function 0ψ10\leq\psi\leq 1 satisfying

ψ(x)={1,|x|<14,0,|x|>13,\psi(x)=\left\{\begin{aligned} &1,&|x|<\frac{1}{4},\\ &0,&|x|>\frac{1}{3},\end{aligned}\right.

and σiC(𝕋2)\sigma_{i}\in C^{\infty}(\mathbb{T}^{2}). The method of [24] also works and the propagation of chaos holds for (1.4) on 𝕋2\mathbb{T}^{2}. The slight difference is that ρ(t):=curlu(t)+1\rho(t):={\rm curl}\,u(t)+1 instead of curlu(t){\rm curl}\,u(t) in the case of 𝕋2\mathbb{T}^{2} in which u(t)u(t) is the solution of Navier Stokes equation (1.3) on 𝕋2\mathbb{T}^{2}.

1.1 Weakly interacting stochastic particle system with singular kernel

The 2-D vortex dynamics (1.2) is a weakly interacting stochastic particle model with independent noise. It is called ”weak” because when the number of particles tends to infinity, the force between any two particles tends to zero.

For the mean-field limit, the case in the presence of smooth interaction kernels has been well studied since McKean’s work [38]. But there is only few references in the case of singular interactions for a long time. Recently, based on the estimation for relative entropy, Jabin and Wang [28] proved the mean-field limit and obtained the optimal convergent rate of distributions for a general class of singular kernels including the Biot-Savart law. Meanwhile, based on modulated energy, Serfaty [47] and Duernckx [16] proved the mean-field limit for deterministic systems with more singular kernels. Combining the two methods, Bresch, Jabin and Wang [6] derived the mean-field limit for the Patlak-Keller-Segel kernel and gave the corresponding convergent rate.

To quantify the difference between empirical measure processes and mean-field limit more precisely, some considered the fluctuation in path space, usually called the central limit theorem (CLT) for Gaussian fluctuations. Results of this type were firstly obtained by Tanaka and Hitsuda [50] and by Tanaka [49] for one dimensional non-singular interacting diffusions. For more general case with regular enough coefficients, it was proved by Fernandez and Méléard [22]. Recently, Wang, Zhao, and Zhu [53] proved CLT for a special class of singular interaction including (1.2).

A further topic is the sample-path large deviation principle for empirical measures. For non-singular interaction kernels only in drift coefficients, Dawson and Gärtner [11] proved the sample-path LDP for general weakly interacting stochastic particles model using measure transformation from independent diffusions. Budhiraja, Dupuis and Fischer [8] further developed this result, allowing the non-singular interaction kernels in both drift and diffusion terms. The method by Feng and Kurtz, based on Hamiltonian-Jocabi theory, also works for smooth kernel (see Section 13.3 of [20]).

For singular interaction kernels, Fontbona [23] proved the sample-path LDP for diffusing particles with electrostatic repulsion in one dimension. Applying the method in [20], at least formally, Feng and Świech [21] gave the rate function of sample-path LDP for empirical measures of (1.2), starting from initial data with finite entropy. However, making this approach rigorous is really subtle, requiring an uniqueness theory for infinite-dimensional Hamiltonian-Jocabi equation, especially when only having initial data with finite energy in hand.

In the present paper, we aim to study the sample-path large deviation principle of (1.2) starting from more rough initial data only with finite energy. We follow the standard approach used in [11, 33, 14], i.e., applying an exponential martingale inequality to prove the upper bound and approximating trajectories with finite action (rate function) by a series of mildly perturbed systems for the proof of lower bound. The main difficulty here comes from the term in the rate function containing the singular kernel 𝒦\mathcal{K}.

We use the idea from [44, 45, 5], introducing an auxiliary functional as a modification of the rate function to make the approximating strategy for the proof of lower bound work. Moreover, during the proof of large deviation upper bound, this auxiliary functional ensures the continuity of singular terms in the Markov generator with respect to weak topology, which is not a problem for the case of non-singular kernels.

As a price, we need a sharp prior estimation for the auxiliary functional for the upper bound establishment. We finally choose the supremum of energy functional and the integral of L2L^{2} norm along time as the auxiliary functional in our paper. Both in the procedure of nice trajectory approximation during the proof of lower bound, and in the prior estimation of auxiliary functional during the proof of upper bound, the singular terms are shown to be bound by the integral of L2L^{2} norm over time, which is our key observation and main tools throughout this paper.

1.2 Notations and definitions

We write C(Ω)C^{\infty}(\Omega) and 𝒟(Ω)\mathcal{D}^{\prime}(\Omega) for the collection of infinitely differentiable functions on Ω\Omega and generalized functions on Ω\Omega.

We use ,\left\langle\cdot,\cdot\right\rangle to represent pairing of linear space and its dual space, such as L2L^{2} inner product, continuous function integral with respect to a measure, and so on.

Denote by 𝒫(𝕋2)\mathcal{P}(\mathbb{T}^{2}) the space of probability measures on 𝕋2\mathbb{T}^{2}. For γ,η𝒫(𝕋2),\gamma,\eta\in\mathcal{P}(\mathbb{T}^{2}), we define the Wasserstein2-2 metric

d(γ,η):=infπΠ(γ,η)𝕋2×𝕋2r(x,y)2π(dx,dy),d(\gamma,\eta):=\inf_{\pi\in\Pi(\gamma,\eta)}\int_{\mathbb{T}^{2}\times\mathbb{T}^{2}}r(x,y)^{2}\pi(dx,dy), (1.6)

where

Π(γ,η)={π𝒫(𝕋2×𝕋2):π(dx,𝕋2)=γ(dx),π(𝕋2,dy)=η(dy)}.\Pi(\gamma,\eta)=\{\pi\in\mathcal{P}(\mathbb{T}^{2}\times\mathbb{T}^{2}):\pi(dx,\mathbb{T}^{2})=\gamma(dx),\pi(\mathbb{T}^{2},dy)=\eta(dy)\}.

Then (𝒫(𝕋2),d)(\mathcal{P}(\mathbb{T}^{2}),d) is a complete separable metric space. See Chapter 7 of [1] or Chapter 7 of [52] for properties of this metric. Given μ𝒫(𝕋2)\mu\in\mathcal{P}(\mathbb{T}^{2}), for any m𝒟(𝕋2)m\in\mathcal{D}^{\prime}(\mathbb{T}^{2}) we define

m1,μ2=supϕC(𝕋2){2ϕ,m𝕋2|ϕ(x)|2𝑑μ}.\|m\|_{-1,\mu}^{2}=\sup_{\phi\in C^{\infty}(\mathbb{T}^{2})}\left\{2\left\langle\phi,m\right\rangle-\int_{\mathbb{T}^{2}}|\nabla\phi(x)|^{2}d\mu\right\}. (1.7)

Let AC((s,t);𝒫(𝕋2))AC((s,t);\mathcal{P}(\mathbb{T}^{2})) be the set of all ρC([s,t];𝒫(𝕋2))\rho\in C([s,t];\mathcal{P}(\mathbb{T}^{2})) satisfying that there exists mL1[s,t]m\in L^{1}[s,t] such that

d(ρ(p),ρ(q))pqm(r)𝑑r,s<pq<t.d(\rho(p),\rho(q))\leq\int_{p}^{q}m(r)dr,\quad\forall s<p\leq q<t.

We refer to smooth mollifier throughout this paper as a non-negative even function that is smooth and its integral over 𝕋2\mathbb{T}^{2} is one.

1.3 Statement of the main result

We introduce the state space

𝒳n:={ρ=1ni=1nδxi:x1,,xn𝕋2,xixj,1i<jn}.\mathcal{X}_{n}:=\left\{\rho=\frac{1}{n}\sum_{i=1}^{n}\delta_{x_{i}}:x_{1},\cdots,x_{n}\in\mathbb{T}^{2},x_{i}\neq x_{j},\forall 1\leq i<j\leq n\right\}.

As shown by [15], for the case Xi(0)=xiX_{i}(0)=x_{i}, the system (1.2) is well-posed for almost every (xi)i=1,,n(x_{i})_{i=1,\cdots,n}, but without an explicit characterization. Actually, we need a well-posed result like Theorem 2.10 in [24], and [48] gives a probabilistic proof which can be directly applied to the case of torus and lead to the following lemma.

Lemma 1.1.

Consider (Xi(0))i=1,,n(X_{i}(0))_{i=1,\cdots,n} as 𝕋2\mathbb{T}^{2}-valued random variables, independent of the family (Bi(t))i=1,,n(B_{i}(t))_{i=1,\cdots,n} of i.i.d.i.i.d. 2D-Brownian motions. Suppose that P(ρn(0)𝒳n)=1P(\rho_{n}(0)\in\mathcal{X}_{n})=1, then there exists an unique strong solution to (2.1) and

P(inf{t:ρn(t)𝒳n}<)=0.P\left(\inf\left\{t:\rho_{n}(t)\notin\mathcal{X}_{n}\right\}<\infty\right)=0.

Now we introduce the energy functional for empirical measures

e0(γ):=12(𝕋2)2\𝑫𝒩(xy)γ(dx)γ(dy),γn1𝒳n,e_{0}(\gamma):=\frac{1}{2}\int_{(\mathbb{T}^{2})^{2}\backslash\boldsymbol{D}}\mathcal{N}(x-y)\gamma(dx)\gamma(dy),\quad\forall\gamma\in\bigcup_{n\geq 1}\mathcal{X}_{n},

where 𝑫={(x,y)(𝕋2)2:x=y}\boldsymbol{D}=\{(x,y)\in(\mathbb{T}^{2})^{2}:x=y\} is the diagonal set. We impose the following superexponential finite energy condition for initial data.

Condition 1.1.
limRlim supn1nlogP(e0(ρn(0))>R)=.\lim_{R\to\infty}\limsup_{n\to\infty}\frac{1}{n}\log P\left(e_{0}(\rho_{n}(0))>R\right)=-\infty.

The removal of diagonal set is not convenient for estimations, so we also introduce the energy functional containing the diagonal set,

e(γ):=12(𝕋2)2𝒩(xy)γ(dx)γ(dy),γ𝒫(𝕋2).e(\gamma):=\frac{1}{2}\int_{(\mathbb{T}^{2})^{2}}\mathcal{N}(x-y)\gamma(dx)\gamma(dy),\quad\forall\gamma\in\mathcal{P}(\mathbb{T}^{2}).

It’s always well defined since 𝒩\mathcal{N} is bounded below. By Fatou’s Lemma, ee is lower semi-continuous functional under the weak topology, and we left more properties of ee into Appendix C.

Let QT:C([0,T];𝒫(𝕋2))Q_{T}:C([0,T];\mathcal{P}(\mathbb{T}^{2}))\mapsto\mathbb{R} be defined by

QT(ρ)=sup0tT(e(ρ(t))+ν20tρ(s)122𝑑s),Q_{T}(\rho)=\sup_{0\leq t\leq T}\left(e(\rho(t))+\frac{\nu}{2}\int_{0}^{t}\|\rho(s)-1\|_{2}^{2}ds\right),

and for ρAC((0,T);𝒫(𝕋2))\rho\in AC((0,T);\mathcal{P}(\mathbb{T}^{2})) and QT(ρ)<Q_{T}(\rho)<\infty, we define

𝔸T(ρ)=14ν0Ttρ(t)div[ρ(t)(𝒦ρ)(t)]+νΔρ(t)1,ρ(t)2𝑑t.\mathbb{A}_{T}(\rho)=\frac{1}{4\nu}\int_{0}^{T}\|\partial_{t}\rho(t)-{\rm div}\,[\rho(t)(\mathcal{K}*\rho)(t)]+\nu\Delta\rho(t)\|_{-1,\rho(t)}^{2}dt.

To check 𝔸T(ρ)\mathbb{A}_{T}(\rho) is well defined, by Theorem 8.3.1 of [1], for ρAC((0,T);𝒫(𝕋2))\rho\in AC((0,T);\mathcal{P}(\mathbb{T}^{2})), we can define tρ𝒟(𝕋2)\partial_{t}\rho\in\mathcal{D}^{\prime}(\mathbb{T}^{2}) for almost every t[0,T]t\in[0,T]. In addition, QT(ρ)<Q_{T}(\rho)<\infty implies that ρ(t)L2(𝕋2)\rho(t)\in L^{2}(\mathbb{T}^{2}) for almost every tt. As a consequence of (1.5) and Young’s inequality, if γL2(𝕋2)\gamma\in L^{2}(\mathbb{T}^{2}), then 𝒦γL2(𝕋2)\mathcal{K}*\gamma\in L^{2}(\mathbb{T}^{2}), so

ϕ,div[γ(𝒦γ)]:=𝕋2ϕ(x)(𝒦γ)(x)γ(x))dx\left\langle\phi,{\rm div}\,[\gamma(\mathcal{K}*\gamma)]\right\rangle:=-\int_{\mathbb{T}^{2}}\nabla\phi(x)\cdot(\mathcal{K}*\gamma)(x)\gamma(x))dx

is well defined for any smooth function ϕ\phi. Thus div[ρ(t)(𝒦ρ(t))]D(𝕋2){\rm div}\,[\rho(t)(\mathcal{K}*\rho(t))]\in D^{\prime}(\mathbb{T}^{2}) for almost every tt.

Our main result is

Theorem 1.2 (Large deviation principle).

Under the condition of Lemma 1.1, suppose ρn(0)\rho_{n}(0), as random variables on 𝒫(𝕋2)\mathcal{P}(\mathbb{T}^{2}), satisfies a large deviation principle with the rate function 0\mathcal{I}_{0} and Condition 1.1 holds. Then for each T>0T>0, the stochastic empirical process {ρn(t):0tT}n=1,2,\{\rho_{n}(t):0\leq t\leq T\}_{n=1,2,\cdots} satisfies a sample-path large deviation principle in the space C([0,T];𝒫(𝕋2))C([0,T];\mathcal{P}(\mathbb{T}^{2})) with a good rate function (usually called action)

T(ρ)={0(ρ(0))+𝔸T(ρ),ifρAC((0,T);𝒫(𝕋2))andQT(ρ)<,,otherwise,\displaystyle\mathcal{I}_{T}(\rho)=\left\{\begin{aligned} &\mathcal{I}_{0}(\rho(0))+\mathbb{A}_{T}(\rho),&\text{if}~{}\rho\in AC((0,T);\mathcal{P}(\mathbb{T}^{2}))~{}\text{and}~{}Q_{T}(\rho)<\infty,\\ &\infty,&otherwise,\end{aligned}\right. (1.8)

that is equivalent to the following two conditions:

1. (Large deviation upper bound) For any closed set AA in C([0,T];𝒫(𝕋2))C([0,T];\mathcal{P}(\mathbb{T}^{2})),

lim supn1nlogP(ρnA)infρAT(ρ);\displaystyle\limsup_{n\to\infty}\frac{1}{n}\log P(\rho_{n}\in A)\leq-\inf_{\rho\in A}\mathcal{I}_{T}(\rho);

2. (Large deviation lower bound) For any open set GG in C([0,T];𝒫(𝕋2))C([0,T];\mathcal{P}(\mathbb{T}^{2})),

lim infn1nlogP(ρnG)infρGT(ρ).\displaystyle\liminf_{n\to\infty}\frac{1}{n}\log P(\rho_{n}\in G)\geq-\inf_{\rho\in G}\mathcal{I}_{T}(\rho).

1.4 Outline of proof

As in the seminar works [11, 33, 14], the sample-path large deviation upper bound can be obtained through an exponential martingale estimation and exponential tightness. Additional treatment is required in the presence of singular terms, i.e. a prior energy estimation to make sure the generator is continuous under weak typology.

The basic strategy in [33, 14] to prove lower bound is to compute the relative entropy between a model with regular perturbation and the original one, which actually provides a lower bound on the probability of the neighborhood of the mean-field limit path of the perturbed process. However, this strategy for lower bound estimation would fail in general when the rate function is not convex, let alone in the presence of singular interaction kernels, unless there is an additional modification for the rate function, as have been done in [44, 45, 5]. As we have stated in the main theorem, we choose QT<Q_{T}<\infty as the modification, which implies finite energy and finite integral of L2L^{2} norm of the distributions over time.

Besides the modification by QT<Q_{T}<\infty, the proof for the lower bound still needs several new inequalities, which can bound the singular terms such as div[ρ(𝒦ρ)]{\rm div}\,[\rho(\mathcal{K}*\rho)] by the integral of L2L^{2} norms along sample path. These inequalities are originated from an energy dissipation structure similar to the classical result for 2D Navier-Stokes equation (e.g. see (3.20) in [51]), in which the integral of L2L^{2} norms along sample path naturally emerges. We can not expect the L2L^{2} norms to be bounded from above since the initial L2L^{2} norm can be infinite, but the integral of L2L^{2} norm along sample path with finite action is proven to be finite, utilizing such an energy dissipation structure. As one may expect, the finite energy condition is required for the energy dissipation structure, and as far as we know, this is the weakest condition that makes the mean-field limit holds for (1.2) with general ν>0\nu>0.

As stated in [5], the rate function with modification makes it harder to prove the upper bound, requiring a prior estimation of QTQ_{T}. However, QTQ_{T} always takes \infty on the empirical measure ρn𝒳n\rho_{n}\in\mathcal{X}_{n}, even if their limit as nn\to\infty has finite QTQ_{T}. Thus we make use of the convolution method with a sequence of smooth mollifiers. By taking specific mollifiers ζn\zeta_{n} converging to Dirac measure with a suitable slow rate, we yield an energy dissipation structure for ζnρn\zeta_{n}*\rho_{n} assuming e(ζnρn(0))e(\zeta_{n}*\rho_{n}(0)) is finite, which gives the control for QT(ζnρn)Q_{T}(\zeta_{n}*\rho_{n}). The key ingredient in the proof is a sharp estimation for the singular term in order to show its vanishing in the energy dissipation structure. Moreover, the energy dissipation structure, as well as related inequalities, can also help to provide a more explicit formula of the rate function than the variational one directly derived from the exponential martingale inequality.

Furthermore, in order to characterize the limit of infintesimal operators of the SDEs with singular terms, we utilize a symmetrization technique to first represent the limit operator defined on arbitrary probability measure, then perform the regularity analysis, and finally come back to the original representation without symmetrization.

In more detail, take ζ:+:\zeta:\mathbb{R}\mapsto\mathbb{R}^{+}:

ζ(x)={Ce114|x|2,|x|<12,0,|x|12,\zeta(x)=\left\{\begin{aligned} &Ce^{-\frac{1}{1-4|x|^{2}}},&|x|<\frac{1}{2},\\ &0,&|x|\geq\frac{1}{2},\end{aligned}\right.

where the constant is to make 02πrζ(r)𝑑r=1\int_{0}^{\infty}2\pi r\zeta(r)dr=1. ζ(x)\zeta(x) is a non-negative smooth function on \mathbb{R} with a compact support. Then let ζn:2\zeta_{n}:\mathbb{R}^{2}\mapsto\mathbb{R} be the specific smooth mollifiers generated by ζ\zeta and defined by

ζn(x)=mn2ζ(mn|x|),\zeta_{n}(x)=m_{n}^{2}\zeta(m_{n}|x|),

where mn,nmn2m_{n}\uparrow\infty,nm_{n}^{-2}\to\infty. Note that the slow convergent rate of mnm_{n} to infinity is crucial for the prior estimation that we will show later.

In Section 2 we will illustrate the proof for the main theorem under the following superexponential finite energy condition for ζnρn\zeta_{n}*\rho_{n} that we will show later.

Condition 1.2.
limRlim supn1nlogP(e(ζnρn(0))>R)=.\lim_{R\to\infty}\limsup_{n\to\infty}\frac{1}{n}\log P\left(e(\zeta_{n}*\rho_{n}(0))>R\right)=-\infty.
Theorem 1.3 (Large deviation principle II).

Under the condition of Lemma 1.1, suppose ρn(0)\rho_{n}(0) satisfies a large deviation principle on 𝒫(𝕋2)\mathcal{P}(\mathbb{T}^{2}) with the rate function 0\mathcal{I}_{0} and Condition 1.2 holds. Then for each T>0T>0, the stochastic empirical process {ρn(t):0tT}n=1,2,\{\rho_{n}(t):0\leq t\leq T\}_{n=1,2,\cdots} satisfies a sample-path large deviation principle in the space C([0,T];𝒫(𝕋2))C([0,T];\mathcal{P}(\mathbb{T}^{2})) with a good rate function given by (1.8).

To conclude the proof, we will also show that finite e0(ρ(0))e_{0}(\rho(0)) is stronger than finite e(ζnρ(0))e(\zeta_{n}*\rho(0)), i.e. the Condition 1.1 implies the Condition 1.2, in Section 3.

In Section 3, we also prove several crucial inequalities for bounding the singular term divρ(𝒦ρ){\rm div}\,\rho(\mathcal{K}*\rho) under convolution and give the prior estimation of QTQ_{T} for the upper bound proof. Section 4 is about regularities of trajectories with finite rate function, which helps us to express the rate function in a more explicit way. In Section 5, we prove the law of large number for the model with certain regular perturbation and perform the ”nice” trajectory approximation strategy for the lower bound proof.

2 Proof of the main theorem

In this section we present the main proof procedure of Theorem 1.3, and we place all the proofs of lemmas into later sections.

2.1 Exponential martingale problem

Our proof of large deviation upper bound is based on the analyses of the exponential martingale problem.

Recall that the dynamics of X(t)X(t) are determined by the stochastic differential equations

dXi(t)=1nji𝒦(XiXj)dt+2νdBi(t),i=1,2,,n.dX_{i}(t)=\frac{1}{n}\sum_{j\neq i}\mathcal{K}(X_{i}-X_{j})dt+\sqrt{2\nu}dB_{i}(t),\;i=1,2,\cdots,n.\\ (2.1)

By Ito’s formula, for each smooth function ϕ\phi on 𝕋2\mathbb{T}^{2}, we have

d[ϕ,ρn(t)]=1ni=1ndϕ(Xi(t))=1ni=1nϕ(Xi(t))dXi(t)+νni=1nΔϕ(Xi(t))dt\displaystyle d\left[\left\langle\phi,\rho_{n}(t)\right\rangle\right]=\frac{1}{n}\sum_{i=1}^{n}d\phi(X_{i}(t))=\frac{1}{n}\sum_{i=1}^{n}\nabla\phi(X_{i}(t))dX_{i}(t)+\frac{\nu}{n}\sum_{i=1}^{n}\Delta\phi(X_{i}(t))dt
=1n2ijϕ(Xi(t))𝒦(Xi(t)Xj(t))dt+2νni=1nϕ(Xi(t))dBi(t)\displaystyle=\frac{1}{n^{2}}\sum_{i\neq j}\nabla\phi(X_{i}(t))\cdot\mathcal{K}(X_{i}(t)-X_{j}(t))dt+\frac{\sqrt{2\nu}}{n}\sum_{i=1}^{n}\nabla\phi(X_{i}(t))dB_{i}(t)
+νni=1nΔϕ(Xi(t))dt\displaystyle+\frac{\nu}{n}\sum_{i=1}^{n}\Delta\phi(X_{i}(t))dt

To write the generator of ρn\rho_{n}, we need to write the singular term 1n2ijϕ(Xi(t))𝒦(Xi(t)Xj(t))\frac{1}{n^{2}}\sum_{i\neq j}\nabla\phi(X_{i}(t))\cdot\mathcal{K}(X_{i}(t)-X_{j}(t)) as a ρn(t)\rho_{n}(t) dependent term. Here we use the observation from Delort [12], proposing an alternative representation for singular term which can be easily extended to general measure.

Recalling 𝑫\boldsymbol{D} is the diagonal set, noting 𝒦(x)=𝒦(x),\mathcal{K}(x)=-\mathcal{K}(-x),

1n2ijϕ(Xi(t))𝒦(Xi(t)Xj(t))\displaystyle\frac{1}{n^{2}}\sum_{i\neq j}\nabla\phi(X_{i}(t))\cdot\mathcal{K}(X_{i}(t)-X_{j}(t)) (2.2)
=12n2ij[ϕ(Xi(t))ϕ(Xj(t))]𝒦(Xi(t)Xj(t))dt\displaystyle=\frac{1}{2n^{2}}\sum_{i\neq j}\left[\nabla\phi(X_{i}(t))-\nabla\phi(X_{j}(t))\right]\cdot\mathcal{K}(X_{i}(t)-X_{j}(t))dt
=12(𝕋2)2\𝑫(ϕ(x)ϕ(y))𝒦(xy)ρn(t,dx)ρn(t,dy)\displaystyle=\frac{1}{2}\int_{(\mathbb{T}^{2})^{2}\backslash\boldsymbol{D}}(\nabla\phi(x)-\nabla\phi(y))\cdot\mathcal{K}(x-y)\rho_{n}(t,dx)\rho_{n}(t,dy)
=12(𝕋2)2\𝑫ϕ(x)ϕ(y)r(x,y)r(x,y)𝒦(xy)ρn(t,dx)ρn(t,dy)\displaystyle=\frac{1}{2}\int_{(\mathbb{T}^{2})^{2}\backslash\boldsymbol{D}}\frac{\nabla\phi(x)-\nabla\phi(y)}{r(x,y)}\cdot r(x,y)\mathcal{K}(x-y)\rho_{n}(t,dx)\rho_{n}(t,dy)

is only dependent on ϕ\nabla\phi and ρn(t)\rho_{n}(t) and linear of ϕ.\nabla\phi. In view of (1.5), w(x,y)=r(x,y)𝒦(xy)w(x,y)=r(x,y)\mathcal{K}(x-y) is a bounded function so that for each φC(𝕋2;2)\varphi\in C^{\infty}(\mathbb{T}^{2};\mathbb{R}^{2}),

|(φ(x)φ(y))𝒦(xy)|C𝒦φ|(\varphi(x)-\varphi(y))\cdot\mathcal{K}(x-y)|\leq C_{\mathcal{K}}\|\nabla\varphi\|_{\infty} (2.3)

is bounded. Hence the final expression of (2.2) is always well defined and can be extended to any γ𝒫(𝕋2)\gamma\in\mathcal{P}(\mathbb{T}^{2}). Now we introduce the symmetrization operator 𝐑:𝒫(𝕋2)𝒟(𝕋2)\mathbf{R}:\mathcal{P}(\mathbb{T}^{2})\mapsto\mathcal{D}^{\prime}(\mathbb{T}^{2}) satisfying

φ,𝐑(γ):=12(𝕋2)2\𝑫(φ(x)φ(y))𝒦(xy)γ(dx)γ(dy),φC1(𝕋2;2).\left\langle\varphi,\mathbf{R}(\gamma)\right\rangle:=\frac{1}{2}\int_{(\mathbb{T}^{2})^{2}\backslash\boldsymbol{D}}(\varphi(x)-\varphi(y))\cdot\mathcal{K}(x-y)\gamma(dx)\gamma(dy),\,\forall\varphi\in C^{1}(\mathbb{T}^{2};\mathbb{R}^{2}). (2.4)

As stated above, (2.4) is always well-defined and

|φ,𝐑(γ)|12C𝒦φ.\left|\left\langle\varphi,\mathbf{R}(\gamma)\right\rangle\right|\leq\frac{1}{2}C_{\mathcal{K}}\|\nabla\varphi\|_{\infty}. (2.5)

In fact 𝐑(γ)\mathbf{R}(\gamma) is an extension for γ(𝒦γ).\gamma(\mathcal{K}*\gamma). If γL2(𝕋2),\gamma\in L^{2}(\mathbb{T}^{2}), as stated before Theorem 1.2, 𝒦γL2(𝕋2)\mathcal{K}*\gamma\in L^{2}(\mathbb{T}^{2}). Then for each φC1(𝕋2;2)\varphi\in C^{1}(\mathbb{T}^{2};\mathbb{R}^{2}),

φ,γ(𝒦γ)=𝕋2φ(x)𝒦(xy)γ(dx)γ(dy)=𝕋2φ(y)𝒦(xy)γ(dx)γ(dy).\left\langle\varphi,\gamma(\mathcal{K}*\gamma)\right\rangle=\int_{\mathbb{T}^{2}}\varphi(x)\cdot\mathcal{K}(x-y)\gamma(dx)\gamma(dy)=-\int_{\mathbb{T}^{2}}\varphi(y)\cdot\mathcal{K}(x-y)\gamma(dx)\gamma(dy).

Hence

𝐑(γ)=γ(𝒦γ),γL2(𝕋2).\mathbf{R}(\gamma)=\gamma(\mathcal{K}*\gamma),\quad\forall\gamma\in L^{2}(\mathbb{T}^{2}). (2.6)

With this notation, after further calculations, we can write the generator for ρn\rho_{n} as

Anf(γ):=i=1kφi,𝐑(γ)iψ+νi=1kΔφi,γiψ+νni,j=1kφiφj,γijψ,A_{n}f(\gamma):=\sum_{i=1}^{k}\left\langle\nabla\varphi_{i},\mathbf{R}(\gamma)\right\rangle\partial_{i}\psi+\nu\sum_{i=1}^{k}\left\langle\Delta\varphi_{i},\gamma\right\rangle\partial_{i}\psi+\frac{\nu}{n}\sum_{i,j=1}^{k}\left\langle\nabla\varphi_{i}\cdot\nabla\varphi_{j},\gamma\right\rangle\partial_{ij}\psi, (2.7)

for each fD0:={f:𝒫(𝕋2),f(γ)=ψ(φ1,γ,,φk,γ),ψC2(k),φiC(𝕋2)}f\in D_{0}:=\{f:\mathcal{P}(\mathbb{T}^{2})\mapsto\mathbb{R},f(\gamma)=\psi(\left\langle\varphi_{1},\gamma\right\rangle,\cdots,\left\langle\varphi_{k},\gamma\right\rangle),\psi\in C^{2}(\mathbb{R}^{k}),\varphi_{i}\in C^{\infty}(\mathbb{T}^{2})\} such that

f(ρn(t))0tAnf(ρn(s))𝑑s=martingale.f(\rho_{n}(t))-\int_{0}^{t}A_{n}f(\rho_{n}(s))ds=\text{martingale.}

Then we define

Hnf(γ):=1nenfAnenf(γ)=νΔγdiv𝐑(γ),δfδγ+ν𝕋2|δfδγ|2𝑑γ\displaystyle H_{n}f(\gamma):=\frac{1}{n}e^{-nf}A_{n}e^{nf}(\gamma)=\left\langle\nu\Delta\gamma-{\rm div}\,\mathbf{R}(\gamma),\frac{\delta f}{\delta\gamma}\right\rangle+\nu\int_{\mathbb{T}^{2}}\left|\nabla\frac{\delta f}{\delta\gamma}\right|^{2}d\gamma (2.8)
+νni,j=1kφiφj,γijψ,\displaystyle+\frac{\nu}{n}\sum_{i,j=1}^{k}\left\langle\nabla\varphi_{i}\cdot\nabla\varphi_{j},\gamma\right\rangle\partial_{ij}\psi,

in which Δγ\Delta\gamma as well as div𝐑(γ){\rm div}\,\mathbf{R}(\gamma) is defined in distribution sense and the variational derivative

δfδγ(x):=i=1kiψ(φ1,γ,,φk,γ)φi(x)\frac{\delta f}{\delta\gamma}(x):=\sum_{i=1}^{k}\partial_{i}\psi(\left\langle\varphi_{1},\gamma\right\rangle,\cdots,\left\langle\varphi_{k},\gamma\right\rangle)\varphi_{i}(x)

is a smooth function on 𝕋2\mathbb{T}^{2}.

At least formally by Ito’s formula, for (ft)t0(f_{t})_{t\geq 0} such that ftD0f_{t}\in D_{0}, the exponential form

exp{n[ft(ρn(t))ft(ρn(0))0tsfs(ρn(s))ds0tHnfs(ρn(s))𝑑s]}\exp\left\{n\left[f_{t}(\rho_{n}(t))-f_{t}(\rho_{n}(0))-\int_{0}^{t}\partial_{s}f_{s}(\rho_{n}(s))ds-\int_{0}^{t}H_{n}f_{s}(\rho_{n}(s))ds\right]\right\}

is always considered to be a martingale. In fact, the proof of [33, 14] only use the case when ϕC([0,T]×𝕋2)\phi\in C^{\infty}([0,T]\times\mathbb{T}^{2}) and ft(ρ)=ϕ(t),ρ.f_{t}(\rho)=\left\langle\phi(t),\rho\right\rangle. To adapt their proof to the more general initial data, we shall use a bigger domain D0D_{0}, which is sufficient to determine the rate function (see Section 3.2 of [20]).

2.2 Proof of upper bound

First, as a standard step to prove large deviation principle, we prove the exponential tightness of ρn\rho_{n}, with which we only need to prove the upper bound for compact sets.

Proposition 2.1 (Exponentially tightness).

Under the condition of Theorem 1.3, for each a,T>0,a,T>0, there exists a compact set K^a,TC([0,T];𝒫(𝕋2))\hat{K}_{a,T}\subset C([0,T];\mathcal{P}(\mathbb{T}^{2})) such that

lim supn1nlogP(ρnK^a,T)a.\limsup_{n\to\infty}\frac{1}{n}\log P(\rho_{n}\notin\hat{K}_{a,T})\leq-a. (2.9)
Proof.

Note D0D_{0} is closed under addition and separates points. In addition, recalling (2.8) and (2.5), for f(γ)=ψ(φ1,γ,,φk,γ),ψC2(k),φiC(𝕋2)f(\gamma)=\psi(\left\langle\varphi_{1},\gamma\right\rangle,\cdots,\left\langle\varphi_{k},\gamma\right\rangle),\psi\in C^{2}(\mathbb{R}^{k}),\varphi_{i}\in C^{\infty}(\mathbb{T}^{2})

|Hnf(γ)|νΔδfδγ+νδfδγ2+C2δfδγ\displaystyle|H_{n}f(\gamma)|\leq\nu\left\|\Delta\frac{\delta f}{\delta\gamma}\right\|_{\infty}+\nu\left\|\nabla\frac{\delta f}{\delta\gamma}\right\|^{2}_{\infty}+C\left\|\nabla^{2}\frac{\delta f}{\delta\gamma}\right\|_{\infty} (2.10)
+νni,j=1kφiφjsup|xi|φi|ijψ(x1,,xk)|\displaystyle+\frac{\nu}{n}\sum_{i,j=1}^{k}\|\nabla\varphi_{i}\|_{\infty}\|\nabla\varphi_{j}\|_{\infty}\sup_{|x_{i}|\leq\|\varphi_{i}\|_{\infty}}|\partial_{ij}\psi(x_{1},\cdots,x_{k})|

is bounded. Since 𝒫(𝕋2)\mathcal{P}(\mathbb{T}^{2}) is a compact space, we conclude the proof by Corollary 4.17 of [20]. ∎

Formally, there are three steps to prove the large deviation upper bound for compact sets. First, by Ito’s formula, for each fD0f\in D_{0}, ϕC([0,T]×𝕋2)\phi\in C^{\infty}([0,T]\times\mathbb{T}^{2}) and t0(0,T)t_{0}\in(0,T),

Mn(t):=exp{n[f(ρn(t0t))0t0tHnf(ρn(s))dst0tttϕ(s),ρn(s)ds\displaystyle M_{n}(t):=\exp\bigg{\{}n\bigg{[}f(\rho_{n}(t_{0}\wedge t))-\int_{0}^{t_{0}\wedge t}H_{n}f(\rho_{n}(s))ds-\int_{t_{0}\wedge t}^{t}\left\langle\partial_{t}\phi(s),\rho_{n}(s)\right\rangle ds
+ϕ(t),ρ(t)ϕ(t0t),ρ(t0t)t0ttHn(ϕ(s),)(ρ(s))ds]}\displaystyle+\left\langle\phi(t),\rho(t)\right\rangle-\left\langle\phi(t_{0}\wedge t),\rho(t_{0}\wedge t)\right\rangle-\int_{t_{0}\wedge t}^{t}H_{n}(\left\langle\phi(s),\cdot\right\rangle)(\rho(s))ds\bigg{]}\bigg{\}}
=exp{n[f(ρn(t0t))0t0tHnf(ρn(s))dst0tttϕ(s),ρn(s)ds\displaystyle=\exp\bigg{\{}n\bigg{[}f(\rho_{n}(t_{0}\wedge t))-\int_{0}^{t_{0}\wedge t}H_{n}f(\rho_{n}(s))ds-\int_{t_{0}\wedge t}^{t}\left\langle\partial_{t}\phi(s),\rho_{n}(s)\right\rangle ds
+ϕ(t),ρ(t)ϕ(t0t),ρ(t0t)νt0ttΔϕ(s),ρn(s)𝑑s\displaystyle+\left\langle\phi(t),\rho(t)\right\rangle-\left\langle\phi(t_{0}\wedge t),\rho(t_{0}\wedge t)\right\rangle-\nu\int_{t_{0}\wedge t}^{t}\left\langle\Delta\phi(s),\rho_{n}(s)\right\rangle ds
t0ttϕ(s),𝐑(ρn(s))dsνt0tt𝕋2|ϕ(s)|2dρn(s)ds]}\displaystyle-\int_{t_{0}\wedge t}^{t}\left\langle\nabla\phi(s),\mathbf{R}(\rho_{n}(s))\right\rangle ds-\nu\int_{t_{0}\wedge t}^{t}\int_{\mathbb{T}^{2}}|\nabla\phi(s)|^{2}d\rho_{n}(s)ds\bigg{]}\bigg{\}}

is a non-negative martingale, hence for each measurable set BC([0,T];𝒫(𝕋2))B\subset C([0,T];\mathcal{P}(\mathbb{T}^{2})),

1nlogP(ρnB)=1nlog𝔼[Mn(T)(Mn(T))1χB(ρn)]\displaystyle\frac{1}{{n}}\log P(\rho_{n}\in B)=\frac{1}{n}\log\mathbb{E}\left[M_{n}(T)(M_{n}(T))^{-1}\chi_{B}(\rho_{n})\right] (2.11)
1nlog𝔼Mn(0)infρB1nlogMn(T)\displaystyle\leq\frac{1}{n}\log\mathbb{E}M_{n}(0)-\inf_{\rho\in B}\frac{1}{n}\log M_{n}(T)
=1nlog𝔼enf(ρn(0))infρB(𝒜n,T(ρ,f,t0,ϕ)+f(ρ(0))),\displaystyle=\frac{1}{n}\log\mathbb{E}e^{nf(\rho_{n}(0))}-\inf_{\rho\in B}\left(\mathcal{A}_{n,T}(\rho,f,t_{0},\phi)+f(\rho(0))\right),

where χ\chi is the indicator function and

𝒜n,T(ρ,f,t0,ϕ):=f(ρ(t0))f(ρ(0))0t0Hnf(ρ(t))𝑑tt0Ttϕ(t),ρ(t)𝑑t\displaystyle\mathcal{A}_{n,T}(\rho,f,t_{0},\phi):=f(\rho(t_{0}))-f(\rho(0))-\int_{0}^{t_{0}}H_{n}f(\rho(t))dt-\int_{t_{0}}^{T}\left\langle\partial_{t}\phi(t),\rho(t)\right\rangle dt (2.12)
+ϕ(T),ρ(T)ϕ(t0),ρ(t0)νt0TΔϕ(t),ρ(t)𝑑t\displaystyle+\left\langle\phi(T),\rho(T)\right\rangle-\left\langle\phi(t_{0}),\rho(t_{0})\right\rangle-\nu\int_{t_{0}}^{T}\left\langle\Delta\phi(t),\rho(t)\right\rangle dt
t0Tϕ(t),𝐑(ρ(t))𝑑tνt0T𝕋2|ϕ(t)|2𝑑ρ(t)𝑑t.\displaystyle-\int_{t_{0}}^{T}\left\langle\nabla\phi(t),\mathbf{R}(\rho(t))\right\rangle dt-\nu\int_{t_{0}}^{T}\int_{\mathbb{T}^{2}}|\nabla\phi(t)|^{2}d\rho(t)dt.

Second, obtain the limit of the right-hand-side of (2.11) when nn goes to infinity (write it as Ft0,T,f,ϕF_{t_{0},T,f,\phi}). Taking the supremum over {(f,t0,ϕ):fD0,0<t0<T,ϕC([0,T]×𝕋2)}\{(f,t_{0},\phi):f\in D_{0},0<t_{0}<T,\phi\in C^{\infty}([0,T]\times\mathbb{T}^{2})\}, and then we have

lim supn1nlogP(ρnB)supf,t0,ϕinfρBFt0,T,f,ϕ(ρ).\displaystyle\limsup_{n\to\infty}\frac{1}{n}\log P(\rho_{n}\in B)\leq-\sup_{f,t_{0},\phi}\inf_{\rho\in B}F_{t_{0},T,f,\phi}(\rho).

Finally, use Lemma A2.3.3A2.3.3 in [32] to exchange the supremum and infimum to obtain the upper bound for any compact set KK, i.e.

lim supn1nlogP(ρnK)infρKsupf,t0,ϕFt0,T,f,ϕ(ρ).\displaystyle\limsup_{n\to\infty}\frac{1}{n}\log P(\rho_{n}\in K)\leq-\inf_{\rho\in K}\sup_{f,t_{0},\phi}F_{t_{0},T,f,\phi}(\rho).

However, the exchange of the supremum and infimum requires that FF is lower semi-continuous under the topology of C([0,T];𝒫(𝕋2))C([0,T];\mathcal{P}(\mathbb{T}^{2})). Hence we do need the singular term 𝐑\mathbf{R} is continuous under weak topology. However, that is not true. Fortunately, 𝐑\mathbf{R} is continuous under weak topology with finite energy, stated in the following lemma.

Lemma 2.2.

If γn,γ𝒫(𝕋2),limnd(γn,γ)=0\gamma_{n},\gamma\in\mathcal{P}(\mathbb{T}^{2}),\lim_{n\to\infty}d(\gamma_{n},\gamma)=0, and supne(ζnγn)<\sup_{n}e(\zeta_{n}*\gamma_{n})<\infty (or supne(γn)<\sup_{n}e(\gamma_{n})<\infty), then for each φn,φC1(𝕋2,2)\varphi_{n},\varphi\in C^{1}(\mathbb{T}^{2},\mathbb{R}^{2}) such that supnφn<\sup_{n}\|\nabla\varphi_{n}\|_{\infty}<\infty and limnφnφ=0\lim_{n\to\infty}\|\varphi_{n}-\varphi\|_{\infty}=0, we have

limnφn,𝐑(γn)=φ,𝐑(γ),φC1(𝕋2,2).\lim_{n\to\infty}\left\langle\varphi_{n},\mathbf{R}(\gamma_{n})\right\rangle=\left\langle\varphi,\mathbf{R}(\gamma)\right\rangle,\,\forall\varphi\in C^{1}(\mathbb{T}^{2},\mathbb{R}^{2}).

The proof of this lemma will be given in Sec.5, since it will also be used for the proof of large deviation lower bound.

Thus, we need a prior energy estimation to bound the energy of the trajectories.

Lemma 2.3 (Prior energy estimation).

Under the condition of Theorem 1.3,

limRlim supn\displaystyle\lim_{R\to\infty}\limsup_{n\to\infty} 1nlogP(QT(ζnρn)>R)=.\displaystyle\frac{1}{n}\log P(Q_{T}(\zeta_{n}*\rho_{n})>R)=-\infty. (2.13)
Proposition 2.4 (Upper bound for compact sets).

Under the condition of Theorem 1.3, for any compact set KK in C([0,T];𝒫(𝕋2))C([0,T];\mathcal{P}(\mathbb{T}^{2})),

lim supn1nlogP(ρnK)infρK(0(ρ(0))+𝔸¯T(ρ)+χQT(ρ)=),\displaystyle\limsup_{n\to\infty}\frac{1}{n}\log P(\rho_{n}\in K)\leq-\inf_{\rho\in K}\left(\mathcal{I}_{0}(\rho(0))+\overline{\mathbb{A}}_{T}(\rho)+\infty\cdot\chi_{Q_{T}(\rho)=\infty}\right),

where

𝔸¯T(ρ):=supϕC([0,T]×𝕋2)(ϕ(T),ρ(T)ϕ(0),ρ(0)0Ttϕ(s),ρ(s)ds\displaystyle\overline{\mathbb{A}}_{T}(\rho):=\sup_{\phi\in C^{\infty}([0,T]\times\mathbb{T}^{2})}\bigg{(}\left\langle\phi(T),\rho(T)\right\rangle-\left\langle\phi(0),\rho(0)\right\rangle-\int_{0}^{T}\left\langle\partial_{t}\phi(s),\rho(s)\right\rangle ds (2.14)
ν0TΔϕ(s),ρ(s)ds0Tϕ(s),𝐑(ρ(s))dsν0T𝕋2|ϕ(s)|2dρ(s)ds).\displaystyle-\nu\int_{0}^{T}\left\langle\Delta\phi(s),\rho(s)\right\rangle ds-\int_{0}^{T}\left\langle\nabla\phi(s),\mathbf{R}(\rho(s))\right\rangle ds-\nu\int_{0}^{T}\int_{\mathbb{T}^{2}}|\nabla\phi(s)|^{2}d\rho(s)ds\bigg{)}.
Proof.

Take K^a,T\hat{K}_{a,T} in Proposition 2.1. For any measurable set BC([0,T];𝒫(𝕋2))B\subset C([0,T];\mathcal{P}(\mathbb{T}^{2})), we define

Bn,R,a={ρ:QT(ζnρ)R}BK^a,T.B_{n,R,a}=\{\rho:Q_{T}(\zeta_{n}*\rho)\leq R\}\cap B\cap\hat{K}_{a,T}.

By Lemma 2.3, as R,R\to\infty,

aR:=lim supn1nlogP(QT(ζnρn)>R).a_{R}^{\prime}:=-\limsup_{n\to\infty}\frac{1}{n}\log P(Q_{T}(\zeta_{n}*\rho_{n})>R)\to\infty. (2.15)

In view of (2.11),

1nlog(ρnBn,R,a)1nlog𝔼enf(ρn(0))infρBn,R,a(𝒜n,T(ρ,f,t0,ϕ)+f(ρ(0))).\frac{1}{n}\log(\rho_{n}\in B_{n,R,a})\leq\frac{1}{n}\log\mathbb{E}e^{nf(\rho_{n}(0))}-\inf_{\rho\in B_{n,R,a}}\left(\mathcal{A}_{n,T}(\rho,f,t_{0},\phi)+f(\rho(0))\right).

On the one hand, for fD0f\in D_{0}, define Hf(γ)=νΔγdiv𝐑(γ),δfδγ+ν𝕋2|δfδγ|2𝑑γHf(\gamma)=\left\langle\nu\Delta\gamma-{\rm div}\,\mathbf{R}(\gamma),\frac{\delta f}{\delta\gamma}\right\rangle+\nu\int_{\mathbb{T}^{2}}\left|\nabla\frac{\delta f}{\delta\gamma}\right|^{2}d\gamma and it’s easy to obtain that |Hnf(γ)Hf(γ)|Cfn|H_{n}f(\gamma)-Hf(\gamma)|\leq\frac{C_{f}}{n} for certain constant CfC_{f}. On the other hand, by Varadhan’s lemma,

Λ0(f):=limn1nlog𝔼ef(ρn(0))\Lambda_{0}(f):=\lim_{n\to\infty}\frac{1}{n}\log\mathbb{E}e^{f(\rho_{n}(0))}

exists. Hence,

lim supn1nlogP(ρnBn,R,a)lim infninfρBn,R,a{𝒜T(ρ,f,t0,ϕ)+f(ρ(0))Λ0(f)},\displaystyle\limsup_{n\to\infty}\frac{1}{{n}}\log P(\rho_{n}\in B_{n,R,a})\leq-\liminf_{n\to\infty}\inf_{\rho\in B_{n,R,a}}\bigg{\{}\mathcal{A}_{T}(\rho,f,t_{0},\phi)+f(\rho(0))-\Lambda_{0}(f)\bigg{\}}, (2.16)

where

𝒜T(ρ,f,t0,ϕ):=f(ρ(t0))f(ρ(0))0t0Hf(ρ(s))𝑑s+ϕ(T),ρ(T)\displaystyle\mathcal{A}_{T}(\rho,f,t_{0},\phi):=f(\rho(t_{0}))-f(\rho(0))-\int_{0}^{t_{0}}Hf(\rho(s))ds+\left\langle\phi(T),\rho(T)\right\rangle
ϕ(t0),ρ(t0)t0Ttϕ(s),ρ(s)𝑑sνt0TΔϕ(s),ρ(s)𝑑s\displaystyle-\left\langle\phi(t_{0}),\rho(t_{0})\right\rangle-\int_{t_{0}}^{T}\left\langle\partial_{t}\phi(s),\rho(s)\right\rangle ds-\nu\int_{t_{0}}^{T}\left\langle\Delta\phi(s),\rho(s)\right\rangle ds
t0Tϕ(s),𝐑(ρ(s))𝑑sνt0T𝕋2|ϕ(s)|2𝑑ρ(s)𝑑s.\displaystyle-\int_{t_{0}}^{T}\left\langle\nabla\phi(s),\mathbf{R}(\rho(s))\right\rangle ds-\nu\int_{t_{0}}^{T}\int_{\mathbb{T}^{2}}|\nabla\phi(s)|^{2}d\rho(s)ds.

Now we want to prove that

lim supn1n\displaystyle\limsup_{n\to\infty}\frac{1}{{n}} logP(ρnBn,R,a)\displaystyle\log P(\rho_{n}\in B_{n,R,a}) (2.17)
infρB{𝒜T(ρ,f,t0,ϕ)+f(ρ(0))Λ0(f)+χQT(ρ)>R}.\displaystyle\leq-\inf_{\rho\in B}\bigg{\{}\mathcal{A}_{T}(\rho,f,t_{0},\phi)+f(\rho(0))-\Lambda_{0}(f)+\infty\cdot\chi_{Q_{T}(\rho)>R}\bigg{\}}.

It is trivial if there exists n0n_{0} such that Bn,R,a=B_{n,R,a}=\emptyset for nn0n\geq n_{0}. Otherwise, pick ρ~nBn,R,a\tilde{\rho}_{n}\in B_{n,R,a} approximating the infimum in (2.16) for each nn satisfying Bn,R,aB_{n,R,a}\neq\emptyset, i.e.

𝒜T(ρ~n,f,t0,ϕ)+f(ρ~n(0))infρBn,R,a{𝒜T(ρ,f,t0,ϕ)+f(ρ(0))}+1n.\mathcal{A}_{T}(\tilde{\rho}_{n},f,t_{0},\phi)+f(\tilde{\rho}_{n}(0))\leq\inf_{\rho\in B_{n,R,a}}\bigg{\{}\mathcal{A}_{T}(\rho,f,t_{0},\phi)+f(\rho(0))\bigg{\}}+\frac{1}{n}.

Since Bn,R,aK^a,TB¯B_{n,R,a}\subset\hat{K}_{a,T}\cap\overline{B} and QTQ_{T} is lower semi-continuous, we can find ρ~B¯K^a,T\tilde{\rho}\in\overline{B}\cap\hat{K}_{a,T}, a limiting point of ρ~n\tilde{\rho}_{n}, with QT(ρ~)RQ_{T}(\tilde{\rho})\leq R. Then by Lemma 2.2 and dominated convergence theorem,

𝒜T(ρ~,f,t0,ϕ)+f(ρ~(0))=lim infn(𝒜T(ρ~n,f,t0,ϕ)+f(ρ~n(0)))\displaystyle\mathcal{A}_{T}(\tilde{\rho},f,t_{0},\phi)+f(\tilde{\rho}(0))=\liminf_{n\to\infty}\left(\mathcal{A}_{T}(\tilde{\rho}_{n},f,t_{0},\phi)+f(\tilde{\rho}_{n}(0))\right)
lim infninfρBn,R,a{𝒜T(ρ,f,t0,ϕ)+f(ρ(0))},\displaystyle\leq\liminf_{n\to\infty}\inf_{\rho\in B_{n,R,a}}\bigg{\{}\mathcal{A}_{T}(\rho,f,t_{0},\phi)+f(\rho(0))\bigg{\}},

which combining with (2.16) implies

lim supn1n\displaystyle\limsup_{n\to\infty}\frac{1}{{n}} logP(ρnBn,R,a)\displaystyle\log P(\rho_{n}\in B_{n,R,a}) (2.18)
infρB¯{𝒜T(ρ,f,t0,ϕ)+f(ρ(0))Λ0(f)+χQT(ρ)>R}.\displaystyle\leq-\inf_{\rho\in\overline{B}}\bigg{\{}\mathcal{A}_{T}(\rho,f,t_{0},\phi)+f(\rho(0))-\Lambda_{0}(f)+\infty\cdot\chi_{Q_{T}(\rho)>R}\bigg{\}}.

Use the fact 𝒜T(ρ,f,t0,ϕ)\mathcal{A}_{T}(\rho,f,t_{0},\phi) is continuous when QT(ρ)RQ_{T}(\rho)\leq R, which can be obtained by Lemma 2.2 and dominated convergence theorem, then we arrive at (2.17)

So far, by (2.9), (2.15) and (2.17), for each a,R>0a,R>0,

lim supn1nlogP(ρnB)max{lim supn1nlogP(ρnBn,R,a),\displaystyle\limsup_{n\to\infty}\frac{1}{{n}}\log P(\rho_{n}\in B)\leq\max\bigg{\{}\limsup_{n\to\infty}\frac{1}{{n}}\log P(\rho_{n}\in B_{n,R,a}),
lim supn1nlogP(ρnK^a,Tc),lim supn1nlogP(QT(ζnρn)>R)}\displaystyle\limsup_{n\to\infty}\frac{1}{{n}}\log P(\rho_{n}\in\hat{K}_{a,T}^{c}),\limsup_{n\to\infty}\frac{1}{{n}}\log P(Q_{T}(\zeta_{n}*\rho_{n})>R)\bigg{\}}
infρBmin{𝒜T(ρ,f,t0,ϕ)+f(ρ(0))Λ0(f)+χQT(ρ)>R,a,aR}.\displaystyle\leq-\inf_{\rho\in B}\min\bigg{\{}\mathcal{A}_{T}(\rho,f,t_{0},\phi)+f(\rho(0))-\Lambda_{0}(f)+\infty\cdot\chi_{Q_{T}(\rho)>R},a,a^{\prime}_{R}\bigg{\}}.

Taking aa\to\infty, we obtain

lim supn1nlogP(ρnB)\displaystyle\limsup_{n\to\infty}\frac{1}{{n}}\log P(\rho_{n}\in B) (2.19)
infρBmin{𝒜T(ρ,f,t0,ϕ)+f(ρ(0))Λ0(f)+χQT(ρ)>R,aR},\displaystyle\leq-\inf_{\rho\in B}\min\bigg{\{}\mathcal{A}_{T}(\rho,f,t_{0},\phi)+f(\rho(0))-\Lambda_{0}(f)+\infty\cdot\chi_{Q_{T}(\rho)>R},a^{\prime}_{R}\bigg{\}},

where the expression inside infimum is lower semi-continuous with respect to ρ\rho. Taking the infimum with respect to f,t0f,t_{0} and ϕ\phi in the RHS of (2.19) over {(f,t0,ϕ):fD0,0<t0<T,ϕC([0,T]×𝕋2)}\{(f,t_{0},\phi):f\in D_{0},0<t_{0}<T,\phi\in C^{\infty}([0,T]\times\mathbb{T}^{2})\}, by Lemma A2.3.3 in [32], we can exchange the order of infimum and supremum for compact sets. As consequence, for any compact set KK, by (2.15), we have

lim supn1nlogP(ρnK)limRinfρKsupf,t0,ϕmin{𝒜T(ρ,f,t0,ϕ)+f(ρ(0))Λ0(f)\displaystyle\limsup_{n\to\infty}\frac{1}{n}\log P(\rho_{n}\in K)\leq-\lim_{R\to\infty}\inf_{\rho\in K}\sup_{f,t_{0},\phi}\min\big{\{}\mathcal{A}_{T}(\rho,f,t_{0},\phi)+f(\rho(0))-\Lambda_{0}(f) (2.20)
+χQT(ρ)>R,aR}\displaystyle+\infty\cdot\chi_{Q_{T}(\rho)>R},a^{\prime}_{R}\big{\}}
limRinfρKsupf,t0,ϕmin{𝒜T(ρ,f,t0,ϕ)+f(ρ(0))Λ0(f)+χQT(ρ)=,aR}\displaystyle\leq-\lim_{R\to\infty}\inf_{\rho\in K}\sup_{f,t_{0},\phi}\min\big{\{}\mathcal{A}_{T}(\rho,f,t_{0},\phi)+f(\rho(0))-\Lambda_{0}(f)+\infty\cdot\chi_{Q_{T}(\rho)=\infty},a^{\prime}_{R}\big{\}}
=infρKsupf,t0,ϕ{𝒜T(ρ,f,t0,ϕ)+f(ρ(0))Λ0(f)+χQT(ρ)=}.\displaystyle=-\inf_{\rho\in K}\sup_{f,t_{0},\phi}\big{\{}\mathcal{A}_{T}(\rho,f,t_{0},\phi)+f(\rho(0))-\Lambda_{0}(f)+\infty\cdot\chi_{Q_{T}(\rho)=\infty}\big{\}}.

Now we only remain to show for each ρC([0,T];𝒫(𝕋2)),\rho\in C([0,T];\mathcal{P}(\mathbb{T}^{2})),

supf,t0,ϕ(𝒜T(ρ,f,t0,ϕ)+f(ρ(0))Λ0(f))0(ρ(0))+𝔸¯T(ρ).\displaystyle\sup_{f,t_{0},\phi}\left(\mathcal{A}_{T}(\rho,f,t_{0},\phi)+f(\rho(0))-\Lambda_{0}(f)\right)\geq\mathcal{I}_{0}(\rho(0))+\overline{\mathbb{A}}_{T}(\rho). (2.21)

By Proposition 3.17 of [20], we can find fnD0f_{n}\in D_{0} such that

limn(fn(ρ(0))Λ0(fn))=0(ρ(0)).\lim_{n\to\infty}\left(f_{n}(\rho(0))-\Lambda_{0}(f_{n})\right)=\mathcal{I}_{0}(\rho(0)). (2.22)

By (2.5), we know |Hf(ρ(t))||Hf(\rho(t))| is bounded. Thus, for ρC([0,T];𝒫(𝕋2)),\rho\in C([0,T];\mathcal{P}(\mathbb{T}^{2})),

supϕC([0,T]×𝕋2)limnlimt0+𝒜T(ρ,fn,t,ϕ)=𝔸¯T(ρ).\sup_{\phi\in C^{\infty}([0,T]\times\mathbb{T}^{2})}\lim_{n\to\infty}\lim_{t\to 0+}\mathcal{A}_{T}(\rho,f_{n},t,\phi)=\overline{\mathbb{A}}_{T}(\rho). (2.23)

Combining (2.22) with (2.23), we have

LHS of (2.21)supϕC([0,T]×𝕋2)limnlimt0+(𝒜T(ρ,fn,t,ϕ)+fn(ρ(0))Λ0(fn))\displaystyle\text{LHS of (\ref{equ_atatat})}\geq\sup_{\phi\in C^{\infty}([0,T]\times\mathbb{T}^{2})}\lim_{n\to\infty}\lim_{t\to 0+}\left(\mathcal{A}_{T}(\rho,f_{n},t,\phi)+f_{n}(\rho(0))-\Lambda_{0}(f_{n})\right)
𝔸¯T(ρ)+0(ρ(0)),\displaystyle\geq\overline{\mathbb{A}}_{T}(\rho)+\mathcal{I}_{0}(\rho(0)),

which together with (2.20) completes the proof. ∎

With the prior estimation QT(ρ)<Q_{T}(\rho)<\infty, we can obtain the upper bound is, as shown in the lemma below, actually equal to T(ρ)\mathcal{I}_{T}(\rho) in (1.8).

Lemma 2.5 (Variational representation of the rate function).

If ρC([0,T];𝒫(𝕋2))\rho\in C([0,T];\mathcal{P}(\mathbb{T}^{2})) such that QT(ρ)<Q_{T}(\rho)<\infty and 𝔸¯T(ρ)<\overline{\mathbb{A}}_{T}(\rho)<\infty, then ρAC((0,T);𝒫(𝕋2))\rho\in AC((0,T);\mathcal{P}(\mathbb{T}^{2})). Conversely, if ρAC((0,T);𝒫(𝕋2))\rho\in AC((0,T);\mathcal{P}(\mathbb{T}^{2})) with QT(ρ)<Q_{T}(\rho)<\infty, then 𝔸T(ρ)=𝔸¯T(ρ).\mathbb{A}_{T}(\rho)=\overline{\mathbb{A}}_{T}(\rho).=

Finally, we obtain the large deviation upper bound.

Proposition 2.6 (Upper bound).

Under the condition of Theorem 1.3, for each closed set AA in C([0,T];𝒫(𝕋2)),C([0,T];\mathcal{P}(\mathbb{T}^{2})),

lim supn1nlogP(ρnA)infρAT(ρ).\limsup_{n\to\infty}\frac{1}{n}\log P(\rho_{n}\in A)\leq-\inf_{\rho\in A}\mathcal{I}_{T}(\rho). (2.24)
Proof.

In view of Proposition 2.4, we have proved the upper bound for compact sets. By Lemma 2.5, we verified this upper bound is equal to infρBT(ρ)-\inf_{\rho\in B}\mathcal{I}_{T}(\rho). Finally we conclude the proof by Proposition 2.1 and Lemma 1.2.18 in [13], ∎

2.3 Proof of lower bound

We firstly study the law of large numbers of a regularly perturbed model, which can be obtained by measure transformation from the original model. For any given vL([0,T]×𝕋2;2)v\in L^{\infty}([0,T]\times\mathbb{T}^{2};\mathbb{R}^{2}), take

Zv(t):=exp[i=1n0tv(s,Xi(s))2ν𝑑Bi(s)n4ν0t𝕋2|v(s)|2𝑑ρn(s)𝑑s].Z^{v}(t):=\exp\left[\sum_{i=1}^{n}\int_{0}^{t}\frac{v(s,X_{i}(s))}{\sqrt{2\nu}}dB_{i}(s)-\frac{n}{4\nu}\int_{0}^{t}\int_{\mathbb{T}^{2}}|v(s)|^{2}d\rho_{n}(s)ds\right].

By Proposition 5.12 of [31], it’s a martingale. Thus we can apply Girsanov formula. Let T\mathcal{F}_{T} be the augmented filtration given by initial data and the Brownian motion (see (2.3) in Section 5.2 of [31]). For ATA\in\mathcal{F}_{T}, let Pv(A)(=PTv(A)):=𝔼[χAZv(T)].P^{v}(A)(=P^{v}_{T}(A)):=\mathbb{E}[\chi_{A}Z^{v}(T)]. Let Wi(t)=Bi(t)v(t,Xi(t))2νW_{i}(t)=B_{i}(t)-\frac{v(t,X_{i}(t))}{\sqrt{2\nu}} and then {Wt}0tT\{W_{t}\}_{0\leq t\leq T} is a 2n2n-dimensional Brownian motion under PvP^{v}. In addition, we have

dXi(t)=1nij𝒦(XiXj)dt+v(t,Xi)dt+2νdWi(t).dX_{i}(t)=\frac{1}{n}\sum_{i\neq j}\mathcal{K}(X_{i}-X_{j})dt+v(t,X_{i})dt+\sqrt{2\nu}dW_{i}(t). (2.25)

Formally, the empirical distribution ρn\rho_{n} of (2.25) would converge to the solution of

tρνΔρ+div(𝐑(ρ))+div(ρv)=0.\partial_{t}\rho-\nu\Delta\rho+{\rm div}\,(\mathbf{R}(\rho))+{\rm div}\,(\rho v)=0. (2.26)

Hence a definition of weak solution is required for characterizing the mean-field limit.

Definition 2.1.

For vL([0,T]×𝕋2;2)v\in L^{\infty}([0,T]\times\mathbb{T}^{2};\mathbb{R}^{2}), we say ρC([0,T];𝒫(𝕋2))\rho\in C([0,T];\mathcal{P}(\mathbb{T}^{2})) is a weak solution of (2.26) if QT(ρ)<Q_{T}(\rho)<\infty and for each ϕC([0,T]×𝕋2)\phi\in C^{\infty}([0,T]\times\mathbb{T}^{2}) and 0s<tT0\leq s<t\leq T,

ϕ(t),ρ(t)ϕ(s),ρ(s)=sttϕ(r),ρ(r)𝑑r+νstΔϕ(r),ρ(r)𝑑r\displaystyle\left\langle\phi(t),\rho(t)\right\rangle-\left\langle\phi(s),\rho(s)\right\rangle=\int_{s}^{t}\left\langle\partial_{t}\phi(r),\rho(r)\right\rangle dr+\nu\int_{s}^{t}\left\langle\Delta\phi(r),\rho(r)\right\rangle dr
+stϕ(r),𝐑(ρ(r))𝑑r+st𝕋2ϕ(r)v(r)𝑑ρ(r)𝑑r.\displaystyle+\int_{s}^{t}\left\langle\nabla\phi(r),\mathbf{R}(\rho(r))\right\rangle dr+\int_{s}^{t}\int_{\mathbb{T}^{2}}\nabla\phi(r)\cdot v(r)d\rho(r)dr.

Then we will prove the law of large numbers under PvP^{v}.

Lemma 2.7.

Let γ𝒫(𝕋2)\gamma\in\mathcal{P}(\mathbb{T}^{2}) with e(γ)<e(\gamma)<\infty and vL([0,T]×𝕋2;2)v\in L^{\infty}([0,T]\times\mathbb{T}^{2};\mathbb{R}^{2}). Then there exists an unique weak solution ργ,v\rho^{\gamma,v} of (2.26) such that ρ(0)=γ\rho(0)=\gamma and

QT(ργ,v)e(γ)+CvT,Q_{T}(\rho^{\gamma,v})\leq e(\gamma)+C_{v}T,

where CvC_{v} is a vv-dependent constant. In addition, for each ε>0,R>0\varepsilon>0,R>0, if a sequence (γn)(\gamma_{n}) satisfies γn𝒳n\gamma_{n}\in\mathcal{X}_{n}, e(ζnγn)Re(\zeta_{n}*\gamma_{n})\leq R and limnd(γn,γ)=0\lim_{n\to\infty}d(\gamma_{n},\gamma)=0, then

limnPγnv(sup0tTd(ρn(t),ργ,v(t))>ε)=0.\lim_{n\to\infty}P^{v}_{\gamma_{n}}\left(\sup_{0\leq t\leq T}d\left(\rho_{n}(t),\rho^{\gamma,v}(t)\right)>\varepsilon\right)=0.
Remark.

It might happen that 𝒳n{η:e(ζnη)R}=\mathcal{X}_{n}\cap\{\eta:e(\zeta_{n}*\eta)\leq R\}=\emptyset for small nn, but it would not happen when n,Rn,R big enough (see Lemma 5.3). Hence, we allow γn\gamma_{n} to be defined only for large nn.

Generally speaking, this result would imply that if the initial data γn\gamma_{n} converging to γ\gamma in weak topology, then the limit of relative entropy of PTvP^{v}_{T} with respect to PTP_{T} under the scaling 1/n1/n would provide a large deviation lower bound, i.e.

lim infn1nlogP(ρB)limn1nH(Pγnv|Pγn)=14ν0T𝕋2|v(t)|2ργ,v(t)𝑑t\liminf_{n\to\infty}\frac{1}{n}\log P(\rho\in B)\geq-\lim_{n\to\infty}\frac{1}{n}H\left(P^{v}_{\gamma_{n}}|P_{\gamma_{n}}\right)=-\frac{1}{4\nu}\int_{0}^{T}\int_{\mathbb{T}^{2}}|v(t)|^{2}\rho^{\gamma,v}(t)dt

for any open set BB containing ργ,v\rho^{\gamma,v}, in which

H(Pγnv|Pγn):=𝔼γnv(logdPγnvdPγn)=𝔼γnv(logZv(T))=n4ν𝔼γnv[0T𝕋2|v(t)|2𝑑ρn(t)𝑑t].H\left(P^{v}_{\gamma_{n}}|P_{\gamma_{n}}\right):=\mathbb{E}^{v}_{\gamma_{n}}\left(\log\frac{dP^{v}_{\gamma_{n}}}{dP_{\gamma_{n}}}\right)=\mathbb{E}^{v}_{\gamma_{n}}\left(\log Z^{v}(T)\right)=\frac{n}{4\nu}\mathbb{E}^{v}_{\gamma_{n}}\left[\int_{0}^{T}\int_{\mathbb{T}^{2}}|v(t)|^{2}d\rho_{n}(t)dt\right].

Moreover, at least formally, by taking the infimum for vv we will obtain exactly the rate function, i.e.

𝔸T(ρ)=infv:ρ=ργ,v14ν0T𝕋2|v(t)|2ρ(t)𝑑t.\mathbb{A}_{T}(\rho)=\inf_{v^{\prime}:\rho=\rho^{\gamma,v^{\prime}}}\frac{1}{4\nu}\int_{0}^{T}\int_{\mathbb{T}^{2}}|v^{\prime}(t)|^{2}\rho(t)dt. (2.27)

Define

𝒥={v=p:p:[0,T]×𝕋2,p(t)C2(𝕋2),p+2p<}.\mathcal{J}=\{v=\nabla p:p:[0,T]\times\mathbb{T}^{2}\mapsto\mathbb{R},p(t)\in C^{2}(\mathbb{T}^{2}),\|\nabla p\|_{\infty}+\|\nabla^{2}p\|_{\infty}<\infty\}.

In our paper, for ρ\rho regular enough, i.e. ρ=ργ,v\rho=\rho^{\gamma,v} for certain v𝒥v\in\mathcal{J}, the infimum in (2.27) is always taken in the case that v=vv^{\prime}=v.

Since our initial data is more general, we need a stronger result stated below involving the stability for initial values.

Lemma 2.8.

Let γ𝒫(𝕋2)\gamma\in\mathcal{P}(\mathbb{T}^{2}) with e(γ)<e(\gamma)<\infty and vL([0,T]×𝕋2;2)v\in L^{\infty}([0,T]\times\mathbb{T}^{2};\mathbb{R}^{2}). For any ε>0,R>0,\varepsilon>0,R>0, there exists δ(0,ε)\delta\in(0,\varepsilon) such that for each sequence (γn)(\gamma_{n}) satisfying γnBδ(γ)𝒳n\gamma_{n}\in B_{\delta}(\gamma)\cap\mathcal{X}_{n} and e(ζnγn)Re(\zeta_{n}*\gamma_{n})\leq R,

limnPγnv(sup0tTd(ρn(t),ργ,v(t))>ε)=0,\lim_{n\to\infty}P^{v}_{\gamma_{n}}\left(\sup_{0\leq t\leq T}d(\rho_{n}(t),\rho^{\gamma,v}(t))>\varepsilon\right)=0, (2.28)

where Bδ(γ)={η𝒫(𝕋2):d(η,γ)<δ}.B_{\delta}(\gamma)=\{\eta\in\mathcal{P}(\mathbb{T}^{2}):d(\eta,\gamma)<\delta\}. In addition, if v𝒥v\in\mathcal{J}, then there exists a constant CvC_{v} dependent on vv such that

lim supn|14ν𝔼γnv[0T𝕋2|v(t)|2𝑑ρn(t)𝑑t]𝔸T(ργ,v)|Cvε.\limsup_{n\to\infty}\left|\frac{1}{4\nu}\mathbb{E}^{v}_{\gamma_{n}}\left[\int_{0}^{T}\int_{\mathbb{T}^{2}}|v(t)|^{2}d\rho_{n}(t)dt\right]-\mathbb{A}_{T}(\rho^{\gamma,v})\right|\leq C_{v}\varepsilon. (2.29)

As one may expect, the set of ”nice” trajectories, consisting of all ργ,v\rho^{\gamma,v} for v𝒥v\in\mathcal{J}, plays a key role in proof of lower bound of LDP. We want to prove each ρ\rho with finite rate function can be approximated by a series of ργ,v,v𝒥\rho^{\gamma,v},v\in\mathcal{J} in the sense stated below.

Lemma 2.9 (Density of nice trajectory).

For each γ𝒫(𝕋2)\gamma\in\mathcal{P}(\mathbb{T}^{2}) with e(γ)<e(\gamma)<\infty, define Fregγ={ρAC((0,T);𝒫(𝕋2)):QT(ρ)<,ρ(0)=γ}.F^{\gamma}_{reg}=\{\rho\in AC((0,T);\mathcal{P}(\mathbb{T}^{2})):Q_{T}(\rho)<\infty,\rho(0)=\gamma\}. If ρFregγ\rho\in F_{reg}^{\gamma} with 𝔸T(ρ)<\mathbb{A}_{T}(\rho)<\infty, there exists vn𝒥v_{n}\in\mathcal{J} such that

limnsup0tTd(ργ,vn(t),ρ(t))=0,lim supn𝔸T(ργ,vn)𝔸T(ρ).\lim_{n\to\infty}\sup_{0\leq t\leq T}d(\rho^{\gamma,v_{n}}(t),\rho(t))=0,\quad\limsup_{n\to\infty}\mathbb{A}_{T}(\rho^{\gamma,v_{n}})\leq\mathbb{A}_{T}(\rho).

Now we are ready to prove the large deviation lower bound. By classical result of large deviation principle (see [40], Proposition 1.15), to prove lower bound for open set, it suffices to prove that lower bound holds for any open ball Bε(ρ)B_{\varepsilon}(\rho) in C([0,T];𝒫(𝕋2))C([0,T];\mathcal{P}(\mathbb{T}^{2})), which is stated below.

Proposition 2.10 (Lower bound).

Under the condition of Theorem 1.3, for any ρC([0,T];𝒫(𝕋2))\rho\in C([0,T];\mathcal{P}(\mathbb{T}^{2})) and ε>0\varepsilon>0,

lim infn1nlogP(sup0tTd(ρn(t),ρ(t))<ε)T(ρ).\liminf_{n\to\infty}\frac{1}{n}\log P\left(\sup_{0\leq t\leq T}d(\rho_{n}(t),\rho(t))<\varepsilon\right)\geq-\mathcal{I}_{T}(\rho).
Proof.

If T(ρ)=\mathcal{I}_{T}(\rho)=\infty, the result is trivial. Otherwise, ρAC((0,T);𝒫(𝕋2))\rho\in AC((0,T);\mathcal{P}(\mathbb{T}^{2})) and QT(ρ)<Q_{T}(\rho)<\infty. By Condition 1.2, we can pick RR big enough such that

lim supn1nlogP(e(ζnρn(0))>R)<0(γ)1.\limsup_{n\to\infty}\frac{1}{n}\log P(e(\zeta_{n}*\rho_{n}(0))>R)<-\mathcal{I}_{0}(\gamma)-1. (2.30)

By LDP of initial data, for each δ>0\delta>0,

lim infn1nlogP(ρn(0)Bδ(γ))0(γ),\liminf_{n\to\infty}\frac{1}{n}\log P(\rho_{n}(0)\in B_{\delta}(\gamma))\geq-\mathcal{I}_{0}(\gamma),

which in combination with (2.30) leads to

lim infn1nlogP(ρn(0)Bδ(γ),e(ζnρn(0))R)0(γ)>.\liminf_{n\to\infty}\frac{1}{n}\log P(\rho_{n}(0)\in B_{\delta}(\gamma),e(\zeta_{n}*\rho_{n}(0))\leq R)\geq-\mathcal{I}_{0}(\gamma)>-\infty. (2.31)

Let 𝒪=Bε(ρ),\mathcal{O}=B_{\varepsilon}(\rho), and assume ρ(0)=γ.\rho(0)=\gamma. By Lemma 2.9, we can select vm𝒥v_{m}\in\mathcal{J} such that ργ,vmBε(ρ)\rho^{\gamma,v_{m}}\in B_{\varepsilon}(\rho) and

limmsup0tTd(ργ,vm(t),ρ(t))=0,lim supm𝔸T(ργ,vm)𝔸T(ρ).\lim_{m\to\infty}\sup_{0\leq t\leq T}d(\rho^{\gamma,v_{m}}(t),\rho(t))=0,\quad\limsup_{m\to\infty}\mathbb{A}_{T}(\rho^{\gamma,v_{m}})\leq\mathbb{A}_{T}(\rho). (2.32)

We claim that for each mm,

lim infn1nlogP(ρn𝒪)𝔸T(ργ,vm)0(γ).\liminf_{n\to\infty}\frac{1}{n}\log P\left(\rho_{n}\in\mathcal{O}\right)\geq-\mathbb{A}_{T}(\rho^{\gamma,v_{m}})-\mathcal{I}_{0}(\gamma). (2.33)

Then by (2.32) we have

lim infn1nlogP(ρn𝒪)𝔸T(ρ)0(γ)=T(ρ),\liminf_{n\to\infty}\frac{1}{n}\log P\left(\rho_{n}\in\mathcal{O}\right)\geq-\mathbb{A}_{T}(\rho)-\mathcal{I}_{0}(\gamma)=-\mathcal{I}_{T}(\rho),

and arrive at the lower bound.

So we just need to verify (2.33). For any small enough ε1<ε\varepsilon_{1}<\varepsilon such that 𝒪1=Bε1(ργ,vm)𝒪\mathcal{O}_{1}=B_{\varepsilon_{1}}(\rho^{\gamma,v_{m}})\subset\mathcal{O}, take δ<ε1\delta<\varepsilon_{1} in Lemma 2.8 such that (2.28) and (2.29) holds replacing ε\varepsilon by ε1\varepsilon_{1}. Note that

P(ρn𝒪\displaystyle P(\rho_{n}\in\mathcal{O} )infγnBδ(γ)𝒳n,e(ζnγn)RPγn(ρn𝒪)\displaystyle)\geq\inf_{\gamma_{n}\in B_{\delta}(\gamma)\cap\mathcal{X}_{n},e(\zeta_{n}*\gamma_{n})\leq R}P_{\gamma_{n}}(\rho_{n}\in\mathcal{O}) (2.34)
P(ρn(0)Bδ(γ),e(ζnρn(0))R).\displaystyle\cdot P(\rho_{n}(0)\in B_{\delta}(\gamma),e(\zeta_{n}*\rho_{n}(0))\leq R).

By (2.31), there exists n0n_{0} such that when Bδ(γ)𝒳nB_{\delta}(\gamma)\cap\mathcal{X}_{n}\neq\emptyset for nn0n\geq n_{0}. Hence for nn0n\geq n_{0}, take γ¯nBδ(γ)𝒳n\overline{\gamma}_{n}\in B_{\delta}(\gamma)\cap\mathcal{X}_{n} such that e(ζnγ¯n)Re(\zeta_{n}*\overline{\gamma}_{n})\leq R and

1nlogPγ¯n(ρn𝒪)infγnBδ(γ)𝒳n,e(ζnγn)R1nlogPγn(ρn𝒪)+1n.\frac{1}{n}\log P_{\overline{\gamma}_{n}}(\rho_{n}\in\mathcal{O})\leq\inf_{\gamma_{n}\in B_{\delta}(\gamma)\cap\mathcal{X}_{n},e(\zeta_{n}*\gamma_{n})\leq R}\frac{1}{n}\log P_{\gamma_{n}}(\rho_{n}\in\mathcal{O})+\frac{1}{n}. (2.35)

In view of (2.31), (2.34) and (2.35),

lim infn1nlogP(ρn𝒪)lim infn1nlogPγ¯n(ρn𝒪)\displaystyle\liminf_{n\to\infty}\frac{1}{n}\log P(\rho_{n}\in\mathcal{O})\geq\liminf_{n\to\infty}\frac{1}{n}\log P_{\overline{\gamma}_{n}}(\rho_{n}\in\mathcal{O}) (2.36)
+lim infn1nlogP(ρn(0)Bδ(γ),e(ζnρn(0))R)\displaystyle+\liminf_{n\to\infty}\frac{1}{n}\log P(\rho_{n}(0)\in B_{\delta}(\gamma),e(\zeta_{n}*\rho_{n}(0))\leq R)
lim infn1nlogPγ¯n(ρn𝒪)0(γ).\displaystyle\geq\liminf_{n\to\infty}\frac{1}{n}\log P_{\overline{\gamma}_{n}}(\rho_{n}\in\mathcal{O})-\mathcal{I}_{0}(\gamma).

Since 𝒪1T\mathcal{O}_{1}\in\mathcal{F}_{T},

Pγ¯n(ρn𝒪1)=Pγ¯nvm(𝒪1)𝔼γ¯nvm[(Zvm(T))1χ𝒪1Pγ¯nvm(𝒪1)].P_{\overline{\gamma}_{n}}(\rho_{n}\in\mathcal{O}_{1})=P_{\overline{\gamma}_{n}}^{v_{m}}(\mathcal{O}_{1})\mathbb{E}_{\overline{\gamma}_{n}}^{v_{m}}\bigg{[}\left(Z^{v_{m}}(T)\right)^{-1}\frac{\chi_{\mathcal{O}_{1}}}{P_{\overline{\gamma}_{n}}^{v_{m}}(\mathcal{O}_{1})}\bigg{]}.

By Jensen’s inequality,

1nlogPγ¯n(ρn𝒪1)(Pγ¯nvm(ρn𝒪1))1Eγ¯nvm[1nlogZvm(T);𝒪1]\displaystyle\frac{1}{n}\log P_{\overline{\gamma}_{n}}(\rho_{n}\in\mathcal{O}_{1})\geq-(P_{\overline{\gamma}_{n}}^{v_{m}}(\rho_{n}\in\mathcal{O}_{1}))^{-1}E_{\overline{\gamma}_{n}}^{v_{m}}\bigg{[}\frac{1}{n}\log Z^{v_{m}}(T);\mathcal{O}_{1}\bigg{]}
+1nlogPγ¯nvm(ρn𝒪1).\displaystyle+\frac{1}{n}\log P_{\overline{\gamma}_{n}}^{v_{m}}(\rho_{n}\in\mathcal{O}_{1}).

By Lemma 2.8, the second expression on the RHS of the above inequality converges to 0. The first one is equal to

1Pγ¯nvm(ρn𝒪1){Eγ¯nvm[1nlogZvm(T)]Eγ¯nvm[1nlogZvm(T);𝒪1c]}.-\frac{1}{P_{\overline{\gamma}_{n}}^{v_{m}}(\rho_{n}\in\mathcal{O}_{1})}\bigg{\{}E_{\overline{\gamma}_{n}}^{v_{m}}\bigg{[}\frac{1}{n}\log Z^{v_{m}}(T)\bigg{]}-E_{\overline{\gamma}_{n}}^{v_{m}}\bigg{[}\frac{1}{n}\log Z^{v_{m}}(T);\mathcal{O}_{1}^{c}\bigg{]}\bigg{\}}. (2.37)

Noting

Zvm(T)=exp[i=1n0Tvm(s,Xi(t))2ν𝑑Wi(t)+n4ν0T𝕋2|vm(t)|2𝑑ρn(t)𝑑t],\displaystyle Z^{v_{m}}(T)=\exp\left[\sum_{i=1}^{n}\int_{0}^{T}\frac{v_{m}(s,X_{i}(t))}{\sqrt{2\nu}}dW_{i}(t)+\frac{n}{4\nu}\int_{0}^{T}\int_{\mathbb{T}^{2}}|v_{m}(t)|^{2}d\rho_{n}(t)dt\right],
Varγ¯nvm[1ni=1n0Tvm(t,Xi(t))2ν𝑑Wi(t)]12nνvm2T,\displaystyle\text{Var}^{v_{m}}_{\overline{\gamma}_{n}}\left[\frac{1}{n}\sum_{i=1}^{n}\int_{0}^{T}\frac{v_{m}(t,X_{i}(t))}{\sqrt{2\nu}}dW_{i}(t)\right]\leq\frac{1}{2n\nu}\|v_{m}\|^{2}_{\infty}T,
|14ν0T𝕋2|vm(t)|2𝑑ρn(t)|14νvm2T,\displaystyle\bigg{|}\frac{1}{4\nu}\int_{0}^{T}\int_{\mathbb{T}^{2}}|v_{m}(t)|^{2}d\rho_{n}(t)\bigg{|}\leq\frac{1}{4\nu}\|v_{m}\|^{2}_{\infty}T,

and by (2.28), the second term in (2.37) converges to zero, and the denominator in the first terms converges to one, so we have

limn|(2.37)+𝔼γ¯nvm[14ν0T𝕋2|vm(t)|2𝑑ρn(t)𝑑t]|=0.\lim_{n\to\infty}\bigg{|}(\ref{equ_lower7})+\mathbb{E}_{\overline{\gamma}_{n}}^{v_{m}}\bigg{[}\frac{1}{4\nu}\int_{0}^{T}\int_{\mathbb{T}^{2}}|v_{m}(t)|^{2}d\rho_{n}(t)dt\bigg{]}\bigg{|}=0.

Therefore, so far we have,

lim infn1nlogPγ¯n(ρn𝒪)lim infn1nlogPγ¯n(ρn𝒪1)\displaystyle\liminf_{n\to\infty}\frac{1}{n}\log P_{\overline{\gamma}_{n}}(\rho_{n}\in\mathcal{O})\geq\liminf_{n\to\infty}\frac{1}{n}\log P_{\overline{\gamma}_{n}}(\rho_{n}\in\mathcal{O}_{1}) (2.38)
lim supn𝔼γ¯nvm[14ν0t𝕋2|vm(t)|2𝑑ρn(t)].\displaystyle\geq-\limsup_{n\to\infty}\mathbb{E}_{\overline{\gamma}_{n}}^{v_{m}}\bigg{[}\frac{1}{4\nu}\int_{0}^{t}\int_{\mathbb{T}^{2}}|v_{m}(t)|^{2}d\rho_{n}(t)\bigg{]}.

Then by (2.29),

lim infn1nlogPγ¯n(ρn𝒪)𝔸T(ργ,pm)Cvmε1.\liminf_{n\to\infty}\frac{1}{n}\log P_{\overline{\gamma}_{n}}(\rho_{n}\in\mathcal{O})\geq-\mathbb{A}_{T}(\rho^{\gamma,p_{m}})-C^{\prime}_{v_{m}}\varepsilon_{1}. (2.39)

So (2.33) follows from (2.36) and (2.39) by taking ε10\varepsilon_{1}\to 0. ∎

3 Prior energy estimation and crucial inequalities

This section investigates some estimations related to the energy functional along with 𝐑(ρ)\mathbf{R}(\rho). In Section 3.1, we prove Condition 1.1 implies Condition 1.2. Section 3.2 contains some crucial inequalities, which will be used throughout the paper. The proof of Lemma 2.3 is provided in Section 3.3.

3.1 Energy with mollification

Recall for x,x\in\mathbb{R},

ζ(x)={Ce114|x|2,|x|<12,0,|x|12,\zeta(x)=\left\{\begin{aligned} &Ce^{-\frac{1}{1-4|x|^{2}}},&|x|<\frac{1}{2},\\ &0,&|x|\geq\frac{1}{2},\end{aligned}\right.

and ζn:2,\zeta_{n}:\mathbb{R}^{2}\mapsto\mathbb{R},

ζn(x)=mn2ζ(mn|x|),\zeta_{n}(x)=m_{n}^{2}\zeta(m_{n}|x|), (3.1)

where mn,nmn2m_{n}\uparrow\infty,nm_{n}^{-2}\to\infty. For convenience, we take m15m_{1}\geq 5. In fact, the estimations in this section hold for each ζ\zeta supported on [12,12]\left[-\frac{1}{2},\frac{1}{2}\right] and satisfying the Condition 3.1 below.

Condition 3.1.

(1) ζ\zeta is a non-negative smooth even function.
(2) For each x>0,x>0, ζ(x)0\zeta^{\prime}(x)\leq 0.
(3) 02πrζ(r)𝑑r=1.\int_{0}^{\infty}2\pi r\zeta(r)dr=1.
(4) There exists a constant CζC_{\zeta} such that ζ(r)Cζζ(r).\zeta(r)\leq-C_{\zeta}\zeta^{\prime}(r).

Define Gn:=ζnζnG_{n}:=\zeta_{n}*\zeta_{n}, which also are smooth mollifiers. It is straightforward to show that Gn(x)=mn2G(mn|x|)G_{n}(x)=m_{n}^{2}G(m_{n}|x|) for certain function GG satisfying Condition 3.1 and being supported on [1,1][-1,1], and

(Gnf)(x)=01mnmn2G(mnr)[Br(x)f𝑑S]𝑑r.(G_{n}*f)(x)=\int_{0}^{\frac{1}{m_{n}}}m_{n}^{2}G(m_{n}r)\left[\int_{\partial B_{r}(x)}fdS\right]dr. (3.2)
Lemma 3.1.

Condition 1.1 implies Condtion 1.2.

Proof.

Let

F(r,x)=12πrBr(x)𝒩(y)𝑑S.F(r,x)=\frac{1}{2\pi r}\int_{\partial B_{r}(x)}\mathcal{N}(y)dS.

Then for 0<r<120<r<\frac{1}{2} and x=(x1,x2)[12,12)2,|x|rx=(x_{1},x_{2})\in\left[-\frac{1}{2},\frac{1}{2}\right)^{2},|x|\neq r,

rF(r,x)=12π02πr𝒩(x1+rcosθ,x2+rsinθ)dθ\displaystyle\partial_{r}F(r,x)=\frac{1}{2\pi}\int_{0}^{2\pi}\partial_{r}\mathcal{N}(x_{1}+r\cos\theta,x_{2}+r\sin\theta)d\theta (3.3)
=12π02π𝒩(x1+rcosθ,x2+rsinθ)(cosθ,sinθ)𝑑θ\displaystyle=\frac{1}{2\pi}\int_{0}^{2\pi}\nabla\mathcal{N}(x_{1}+r\cos\theta,x_{2}+r\sin\theta)\cdot(\cos\theta,\sin\theta)d\theta
=12πrBr(x)𝒩(y)n𝑑S=12πrBr(x)Δ𝒩(x)𝑑x=r212πrχr>|x|,\displaystyle=\frac{1}{2\pi r}\int_{\partial B_{r}(x)}\nabla\mathcal{N}(y)\cdot\vec{n}dS=\frac{1}{2\pi r}\int_{B_{r}(x)}\Delta\mathcal{N}(x)dx=\frac{r}{2}-\frac{1}{2\pi r}\chi_{r>|x|},

where we used (1.1). It’s easy to check limr0+F(r,x)=𝒩(x)\lim_{r\to 0+}F(r,x)=\mathcal{N}(x). Hence, for x[12,12)2x\in\left[-\frac{1}{2},\frac{1}{2}\right)^{2} and r[0,12)r\in[0,\frac{1}{2}), integrate (3.3) along [0,r][0,r], then we have

F(r,x)=𝒩(x)+r2412πmax{log(r)log(|x|),0).F(r,x)=\mathcal{N}(x)+\frac{r^{2}}{4}-\frac{1}{2\pi}\max\{\log(r)-\log(|x|),0).

By (3.2),

(Gn𝒩)(x)𝒩(x)=01mn2πrmn2G(mnr)[F(r,x)𝒩(x)]𝑑r\displaystyle(G_{n}*\mathcal{N})(x)-\mathcal{N}(x)=\int_{0}^{\frac{1}{m_{n}}}2\pi rm_{n}^{2}G(m_{n}r)\big{[}F(r,x)-\mathcal{N}(x)\big{]}dr (3.4)
=01mnrmn2G(mnr)max{log(r)log(|x|),0)𝑑r+1201mnπr3mn2G(mnr)𝑑r.\displaystyle=-\int_{0}^{\frac{1}{m_{n}}}rm_{n}^{2}G(m_{n}r)\max\{\log(r)-\log(|x|),0)dr+\frac{1}{2}\int_{0}^{\frac{1}{m_{n}}}\pi r^{3}m_{n}^{2}G(m_{n}r)dr.

Since 01mnπr3mn2G(mnr)𝑑r=mn201πr3G(r)𝑑r\int_{0}^{\frac{1}{m_{n}}}\pi r^{3}m_{n}^{2}G(m_{n}r)dr=m_{n}^{-2}\int_{0}^{1}\pi r^{3}G(r)dr, then there exists a constant CC such that

(Gn𝒩)(x)𝒩(x)+Cmn2,x2.(G_{n}*\mathcal{N})(x)\leq\mathcal{N}(x)+\frac{C}{m_{n}^{2}},\quad\forall x\notin\mathbb{Z}^{2}. (3.5)

By (1.5), for 0<r<120<r<\frac{1}{2}, F(r,0)12πlogr+CF(r,0)\leq-\frac{1}{2\pi}\log r+C is well defined. Then also due to (3.2),

(Gn𝒩)(0)=01mn2πrmn2G(mnr)F(r,0)𝑑r12πlog(mn)+C.\displaystyle(G_{n}*\mathcal{N})(0)=\int_{0}^{\frac{1}{m_{n}}}2\pi rm_{n}^{2}G(m_{n}r)F(r,0)dr\leq\frac{1}{2\pi}\log(m_{n})+C.

Hence,

e(ζnρn(0))=12𝕋2(Gn𝒩)(xy)ρn(0,dx)ρn(0,dy)\displaystyle e(\zeta_{n}*\rho_{n}(0))=\frac{1}{2}\int_{\mathbb{T}^{2}}(G_{n}*\mathcal{N})(x-y)\rho_{n}(0,dx)\rho_{n}(0,dy)
=12n2i,j=1n(Gn𝒩)(Xi(0)Xj(0))\displaystyle=\frac{1}{2n^{2}}\sum_{i,j=1}^{n}(G_{n}*\mathcal{N})(X_{i}(0)-X_{j}(0))
=12n(Gn𝒩)(0)+12n2ijn(Gn𝒩)(Xi(0)Xj(0))\displaystyle=\frac{1}{2n}(G_{n}*\mathcal{N})(0)+\frac{1}{2n^{2}}\sum_{i\neq j}^{n}(G_{n}*\mathcal{N})(X_{i}(0)-X_{j}(0))
12n(Gn𝒩)(0)+12n2ijn𝒩(Xi(0)Xj(0))+C2mn2(by (3.5))\displaystyle\leq\frac{1}{2n}(G_{n}*\mathcal{N})(0)+\frac{1}{2n^{2}}\sum_{i\neq j}^{n}\mathcal{N}(X_{i}(0)-X_{j}(0))+\frac{C}{2m_{n}^{2}}\quad(\text{by (\ref{equ_gnn3})})
=12n(Gn𝒩)(0)+e0(ρn(0))+C2mn2e0(ρn(0))+log(mn)2πn+C2n+C2mn2,\displaystyle=\frac{1}{2n}(G_{n}*\mathcal{N})(0)+e_{0}(\rho_{n}(0))+\frac{C}{2m_{n}^{2}}\leq e_{0}(\rho_{n}(0))+\frac{\log(m_{n})}{2\pi n}+\frac{C}{2n}+\frac{C}{2m_{n}^{2}},

and the conclusion follows. ∎

3.2 Convolution estimations

Lemma 3.2.

There exists a constant C0C_{0} such that for x[12,12)2x\in\left[-\frac{1}{2},\frac{1}{2}\right)^{2}, |x||(Gn𝒦)(x)𝒦(x)|C0G(mn|x|).|x||(G_{n}*\mathcal{K})(x)-\mathcal{K}(x)|\leq C_{0}G(m_{n}|x|).

Proof.

Take the gradient of (3.4), and we have

(Gn𝒩)(x)𝒩(x)=x|x|2min{|x|,1mn}1mnmn2rG(mnr)𝑑r.\nabla(G_{n}*\mathcal{N})(x)-\nabla\mathcal{N}(x)=\frac{x}{|x|^{2}}\int_{\min\{|x|,\frac{1}{m_{n}}\}}^{\frac{1}{m_{n}}}m_{n}^{2}rG(m_{n}r)dr.

So

|x|[(Gn𝒩)(x)𝒩(x)]=x|x|min{mn|x|,1}1rG(r)𝑑r.|x|\big{[}\nabla(G_{n}*\mathcal{N})(x)-\nabla\mathcal{N}(x)\big{]}=\frac{x}{|x|}\int_{\min\{m_{n}|x|,1\}}^{1}rG(r)dr. (3.6)

By (4) of Condition 3.1, there exists CG>0C_{G}>0 such that

s1rG(r)𝑑rCGs1[rG(r)G(r)]𝑑r+CGs1G(r)𝑑r\displaystyle\int_{s}^{1}rG(r)dr\leq C_{G}\int_{s}^{1}\left[-rG^{\prime}(r)-G(r)\right]dr+C_{G}\int_{s}^{1}G(r)dr
CGsG(s)CG2s1G(r)𝑑r(CG+CG2)G(s),\displaystyle\leq C_{G}sG(s)-C_{G}^{2}\int_{s}^{1}G^{\prime}(r)dr\leq\left(C_{G}+C_{G}^{2}\right)G(s),

which means there exists C0C_{0} such that s1rG(r)𝑑rC0G(s).\int_{s}^{1}rG(r)dr\leq C_{0}G(s). By (3.6) and noting that 𝒦=𝒩\mathcal{K}=-\nabla^{\perp}\mathcal{N}, we conclude the proof. ∎

Similar to the way of obtaining (2.6), for φC1(𝕋2;2)\varphi\in C^{1}(\mathbb{T}^{2};\mathbb{R}^{2}),

φ,γ(Gn𝒦γ)=12(𝕋2)2\𝑫[φ(x)φ(y)](Gn𝒦)(xy)γ(dx)γ(dy),\left\langle\varphi,\gamma(G_{n}*\mathcal{K}*\gamma)\right\rangle=\frac{1}{2}\int_{(\mathbb{T}^{2})^{2}\backslash\boldsymbol{D}}[\varphi(x)-\varphi(y)]\cdot(G_{n}*\mathcal{K})(x-y)\gamma(dx)\gamma(dy), (3.7)

so we can estimate the difference between 𝐑(γ)\mathbf{R}(\gamma) and γ(Gn𝒦γ)\gamma(G_{n}*\mathcal{K}*\gamma).

Lemma 3.3.

For any R>0R>0, there exists a sequence cn0c_{n}\downarrow 0 such that for each γ𝒫(𝕋2)\gamma\in\mathcal{P}(\mathbb{T}^{2}), satisfying e(γ)R<e(\gamma)\leq R<\infty,

|Gn𝒩γ,𝐑(γ)|cnζnγ22.\big{|}\left\langle\nabla G_{n}*\mathcal{N}*\gamma,\mathbf{R}(\gamma)\right\rangle\big{|}\leq c_{n}\|\zeta_{n}*\gamma\|_{2}^{2}.
Proof.

Step 1.
First, we firstly prove that for each η,γ𝒫(𝕋2)\eta,\gamma\in\mathcal{P}(\mathbb{T}^{2}), there exists a constant C1C_{1} not dependent on η,γ,k\eta,\gamma,k and nn such that for 1kmn321\leq k\leq\frac{m_{n}-3}{2},

|Gn𝒩γ,div𝐑(η)div[η(Gn𝒦η)]|\displaystyle\big{|}\left\langle G_{n}*\mathcal{N}*\gamma,{\rm div}\,\mathbf{R}(\eta)-{\rm div}\,[\eta(G_{n}*\mathcal{K}*\eta)]\right\rangle\big{|} (3.8)
C1[1k2+supzγ(Bk+1mn(z))]ζnη22.\displaystyle\leq C_{1}\bigg{[}\frac{1}{k^{2}}+\sup_{z}\gamma\left(B_{\frac{k+1}{m_{n}}}(z)\right)\bigg{]}\|\zeta_{n}*\eta\|_{2}^{2}.

Noticing that Gn(x)=0G_{n}(x)=0 if |x|1mn|x|\geq\frac{1}{m_{n}} and

(𝕋2)2Gn(xy)η(dx)η(dy)=ζnη22,\int_{(\mathbb{T}^{2})^{2}}G_{n}(x-y)\eta(dx)\eta(dy)=\|\zeta_{n}*\eta\|_{2}^{2},

together with (2.4), (3.7) and Lemma 3.2, it’s enough to show there exists a constant C1C_{1} such that for each xy,x\neq y, r(x,y)<1mnr(x,y)<\frac{1}{m_{n}},

|(Gn𝒩γ)(y)(Gn𝒩γ)(x)|r(x,y)C1mn2[1k2+supzγ(Bk+1mn(z))].\frac{|\nabla(G_{n}*\mathcal{N}*\gamma)(y)-\nabla(G_{n}*\mathcal{N}*\gamma)(x)|}{r(x,y)}\leq C_{1}m_{n}^{2}\bigg{[}\frac{1}{k^{2}}+\sup_{z}\gamma\left(B_{\frac{k+1}{m_{n}}}(z)\right)\bigg{]}. (3.9)

Since 𝒩\mathcal{N} can be seen as a periodic function, without loss of generality, we assume xi<yi<xi+12x_{i}<y_{i}<x_{i}+\frac{1}{2} (i=1,2)(i=1,2), so that r(x,y)=|xy|r(x,y)=|x-y|. Let z0=x+y2z_{0}=\frac{x+y}{2}. Then

|(Gn𝒩γ)(y)(Gn𝒩γ)(x)||xy|\displaystyle\frac{|\nabla(G_{n}*\mathcal{N}*\gamma)(y)-\nabla(G_{n}*\mathcal{N}*\gamma)(x)|}{|x-y|} (3.10)
Bk+1mn(z0)|(Gn𝒩)(yz)(Gn𝒩)(xz)||xy|γ(dz)\displaystyle\leq\int_{B_{\frac{k+1}{m_{n}}}(z_{0})}\frac{|(G_{n}*\nabla\mathcal{N})(y-z)-(G_{n}*\nabla\mathcal{N})(x-z)|}{|x-y|}\gamma(dz)
+Bk+1mn(z0)c|(Gn𝒩)(yz)(Gn𝒩)(xz)||xy|γ(dz)\displaystyle+\int_{B_{\frac{k+1}{m_{n}}}(z_{0})^{c}}\frac{|(G_{n}*\nabla\mathcal{N})(y-z)-(G_{n}*\nabla\mathcal{N})(x-z)|}{|x-y|}\gamma(dz)
supzBk+1mn(z0)|(Gn𝒩)(yz)(Gn𝒩)(xz)||xy|γ(Bk+1mn(z0))\displaystyle\leq\sup_{z\in B_{\frac{k+1}{m_{n}}}(z_{0})}\frac{|(G_{n}*\nabla\mathcal{N})(y-z)-(G_{n}*\nabla\mathcal{N})(x-z)|}{|x-y|}\gamma\left(B_{\frac{k+1}{m_{n}}}(z_{0})\right)
+supzBk+1mn(z0)|𝒩(yz)𝒩(xz)||xy|,\displaystyle+\sup_{z\notin B_{\frac{k+1}{m_{n}}}(z_{0})}\frac{|\nabla\mathcal{N}(y-z)-\nabla\mathcal{N}(x-z)|}{|x-y|},

where we used (3.6) to obtain

𝒩(x)=Gn𝒩(x), if r(x,2)>1mn.\nabla\mathcal{N}(x)=G_{n}*\nabla\mathcal{N}(x),\text{ if }r(x,\mathbb{Z}^{2})>\frac{1}{m_{n}}.

As mentioned in [28], i𝒩W˙1,(𝕋2)\partial_{i}\mathcal{N}\in\dot{W}^{-1,\infty}(\mathbb{T}^{2}), i.e. there exist Ai,jL(𝕋2)(i,j=1,2)A_{i,j}\in L^{\infty}(\mathbb{T}^{2})(i,j=1,2) such that i𝒩=j=1,2jAi,j\partial_{i}\mathcal{N}=\sum_{j=1,2}\partial_{j}A_{i,j}. Therefore,

[(Gn𝒩)(yz)(Gn𝒩)(xz)]i=j[jGnAi,j(yz)jGnAi,j(xz)],[(G_{n}*\nabla\mathcal{N})(y-z)-(G_{n}*\nabla\mathcal{N})(x-z)]_{i}=\sum_{j}\left[\partial_{j}G_{n}*A_{i,j}(y-z)-\partial_{j}G_{n}*A_{i,j}(x-z)\right],

So

|(Gn𝒩)i(yz)(Gn𝒩)i(xz)||xy|\displaystyle\frac{|(G_{n}*\nabla\mathcal{N})_{i}(y-z)-(G_{n}*\nabla\mathcal{N})_{i}(x-z)|}{|x-y|}
B2mn(z0z)j|jGn(yzw)jGn(xzw)||xy|Ai,j(w)𝑑w\displaystyle\leq\int_{B_{\frac{2}{m_{n}}}(z_{0}-z)}\frac{\sum_{j}|\partial_{j}G_{n}(y-z-w)-\partial_{j}G_{n}(x-z-w)|}{|x-y|}A_{i,j}(w)dw
Csup|2Gn|mn2Cmn2.\displaystyle\leq\frac{C\sup|\nabla^{2}G_{n}|}{m_{n}^{2}}\leq Cm_{n}^{2}.

Since |xz0|=|yz0|<12mn|x-z_{0}|=|y-z_{0}|<\frac{1}{2m_{n}}, for zBk+1mnc(z0)z\in B^{c}_{\frac{k+1}{m_{n}}}(z_{0}) we have

x,yBk+3/2mn(z)\B¯k+1/2mn(z)B1/2(z)\B¯kmn(z),x,y\in B_{\frac{k+3/2}{m_{n}}}(z)\backslash\overline{B}_{\frac{k+1/2}{m_{n}}}(z)\subset B_{1/2}(z)\backslash\overline{B}_{\frac{k}{m_{n}}}(z),

and thus by (1.5),

|𝒩(yz)𝒩(xz)||xy|sup12>|w|>kmn|2𝒩(w)|Cmn2k2|xy|.|\nabla\mathcal{N}(y-z)-\nabla\mathcal{N}(x-z)|\leq|x-y|\sup_{\frac{1}{2}>|w|>\frac{k}{m_{n}}}|\nabla^{2}\mathcal{N}(w)|\leq C\frac{m_{n}^{2}}{k^{2}}|x-y|.

Therefore (3.10) can deduce (3.9) and we arrive at (3.8).
Step 2.
Take η=γ\eta=\gamma in (3.8). Noting that (Gn𝒦γ)(Gn𝒩γ)=0(G_{n}*\mathcal{K}*\gamma)\cdot(G_{n}*\nabla\mathcal{N}*\gamma)=0, for 1kmn321\leq k\leq\frac{m_{n}-3}{2}, we have

|Gn𝒩γ,𝐑(γ)|C1[1k2+supzγ(Bk+1mn(z))]ζnγ22.\big{|}\left\langle\nabla G_{n}*\mathcal{N}*\gamma,\mathbf{R}(\gamma)\right\rangle\big{|}\leq C_{1}\bigg{[}\frac{1}{k^{2}}+\sup_{z}\gamma\left(B_{\frac{k+1}{m_{n}}}(z)\right)\bigg{]}\|\zeta_{n}*\gamma\|_{2}^{2}.

By (C.9),

supzγ(Bk+1mn(z))(C𝒩+4πe(γ)log(mn/2)log(k+1))12.\sup_{z}\gamma\left(B_{\frac{k+1}{m_{n}}}(z)\right)\leq\bigg{(}\frac{C_{\mathcal{N}}+4\pi e(\gamma)}{\log(m_{n}/2)-\log(k+1)}\bigg{)}^{\frac{1}{2}}.

Taking

cn=C1min1k<mn32[1k2+(C𝒩+4πRlog(mn/2)log(k+1))12],c_{n}=C_{1}\min_{1\leq k<\frac{m_{n}-3}{2}}\left[\frac{1}{k^{2}}+\bigg{(}\frac{C_{\mathcal{N}}+4\pi R}{\log(m_{n}/2)-\log(k+1)}\bigg{)}^{\frac{1}{2}}\right],

then

|Gn𝒩γ,𝐑(γ)|cnζnγ22.\big{|}\left\langle\nabla G_{n}*\mathcal{N}*\gamma,\mathbf{R}(\gamma)\right\rangle\big{|}\leq c_{n}\|\zeta_{n}*\gamma\|_{2}^{2}.

Through a simple calculation, we can check limncn=0.\lim_{n\to\infty}c_{n}=0.

Remark.

The rate for mnm_{n}\to\infty is not needed in the proof of Lemmas 3.2 and 3.3, i.e., we haven’t used the fact that nmn2nm_{n}^{-2}\to\infty.

We also need a generalised version of Ladyzhenskaya’s inequality, which is often used to study two-dimensional Navier-Stokes equation.

Lemma 3.4.

There exists C1,C2C_{1},C_{2} such that for each γ,η𝒫(𝕋2)L2(𝕋2)\gamma,\eta\in\mathcal{P}(\mathbb{T}^{2})\cap L^{2}(\mathbb{T}^{2}),

𝒦(γη)42=𝒩(γη)42C1γη2,\|\mathcal{K}*(\gamma-\eta)\|_{4}^{2}=\|\nabla\mathcal{N}*(\gamma-\eta)\|_{4}^{2}\leq C_{1}\|\gamma-\eta\|_{2},
𝕋2|𝒩γ|2𝑑γ=𝕋2|𝒦γ|2𝑑γC2γ122.\int_{\mathbb{T}^{2}}|\nabla\mathcal{N}*\gamma|^{2}d\gamma=\int_{\mathbb{T}^{2}}|\mathcal{K}*\gamma|^{2}d\gamma\leq C_{2}\|\gamma-1\|_{2}^{2}.
Proof.

Let ϕCc(2)\phi\in C^{\infty}_{c}(\mathbb{R}^{2}) be a radial function such that 0ϕ1,ϕ(x)=10\leq\phi\leq 1,\phi(x)=1 for |x|12|x|\leq\frac{1}{2} and ϕ(x)=0\phi(x)=0 for |x|1.|x|\geq 1. We define for 12>B>0,\frac{1}{2}>B>0, ϕB(x)=ϕ(x/B)\phi_{B}(x)=\phi(x/B), and 𝒦1,B=ϕB𝒦,𝒦2,B=(1ϕB)𝒦.\mathcal{K}_{1,B}=\phi_{B}\mathcal{K},\mathcal{K}_{2,B}=(1-\phi_{B})\mathcal{K}. We start with proving that for 1p<2,q>2,1\leq p<2,q>2, there exist constants Cq,CpC_{q},C^{\prime}_{p}, such that

𝒦2,BqCqB2q1,𝒦1,BpCpB2p1.\|\mathcal{K}_{2,B}\|_{q}\leq C_{q}B^{\frac{2}{q}-1},\quad\|\mathcal{K}_{1,B}\|_{p}\leq C^{\prime}_{p}B^{\frac{2}{p}-1}. (3.11)

By (1.5), we can find C0>0C_{0}>0 such that for each xB12((0,0))\{(0,0)},x\in B_{\frac{1}{2}}((0,0))\backslash\{(0,0)\},

|𝒦(x)|C0|x|1,|𝒦(x)|C0|x|2.|\mathcal{K}(x)|\leq C_{0}|x|^{-1},\quad|\nabla\mathcal{K}(x)|\leq C_{0}|x|^{-2}. (3.12)

So there exists CqC_{q} such that

𝒦2,BqC0(|x|>B2|x|q𝑑x)1/qCqB2/q1.\|\mathcal{K}_{2,B}\|_{q}\leq C_{0}\left(\int_{|x|>\frac{B}{2}}|x|^{-q}dx\right)^{1/q}\leq C_{q}B^{2/q-1}.

Also we have

𝒦1,BppC0p[1/2,1/2]2ϕ(x/B)p|x|p𝑑xC0pB2p|y|1ϕ(y)p|y|p𝑑y.\|\mathcal{K}_{1,B}\|_{p}^{p}\leq C_{0}^{p}\int_{[-1/2,1/2]^{2}}\phi(x/B)^{p}|x|^{-p}dx\leq C_{0}^{p}B^{2-p}\int_{|y|\leq 1}\phi(y)^{p}|y|^{-p}dy.

Let Cp=C0(|y|1|y|p𝑑y)1pC^{\prime}_{p}=C_{0}\left(\int_{|y|\leq 1}|y|^{-p}dy\right)^{\frac{1}{p}}. Then we have

𝒦1,BpCpB2p1.\|\mathcal{K}_{1,B}\|_{p}\leq C^{\prime}_{p}B^{\frac{2}{p}-1}.

Turn to the proof of the desired inequalities. A consequence of (3.11) along with Young’s inequality implies that there exists a constant CC, such that

(ϕB𝒦)(γη)4𝒦1,B43γη2CB12γη2,\|(\phi_{B}\mathcal{K})*(\gamma-\eta)\|_{4}\leq\|\mathcal{K}_{1,B}\|_{{\frac{4}{3}}}\|\gamma-\eta\|_{2}\leq CB^{\frac{1}{2}}\|\gamma-\eta\|_{2},

and

[(1ϕB)𝒦](γη)4𝒦2,B4γη1C/B12.\|[(1-\phi_{B})\mathcal{K}]*(\gamma-\eta)\|_{4}\leq\|\mathcal{K}_{2,B}\|_{4}\|\gamma-\eta\|_{1}\leq C/B^{\frac{1}{2}}.

If γη2>2\|\gamma-\eta\|_{2}>2, we take B=γη21B=\|\gamma-\eta\|_{2}^{-1} and have

𝒦(γη)4(ϕB𝒦)(γη)4+[(1ϕB)𝒦](γη)42Cγη212.\|\mathcal{K}*(\gamma-\eta)\|_{4}\leq\|(\phi_{B}\mathcal{K})*(\gamma-\eta)\|_{4}+\|[(1-\phi_{B})\mathcal{K}]*(\gamma-\eta)\|_{4}\leq 2C\|\gamma-\eta\|_{2}^{\frac{1}{2}}.

If γη22\|\gamma-\eta\|_{2}\leq 2, by Young’s inequality,

𝒦(γη)4𝒦43γη22𝒦43γη212,\|\mathcal{K}*(\gamma-\eta)\|_{4}\leq\|\mathcal{K}\|_{{\frac{4}{3}}}\|\gamma-\eta\|_{2}\leq\sqrt{2}\|\mathcal{K}\|_{{\frac{4}{3}}}\|\gamma-\eta\|_{2}^{\frac{1}{2}},

in which 𝒦43<\|\mathcal{K}\|_{\frac{4}{3}}<\infty due to (1.5). So we can pick C1=max{2𝒦43,2C}C_{1}=\max\left\{\sqrt{2}\|\mathcal{K}\|_{{\frac{4}{3}}},2C\right\} such that

𝒦(γη)42C1γη2.\|\mathcal{K}*(\gamma-\eta)\|_{4}^{2}\leq C_{1}\|\gamma-\eta\|_{2}.

To obtain the second inequality, noting that 𝕋2𝒦(x)𝑑x=0\int_{\mathbb{T}^{2}}\mathcal{K}(x)dx=0 and by young’s inequality, we have

𝒦γ2=𝒦(γ1)2𝒦1γ12.\|\mathcal{K}*\gamma\|_{2}=\|\mathcal{K}*(\gamma-1)\|_{2}\leq\|\mathcal{K}\|_{1}\|\gamma-1\|_{2}.

Hence by Holder’s inequality,

𝕋2|𝒦γ|2𝑑γ=𝕋2|𝒦γ|2𝑑x+𝕋2|𝒦γ|2d(γ1)𝒦γ22+𝒦γ42γ12\displaystyle\int_{\mathbb{T}^{2}}|\mathcal{K}*\gamma|^{2}d\gamma=\int_{\mathbb{T}^{2}}|\mathcal{K}*\gamma|^{2}dx+\int_{\mathbb{T}^{2}}|\mathcal{K}*\gamma|^{2}d(\gamma-1)\leq\|\mathcal{K}*\gamma\|_{2}^{2}+\|\mathcal{K}*\gamma\|_{4}^{2}\|\gamma-1\|_{2}
(𝒦12+C1)γ122.\displaystyle\leq(\|\mathcal{K}\|_{1}^{2}+C_{1})\|\gamma-1\|_{2}^{2}.

Corollary 3.5.

For each δ>0\delta>0, γ𝒫(𝕋2)\gamma\in\mathcal{P}(\mathbb{T}^{2}), smooth mollifier JJ and γ\gamma-measurable function φ\varphi satisfying 𝕋2|φ|2𝑑γ<\int_{\mathbb{T}^{2}}|\varphi|^{2}d\gamma<\infty, one has

|𝕋2φ(JJ𝒦γ)𝑑γ|δJγ122+C14δ𝕋2|φ|2𝑑γ,\left|\int_{\mathbb{T}^{2}}\varphi\cdot(J*J*\mathcal{K}*\gamma)d\gamma\right|\leq\delta\|J*\gamma-1\|_{2}^{2}+\frac{C_{1}}{4\delta}\int_{\mathbb{T}^{2}}|\varphi|^{2}d\gamma,

where C1C_{1} is the constant in Lemma 3.4.

Proof.

By Jensen’s inequality

|JJ𝒦γ|2(x)(J|J𝒦γ|2)(x).|J*J*\mathcal{K}*\gamma|^{2}(x)\leq(J*|J*\mathcal{K}*\gamma|^{2})(x).

Then using abδC1a2+C14δb2ab\leq\frac{\delta}{C_{1}}a^{2}+\frac{C_{1}}{4\delta}b^{2} and by Lemma 3.4, we have

|𝕋2φ(JJ𝒦γ)𝑑γ|δC1𝕋2|JJ𝒦γ|2𝑑γ+C14δ𝕋2|φ|2𝑑γ\displaystyle\left|\int_{\mathbb{T}^{2}}\varphi\cdot(J*J*\mathcal{K}*\gamma)d\gamma\right|\leq\frac{\delta}{C_{1}}\int_{\mathbb{T}^{2}}|J*J*\mathcal{K}*\gamma|^{2}d\gamma+\frac{C_{1}}{4\delta}\int_{\mathbb{T}^{2}}|\varphi|^{2}d\gamma
δC1𝕋2|J𝒦γ|2d(Jγ)+C14δ𝕋2|φ|2𝑑γ\displaystyle\leq\frac{\delta}{C_{1}}\int_{\mathbb{T}^{2}}|J*\mathcal{K}*\gamma|^{2}d(J*\gamma)+\frac{C_{1}}{4\delta}\int_{\mathbb{T}^{2}}|\varphi|^{2}d\gamma
δJγ122+C14δ𝕋2|φ|2𝑑γ.\displaystyle\leq\delta\|J*\gamma-1\|_{2}^{2}+\frac{C_{1}}{4\delta}\int_{\mathbb{T}^{2}}|\varphi|^{2}d\gamma.

3.3 Proof of Lemma 2.3

We need a generalization of the Doob submartingale inequality.

Lemma 3.6.

If M(t)M(t) is a positive continuous local martingale, then for each l,l\in\mathbb{R},

P(sup0tTlogM(t)l)𝔼M(0)el.P\left(\sup_{0\leq t\leq T}\log M(t)\geq l\right)\leq\frac{\mathbb{E}M(0)}{e^{l}}.
Proof.

Since M(t)M(t) is a positive local martingale, it’s a supermartingale. Let τ=inf{t:M(t)>el}T\tau=\inf\{t:M(t)>e^{l}\}\wedge T. Then M(tτ)M(t\wedge\tau) is a non-negative supermartingale. Hence

elP(sup0tTM(t)el)𝔼(M(τ)χM(τ)el)𝔼(M(τ))𝔼(M(0)).e^{l}P\left(\sup_{0\leq t\leq T}M(t)\geq e^{l}\right)\leq\mathbb{E}(M(\tau)\chi_{M(\tau)\geq e^{l}})\leq\mathbb{E}(M(\tau))\leq\mathbb{E}(M(0)).

Now we are ready to give a quantitative version of Lemma 2.3, which together with Condition 1.2 implies Lemma 2.3.

Lemma 3.7.

There exists constants λ>0\lambda>0 such that for each sequence (ηn)(\eta_{n}) satisfying ηn𝒳n\eta_{n}\in\mathcal{X}_{n} and e(ζnηn)R,e(\zeta_{n}*\eta_{n})\leq R,

lim supn\displaystyle\limsup_{n\to\infty} 1nlogPηn(sup0<tT(e(ζnρn(t))\displaystyle\frac{1}{n}\log P_{\eta_{n}}\bigg{(}\sup_{0<t\leq T}\bigg{(}e(\zeta_{n}*\rho_{n}(t)) (3.13)
+ν20tζnρn(s)122ds)>l)λ(lR)\displaystyle+\frac{\nu}{2}\int_{0}^{t}\|\zeta_{n}*\rho_{n}(s)-1\|_{2}^{2}ds\bigg{)}>l\bigg{)}\leq-\lambda(l-R)

for any ll\in\mathbb{R}.

Proof.

Define 𝒩n:=Gn𝒩\mathcal{N}_{n}:=G_{n}*\mathcal{N}, ωn:=νn(Gn(0)1)\omega_{n}:=\frac{\nu}{n}(G_{n}(0)-1). By Ito’s formula and using the fact Δ(Gn𝒩)=Gn1-\Delta(G_{n}*\mathcal{N})=G_{n}-1, we have

d[12𝒩nρn(t),ρn(t)]=12n2ijd𝒩n(Xi(t)Xj(t))\displaystyle d\left[\frac{1}{2}\left\langle\mathcal{N}_{n}*\rho_{n}(t),\rho_{n}(t)\right\rangle\right]=\frac{1}{2n^{2}}\sum_{i\neq j}d\mathcal{N}_{n}(X_{i}(t)-X_{j}(t))
=12n2ij𝒩n(Xi(t)Xj(t))(dXi(t)dXj(t))+νn2ijΔ𝒩n(Xi(t)Xj(t))dt\displaystyle=\frac{1}{2n^{2}}\sum_{i\neq j}\nabla\mathcal{N}_{n}(X_{i}(t)-X_{j}(t))(dX_{i}(t)-dX_{j}(t))+\frac{\nu}{n^{2}}\sum_{i\neq j}\Delta\mathcal{N}_{n}(X_{i}(t)-X_{j}(t))dt
=1n2ij𝒩n(Xi(t)Xj(t))dXi(t)+νn2ijΔ𝒩n(Xi(t)Xj(t))dt\displaystyle=\frac{1}{n^{2}}\sum_{i\neq j}\nabla\mathcal{N}_{n}(X_{i}(t)-X_{j}(t))dX_{i}(t)+\frac{\nu}{n^{2}}\sum_{i\neq j}\Delta\mathcal{N}_{n}(X_{i}(t)-X_{j}(t))dt
=1n2i=1nj=1nki𝒩n(Xi(t)Xj(t))(𝒦(Xi(t)Xk(t))dt+2νdBi(t))\displaystyle=\frac{1}{n^{2}}\sum_{i=1}^{n}\sum_{j=1}^{n}\sum_{k\neq i}\nabla\mathcal{N}_{n}(X_{i}(t)-X_{j}(t))\left(\mathcal{K}(X_{i}(t)-X_{k}(t))dt+\sqrt{2\nu}dB_{i}(t)\right)
+νn2i,j=1nΔ𝒩n(Xi(t)Xj(t))dtνnΔ𝒩n(0)dt\displaystyle+\frac{\nu}{n^{2}}\sum_{i,j=1}^{n}\Delta\mathcal{N}_{n}(X_{i}(t)-X_{j}(t))dt-\frac{\nu}{n}\Delta\mathcal{N}_{n}(0)dt
=𝒩nρn(t),𝐑(ρn(t))dtνζnρn(t)122dt\displaystyle=\left\langle\nabla\mathcal{N}_{n}*\rho_{n}(t),\mathbf{R}(\rho_{n}(t))\right\rangle dt-\nu\|\zeta_{n}*\rho_{n}(t)-1\|_{2}^{2}dt
+2νn2i,j=1n𝒩n(Xi(t)Xj(t))dBi(t)+ωndt.\displaystyle+\frac{\sqrt{2\nu}}{n^{2}}\sum_{i,j=1}^{n}\nabla\mathcal{N}_{n}(X_{i}(t)-X_{j}(t))dB_{i}(t)+\omega_{n}dt.

After further calculation, for any λ>0,\lambda>0,

exp{nλ[e(ζnρn(t))e(ζnρn(0))+ν0tζnρn(s)122ds\displaystyle\exp\bigg{\{}n\lambda\bigg{[}e(\zeta_{n}*\rho_{n}(t))-e(\zeta_{n}*\rho_{n}(0))+\nu\int_{0}^{t}\|\zeta_{n}*\rho_{n}(s)-1\|_{2}^{2}ds (3.14)
0t𝒩nρn(s),𝐑(ρn(s))dsλν0t|𝒩nρn(s)|2,ρn(s)dsωnt]}\displaystyle-\int_{0}^{t}\left\langle\nabla\mathcal{N}_{n}*\rho_{n}(s),\mathbf{R}(\rho_{n}(s))\right\rangle ds-\lambda\nu\int_{0}^{t}\left\langle|\nabla\mathcal{N}_{n}*\rho_{n}(s)|^{2},\rho_{n}(s)\right\rangle ds-\omega_{n}t\bigg{]}\bigg{\}}

is a positive continuous martingale. By Lemma 3.6, for each ηn𝒳n\eta_{n}\in\mathcal{X}_{n} with e(ζnηn)R,e(\zeta_{n}*\eta_{n})\leq R,

Pηn{sup0<tT[e(ζnρn(t))0t𝒩nρn(s),𝐑(ρn(s))ds\displaystyle P_{\eta_{n}}\bigg{\{}\sup_{0<t\leq T}\bigg{[}e(\zeta_{n}*\rho_{n}(t))-\int_{0}^{t}\left\langle\nabla\mathcal{N}_{n}*\rho_{n}(s),\mathbf{R}(\rho_{n}(s))\right\rangle ds (3.15)
+ν0tζnρn(s)122dsνλ0t𝕋2|𝒩nρn(s)|2ρn(s,dx)ds]>l}\displaystyle+\nu\int_{0}^{t}\|\zeta_{n}*\rho_{n}(s)-1\|_{2}^{2}ds-\nu\lambda\int_{0}^{t}\int_{\mathbb{T}^{2}}|\nabla\mathcal{N}_{n}*\rho_{n}(s)|^{2}\rho_{n}(s,dx)ds\bigg{]}>l\bigg{\}}
enλ(lRωnT).\displaystyle\leq e^{-n\lambda(l-R-\omega_{n}T)}.

By Jensen inequality

|𝒩nρn(t,x)|2=|ζnζn𝒩ρn(t,x)|2(ζn|ζn𝒩ρn(t,)|2)(x),|\nabla\mathcal{N}_{n}*\rho_{n}(t,x)|^{2}=|\zeta_{n}*\zeta_{n}*\nabla\mathcal{N}*\rho_{n}(t,x)|^{2}\leq\left(\zeta_{n}*|\zeta_{n}*\nabla\mathcal{N}*\rho_{n}(t,\cdot)|^{2}\right)(x),

Then by Lemma 3.4, there exists C2>0C_{2}>0 such that

𝕋2|𝒩nρn(t)|2ρn(t,dx)𝕋2|𝒩ζnρn(t)|2(ζnρn)(t,dx)C2ζnρn(t)122,\int_{\mathbb{T}^{2}}|\nabla\mathcal{N}_{n}*\rho_{n}(t)|^{2}\rho_{n}(t,dx)\leq\int_{\mathbb{T}^{2}}|\nabla\mathcal{N}*\zeta_{n}*\rho_{n}(t)|^{2}(\zeta_{n}*\rho_{n})(t,dx)\leq C_{2}\|\zeta_{n}*\rho_{n}(t)-1\|_{2}^{2}, (3.16)

Combining (3.16) with (3.15) and taking λ\lambda such that C2λ=13C_{2}\lambda=\frac{1}{3}, we have

Pηn{sup0<tT(e(ζnρn(t))0t𝒩nρn(s),𝐑(ρn(s))ds\displaystyle P_{\eta_{n}}\bigg{\{}\sup_{0<t\leq T}\bigg{(}e(\zeta_{n}*\rho_{n}(t))-\int_{0}^{t}\left\langle\nabla\mathcal{N}_{n}*\rho_{n}(s),\mathbf{R}(\rho_{n}(s))\right\rangle ds
+2ν30tζnρn(s)122ds)>l}enλ(lRωnT).\displaystyle+\frac{2\nu}{3}\int_{0}^{t}\|\zeta_{n}*\rho_{n}(s)-1\|_{2}^{2}ds\bigg{)}>l\bigg{\}}\leq e^{-n\lambda(l-R-\omega_{n}T)}.

For each ε>0\varepsilon>0, define En,l,εE_{n,l,\varepsilon} as a subset of C([0,T];𝒫(𝕋2))C([0,T];\mathcal{P}(\mathbb{T}^{2})) by

En,l,ε:={ρ:sup0<tT(e(ζnρ(t))0t𝒩nρ(s),𝐑(ρ(s))ds\displaystyle E_{n,l,\varepsilon}:=\bigg{\{}\rho:\sup_{0<t\leq T}\bigg{(}e(\zeta_{n}*\rho(t))-\int_{0}^{t}\left\langle\nabla\mathcal{N}_{n}*\rho(s),\mathbf{R}(\rho(s))\right\rangle ds
+2ν30tζnρ(s)122ds)lε}.\displaystyle+\frac{2\nu}{3}\int_{0}^{t}\|\zeta_{n}*\rho(s)-1\|_{2}^{2}ds\bigg{)}\leq l-\varepsilon\bigg{\}}.

By Lemma 3.3, there exists a sequence cn0c_{n}\downarrow 0 only dependent on ll such that if e(ζnρ(s))le(\zeta_{n}*\rho(s))\leq l, then

|𝒩nρ(s),𝐑(ρ(s))|cnζnρ(s)22=cnζnρ(s)122+cn.\big{|}\left\langle\nabla\mathcal{N}_{n}*\rho(s),\mathbf{R}(\rho(s))\big{|}\right\rangle\leq c_{n}\|\zeta_{n}*\rho(s)\|_{2}^{2}=c_{n}\|\zeta_{n}*\rho(s)-1\|_{2}^{2}+c_{n}. (3.17)

Take nn big enough such that 2ν3cn>ν2\frac{2\nu}{3}-c_{n}>\frac{\nu}{2} and cn<εc_{n}<\varepsilon. Define

τl:=inf{t:e(ζnρ(t))+ν20tζnρ(s)122𝑑s>l}T.\tau_{l}:=\inf\left\{t:e(\zeta_{n}*\rho(t))+\frac{\nu}{2}\int_{0}^{t}\|\zeta_{n}*\rho(s)-1\|_{2}^{2}ds>l\right\}\wedge T.

We claim that if ρEn,l,ε\rho\in E_{n,l,\varepsilon} and e(ζnρ(0))le(\zeta_{n}*\rho(0))\leq l, then τl=T\tau_{l}=T.

To prove it by contradiction, suppose τl<T\tau_{l}<T. Since ρC([0,T];𝒫(𝕋2))\rho\in C([0,T];\mathcal{P}(\mathbb{T}^{2})),

e(ζnρ(τl))+ν20τlζnρ(s)122𝑑s=l.e(\zeta_{n}*\rho(\tau_{l}))+\frac{\nu}{2}\int_{0}^{\tau_{l}}\|\zeta_{n}*\rho(s)-1\|_{2}^{2}ds=l. (3.18)

Noting ρEn,l,ε\rho\in E_{n,l,\varepsilon}, we have

e(ζnρ(τl))0τl𝒩nρ(s),𝐑(ρ(s))𝑑s+2ν30τlζnρ(s)122𝑑slε.e(\zeta_{n}*\rho(\tau_{l}))-\int_{0}^{\tau_{l}}\left\langle\nabla\mathcal{N}_{n}*\rho(s),\mathbf{R}(\rho(s))\right\rangle ds+\frac{2\nu}{3}\int_{0}^{\tau_{l}}\|\zeta_{n}*\rho(s)-1\|_{2}^{2}ds\leq l-\varepsilon. (3.19)

However, (3.18), (3.17) and (3.19) imply

lcne(ζnρ(τl))+(2ν3cn)0τlζnρ(s)122𝑑scnlε,l-c_{n}\leq e(\zeta_{n}*\rho(\tau_{l}))+\left(\frac{2\nu}{3}-c_{n}\right)\int_{0}^{\tau_{l}}\|\zeta_{n}*\rho(s)-1\|_{2}^{2}ds-c_{n}\leq l-\varepsilon,

which is a contradiction. Hence, for each ρEn,l,ε\rho\in E_{n,l,\varepsilon} with e(ζnρ(0))le(\zeta_{n}*\rho(0))\leq l, there exists n1n_{1} such that if n>n1n>n_{1} then QT(ζnρn)l.Q_{T}(\zeta_{n}*\rho_{n})\leq l.

Therefore,

Pηn(QT(ζnρn)>l)Pηn(ζnρn(En,l,ε)c)enλ(lεRωnT).P_{\eta_{n}}\left(Q_{T}(\zeta_{n}*\rho_{n})>l\right)\leq P_{\eta_{n}}\left(\zeta_{n}*\rho_{n}\in\left(E_{n,l,\varepsilon}\right)^{c}\right)\leq e^{-n\lambda(l-\varepsilon-R-\omega_{n}T)}.

Recall nmn2,nm_{n}^{-2}\to\infty, so lim supnωnlimnνmn2G𝒩1n=0\limsup_{n\to\infty}\omega_{n}\leq\lim_{n\to\infty}\frac{\nu m_{n}^{2}\|G\|_{\infty}\|\mathcal{N}\|_{1}}{n}=0. By the arbitrariness of ε\varepsilon the conclusion follows. ∎

4 Regularity of trajectories with finite rate function

In this section, we study the regularity of trajectories with finite rate function to give a more direct expression for the rate function and prove Lemma 2.5. These regularity results are also preparations for the proof of subsequent lemmas in the next section.

4.1 Weighted Sobolev space and Riesz representation

To obtain the explicit form of rate function, we need a notation of weighted Sobolev space Hρ1([0,T]×𝕋2).H^{1}_{\rho}([0,T]\times\mathbb{T}^{2}). For ϕC([0,T]×𝕋2)\phi\in C^{\infty}([0,T]\times\mathbb{T}^{2}) and ρC([0,T];𝒫(𝕋2))\rho\in C([0,T];\mathcal{P}(\mathbb{T}^{2})), define the norm

ϕ1,ρ,T2=0T𝕋2|ϕ(t)|2𝑑ρ(t)𝑑t.\|\phi\|_{1,\rho,T}^{2}=\int_{0}^{T}\int_{\mathbb{T}^{2}}|\nabla\phi(t)|^{2}d\rho(t)dt.

Define Hρ1([0,T]×𝕋2)H^{1}_{\rho}([0,T]\times\mathbb{T}^{2}) as the completion of C0([0,T]×𝕋2)={ϕC([0,T]×𝕋2):0T𝕋2ϕ(t,x)𝑑x𝑑t=0}C_{0}^{\infty}([0,T]\times\mathbb{T}^{2})=\{\phi\in C^{\infty}([0,T]\times\mathbb{T}^{2}):\int_{0}^{T}\int_{\mathbb{T}^{2}}\phi(t,x)dxdt=0\} under 1,ρ,T.\|\cdot\|_{1,\rho,T}. That is a Hilbert space with inner product

p1,p21,ρ,T=14(p1+p21,ρ,T2p1p21,ρ,T2).\left\langle p_{1},p_{2}\right\rangle_{1,\rho,T}=\frac{1}{4}\left(\|p_{1}+p_{2}\|^{2}_{1,\rho,T}-\|p_{1}-p_{2}\|^{2}_{1,\rho,T}\right).

Then for any pHρ1([0,T]×𝕋2)p\in H^{1}_{\rho}([0,T]\times\mathbb{T}^{2}), there exists a function ^p\hat{\nabla}p defined on [0,T]×𝕋2[0,T]\times\mathbb{T}^{2} such that

0T𝕋2|^p(t)|2𝑑ρ(t)𝑑t<,\int_{0}^{T}\int_{\mathbb{T}^{2}}|\hat{\nabla}p(t)|^{2}d\rho(t)dt<\infty,

and

p,ϕ1,ρ,T=0T𝕋2^p(t,x)ϕ(t,x)𝑑ρ(t)𝑑t,ϕC([0,T]×𝕋2).\left\langle p,\phi\right\rangle_{1,\rho,T}=\int_{0}^{T}\int_{\mathbb{T}^{2}}\hat{\nabla}p(t,x)\cdot\nabla\phi(t,x)d\rho(t)dt,\quad\forall\phi\in C^{\infty}([0,T]\times\mathbb{T}^{2}).

The inner product thus can be written as

p1,p21,ρ,T=0T𝕋2^p1(t,x)^p2(t,x)𝑑ρ(t)𝑑t,p1,p2Hρ1([0,T]×𝕋2).\left\langle p_{1},p_{2}\right\rangle_{1,\rho,T}=\int_{0}^{T}\int_{\mathbb{T}^{2}}\hat{\nabla}p_{1}(t,x)\cdot\hat{\nabla}p_{2}(t,x)d\rho(t)dt,\quad\forall p_{1},p_{2}\in H^{1}_{\rho}([0,T]\times\mathbb{T}^{2}).

Recall that

𝔸¯T(ρ):=supϕC([0,T]×𝕋2)(ϕ(T),ρ(T)ϕ(0),ρ(t0)0Ttϕ(s),ρ(s)ds\displaystyle\overline{\mathbb{A}}_{T}(\rho):=\sup_{\phi\in C^{\infty}([0,T]\times\mathbb{T}^{2})}\bigg{(}\left\langle\phi(T),\rho(T)\right\rangle-\left\langle\phi(0),\rho(t_{0})\right\rangle-\int_{0}^{T}\left\langle\partial_{t}\phi(s),\rho(s)\right\rangle ds
ν0TΔϕ(s),ρ(s)ds0Tϕ(s),𝐑(ρ(s))dsν0T𝕋2|ϕ(s)|2dρ(s)ds).\displaystyle-\nu\int_{0}^{T}\left\langle\Delta\phi(s),\rho(s)\right\rangle ds-\int_{0}^{T}\left\langle\nabla\phi(s),\mathbf{R}(\rho(s))\right\rangle ds-\nu\int_{0}^{T}\int_{\mathbb{T}^{2}}|\nabla\phi(s)|^{2}d\rho(s)ds\bigg{)}.

We claim that 𝔸¯T(ρ)<\overline{\mathbb{A}}_{T}(\rho)<\infty means tρνΔρ+div𝐑(ρ)\partial_{t}\rho-\nu\Delta\rho+{\rm div}\,\mathbf{R}(\rho) can be seen as a bounded linear operator on Hρ1([0,T]×𝕋2)H^{1}_{\rho}([0,T]\times\mathbb{T}^{2}). To obtain that, for ϕC([0,T]×𝕋2)\phi\in C^{\infty}([0,T]\times\mathbb{T}^{2}) define

Lρϕ:=tρνΔρ+div𝐑(ρ),ϕ=ϕ(T),ρ(T)ϕ(0),ρ(0)\displaystyle L_{\rho}\phi:=\left\langle\partial_{t}\rho-\nu\Delta\rho+{\rm div}\,\mathbf{R}(\rho),\phi\right\rangle=\left\langle\phi(T),\rho(T)\right\rangle-\left\langle\phi(0),\rho(0)\right\rangle
0Ttϕ(t),ρ(t)𝑑tν0TΔϕ(t),ρ(t)𝑑t0Tϕ(t),𝐑(ρ(t))𝑑t.\displaystyle-\int_{0}^{T}\left\langle\partial_{t}\phi(t),\rho(t)\right\rangle dt-\nu\int_{0}^{T}\left\langle\Delta\phi(t),\rho(t)\right\rangle dt-\int_{0}^{T}\left\langle\nabla\phi(t),\mathbf{R}(\rho(t))\right\rangle dt.

By the definition of 𝔸¯T\overline{\mathbb{A}}_{T}, for a test function ϕ\phi, taking ϕk=kϕ\phi_{k}=k\phi, we have

supk(Lρ(kϕ)νkϕ1,ρ,T2)=(Lρϕ)24νϕ1,ρ,T2𝔸¯T(ρ),\sup_{k}\left(L_{\rho}(k\phi)-\nu\|k\phi\|_{1,\rho,T}^{2}\right)=\frac{(L_{\rho}\phi)^{2}}{4\nu\|\phi\|_{1,\rho,T}^{2}}\leq\overline{\mathbb{A}}_{T}(\rho),

implying LρL_{\rho} is a bounded operator on Hρ1([0,T]×𝕋2)H^{1}_{\rho}([0,T]\times\mathbb{T}^{2}).

By Riesz representation theorem, if 𝔸¯T(ρ)<\overline{\mathbb{A}}_{T}(\rho)<\infty, there exists pHρ1([0,T]×𝕋2)p\in H^{1}_{\rho}([0,T]\times\mathbb{T}^{2}) such that

0T𝕋2|^p(t)|2𝑑ρ(t)𝑑t4ν𝔸¯T(ρ)\int_{0}^{T}\int_{\mathbb{T}^{2}}|\hat{\nabla}p(t)|^{2}d\rho(t)dt\leq 4\nu\overline{\mathbb{A}}_{T}(\rho)

and Lρϕ=ϕ,p1,ρ,TL_{\rho}\phi=\left\langle\phi,p\right\rangle_{1,\rho,T}, i.e. for each ϕC([0,T]×𝕋2),\phi\in C^{\infty}([0,T]\times\mathbb{T}^{2}),

ϕ(T),ρ(T)ϕ(0),ρ(0)=0Ttϕ(r),ρ(r)𝑑r+νstΔϕ(r),ρ(r)𝑑r\displaystyle\left\langle\phi(T),\rho(T)\right\rangle-\left\langle\phi(0),\rho(0)\right\rangle=\int_{0}^{T}\left\langle\partial_{t}\phi(r),\rho(r)\right\rangle dr+\nu\int_{s}^{t}\left\langle\Delta\phi(r),\rho(r)\right\rangle dr (4.1)
+0Tϕ(r)^p(r),ρ(r)𝑑r+0Tϕ(r),𝐑(ρ(r))𝑑r.\displaystyle+\int_{0}^{T}\left\langle\nabla\phi(r)\cdot\hat{\nabla}p(r),\rho(r)\right\rangle dr+\int_{0}^{T}\left\langle\nabla\phi(r),\mathbf{R}(\rho(r))\right\rangle dr.

In addition, by (4.1), taking smooth test function approximating 12ν^p\frac{1}{2\nu}\hat{\nabla}p in the definition of 𝔸¯T(ρ)\overline{\mathbb{A}}_{T}(\rho), we finally have

𝔸¯T(ρ)=14ν0T𝕋2|^p(t)|2𝑑ρ(t)𝑑t.\overline{\mathbb{A}}_{T}(\rho)=\frac{1}{4\nu}\int_{0}^{T}\int_{\mathbb{T}^{2}}|\hat{\nabla}p(t)|^{2}d\rho(t)dt. (4.2)

More properties of weighted Sobolev space is included in Appendix A.

4.2 Production estimations of energy and entropy

In this subsetion we will prove production estimations of energy and entropy.

Lemma 4.1.

Suppose ρC([0,T];𝒫(𝕋2))\rho\in C([0,T];\mathcal{P}(\mathbb{T}^{2})), QT(ρ)<Q_{T}(\rho)<\infty and 𝔸¯T(ρ)<.\overline{\mathbb{A}}_{T}(\rho)<\infty. Then there exists ^p:[0,T]×𝕋2,\hat{\nabla}p:[0,T]\times\mathbb{T}^{2}\mapsto\mathbb{R}, such that (4.2) holds and for each ϕC([0,T]×𝕋2)\phi\in C^{\infty}([0,T]\times\mathbb{T}^{2}),

ϕ(t),ρ(t)ϕ(s),ρ(s)=sttϕ(r),ρ(r)𝑑r+νstΔϕ(r),ρ(r)𝑑r\displaystyle\left\langle\phi(t),\rho(t)\right\rangle-\left\langle\phi(s),\rho(s)\right\rangle=\int_{s}^{t}\left\langle\partial_{t}\phi(r),\rho(r)\right\rangle dr+\nu\int_{s}^{t}\left\langle\Delta\phi(r),\rho(r)\right\rangle dr (4.3)
+stϕ(r)^p(r),ρ(r)𝑑r+stϕ(r),𝐑(ρ(r))𝑑r.\displaystyle+\int_{s}^{t}\left\langle\nabla\phi(r)\cdot\hat{\nabla}p(r),\rho(r)\right\rangle dr+\int_{s}^{t}\left\langle\nabla\phi(r),\mathbf{R}(\rho(r))\right\rangle dr.

In addition, for 0s<tT,0\leq s<t\leq T,

e(ρ(t))e(ρ(s))=νstρ(r)122𝑑r+st𝕋2(𝒩ρ)(r,x)^p(r,x)ρ(r,x)𝑑r.e(\rho(t))-e(\rho(s))=-\nu\int_{s}^{t}\|\rho(r)-1\|_{2}^{2}dr+\int_{s}^{t}\int_{\mathbb{T}^{2}}(\nabla\mathcal{N}*\rho)(r,x)\cdot\hat{\nabla}p(r,x)\rho(r,x)dr. (4.4)
Proof.

Since 𝔸¯T(ρ)<\overline{\mathbb{A}}_{T}(\rho)<\infty, as discussed in the previous subsection, there exists ^p\hat{\nabla}p satisfying (4.2) and (4.1) holds for each ϕC([0,T]×𝕋2)\phi\in C^{\infty}([0,T]\times\mathbb{T}^{2}). Since ρC([0,T];𝒫(𝕋2))\rho\in C([0,T];\mathcal{P}(\mathbb{T}^{2})), it’s standard to see (4.3) holds for each ϕC([0,T]×𝕋2)\phi\in C^{\infty}([0,T]\times\mathbb{T}^{2}) by smooth approximation of truncated functions only supported on [s,t][s,t].

Take a smooth mollifier JJ with compact support and define Jn(x):=n2J(xn)J_{n}(x):=n^{2}J\left(\frac{x}{n}\right). Taking ϕ(y)=Jn(xy)\phi(y)=J_{n}(x-y) in (4.3), noting ρ(t)L2(𝕋2)\rho(t)\in L^{2}(\mathbb{T}^{2}) for almost every tt and by (2.6), we can check that ρ^n(t,x):=(Jnρ)(t,x)\hat{\rho}_{n}(t,x):=(J_{n}*\rho)(t,x) is absolutely continuous with respect to tt and for a.e.-tt

tρ^n(t,x)νΔρ^n(t,x)+[Jndiv(ρ(t)u(t))](x)+[Jndiv(ρ(t)^p(t))](x)=0,\partial_{t}\hat{\rho}_{n}(t,x)-\nu\Delta\hat{\rho}_{n}(t,x)+[J_{n}*{\rm div}\,(\rho(t)u(t))](x)+[J_{n}*{\rm div}\,(\rho(t)\hat{\nabla}p(t))](x)=0, (4.5)

where u(t)=𝒦ρ(t)u(t)=\mathcal{K}*\rho(t).

Now we show e(ρ^n(t))e(\hat{\rho}_{n}(t)) is absolutely continuous with respect to tt. Notice that ρ^n\|\hat{\rho}_{n}\|_{\infty} and tρ^n\|\partial_{t}\hat{\rho}_{n}\|_{\infty} are bounded by constants which only depend on JnJ_{n}, then for 0s<tT,0\leq s<t\leq T,

|e(ρ^n(t))e(ρ^n(s))|\displaystyle|e(\hat{\rho}_{n}(t))-e(\hat{\rho}_{n}(s))|
=12|(𝕋2)2𝒩(xy)ρ^n(t,x)ρ^n(t,y)𝑑x𝑑y(𝕋2)2𝒩(xy)ρ^n(s,x)ρ^n(s,y)𝑑x𝑑y|\displaystyle=\frac{1}{2}\bigg{|}\int_{(\mathbb{T}^{2})^{2}}\mathcal{N}(x-y)\hat{\rho}_{n}(t,x)\hat{\rho}_{n}(t,y)dxdy-\int_{(\mathbb{T}^{2})^{2}}\mathcal{N}(x-y)\hat{\rho}_{n}(s,x)\hat{\rho}_{n}(s,y)dxdy\bigg{|}
=12|(𝕋2)2𝒩(xy)[ρ^n(t,x)(ρ^n(t,y)ρ^n(s,y))(ρ^n(t,x)ρ^n(s,x))ρ^n(s,y)]𝑑x𝑑y|\displaystyle=\frac{1}{2}\bigg{|}\int_{(\mathbb{T}^{2})^{2}}\mathcal{N}(x-y)\big{[}\hat{\rho}_{n}(t,x)(\hat{\rho}_{n}(t,y)-\hat{\rho}_{n}(s,y))-(\hat{\rho}_{n}(t,x)-\hat{\rho}_{n}(s,x))\hat{\rho}_{n}(s,y)\big{]}dxdy\bigg{|}
(ts)ρ^ntρ^n(𝕋2)2|𝒩(xy)|𝑑x𝑑yCn(ts).\displaystyle\leq(t-s)\|\hat{\rho}_{n}\|_{\infty}\|\partial_{t}\hat{\rho}_{n}\|_{\infty}\int_{(\mathbb{T}^{2})^{2}}|\mathcal{N}(x-y)|dxdy\leq C_{n}(t-s).

Then by directly taking derivative and dominated convergence theorem, we have

te(ρ^n(t,x))=(𝕋2)2𝒩(xy)ρ^n(t,y)tρ^n(t,x)dxdy.\partial_{t}e(\hat{\rho}_{n}(t,x))=\int_{(\mathbb{T}^{2})^{2}}\mathcal{N}(x-y)\hat{\rho}_{n}(t,y)\partial_{t}\hat{\rho}_{n}(t,x)dxdy.

By (4.5) and noticing Δ(JnJn)𝒩=1JnJn\Delta(J_{n}*J_{n})*\mathcal{N}=1-J_{n}*J_{n} as well as (𝕋2)2(JnJn)(xy)ρ(t,dx)ρ(t,dy)1=Jnρ(t)122,\int_{(\mathbb{T}^{2})^{2}}(J_{n}*J_{n})(x-y)\rho(t,dx)\rho(t,dy)-1=\|J_{n}*\rho(t)-1\|_{2}^{2},

e(Jnρ(t))e(Jnρ(s))=νstJnρ(r)122𝑑r\displaystyle e(J_{n}*\rho(t))-e(J_{n}*\rho(s))=-\nu\int_{s}^{t}\|J_{n}*\rho(r)-1\|_{2}^{2}dr (4.6)
+st𝕋2(JnJn𝒩ρ)(r,x)u(r,x)ρ(r,x)𝑑x𝑑r\displaystyle+\int_{s}^{t}\int_{\mathbb{T}^{2}}(J_{n}*J_{n}*\nabla\mathcal{N}*\rho)(r,x)\cdot u(r,x)\rho(r,x)dxdr
+st𝕋2(JnJn𝒩ρ)(r,x)^p(r,x)ρ(r,dx)𝑑s.\displaystyle+\int_{s}^{t}\int_{\mathbb{T}^{2}}(J_{n}*J_{n}*\nabla\mathcal{N}*\rho)(r,x)\cdot\hat{\nabla}p(r,x)\rho(r,dx)ds.

To use (4.6) to show (4.4), we just need to prove that

limnst𝕋2(JnJn𝒩ρ)(r,x)u(r,x)ρ(r,x)𝑑x𝑑r=0,\displaystyle\lim_{n\to\infty}\int_{s}^{t}\int_{\mathbb{T}^{2}}(J_{n}*J_{n}*\nabla\mathcal{N}*\rho)(r,x)\cdot u(r,x)\rho(r,x)dxdr=0, (4.7)
limnst𝕋2(JnJn𝒩ρ)(r,x)^p(r,x)ρ(r,x)𝑑x𝑑r\displaystyle\lim_{n\to\infty}\int_{s}^{t}\int_{\mathbb{T}^{2}}(J_{n}*J_{n}*\nabla\mathcal{N}*\rho)(r,x)\cdot\hat{\nabla}p(r,x)\rho(r,x)dxdr (4.8)
=st𝕋2(𝒩ρ)(r,x)^p(r,x)ρ(r,x)𝑑x𝑑r,\displaystyle=\int_{s}^{t}\int_{\mathbb{T}^{2}}(\nabla\mathcal{N}*\rho)(r,x)\cdot\hat{\nabla}p(r,x)\rho(r,x)dxdr,

and the convergence of rest terms can be proved by Fatou’s Lemma and Jensen’s inequality.

Note that ρ(r)2<\|\rho(r)\|_{2}<\infty for a.e. rr since QT(ρ)<Q_{T}(\rho)<\infty. By Lemma 3.4, for a.e. rr, 𝒩ρ(r)L4(𝕋2)\nabla\mathcal{N}*\rho(r)\in L^{4}(\mathbb{T}^{2}) so that by property of smooth mollifiers,

limn(JnJn𝒩ρ)(r)𝒩ρ(r)4=0.\lim_{n\to\infty}\|(J_{n}*J_{n}*\nabla\mathcal{N}*\rho)(r)-\nabla\mathcal{N}*\rho(r)\|_{4}=0. (4.9)

By Holder’s inequality, for a.e. rr,

|𝕋2(JnJn𝒩ρ)(r,x)u(r,x)ρ(r,x)𝑑x|\displaystyle\left|\int_{\mathbb{T}^{2}}(J_{n}*J_{n}*\nabla\mathcal{N}*\rho)(r,x)\cdot u(r,x)\rho(r,x)dx\right| (4.10)
=|𝕋2(JnJn𝒩ρ𝒩ρ)(r,x)u(r,x)ρ(r,x)𝑑x|\displaystyle=\left|\int_{\mathbb{T}^{2}}(J_{n}*J_{n}*\nabla\mathcal{N}*\rho-\nabla\mathcal{N}*\rho)(r,x)\cdot u(r,x)\rho(r,x)dx\right|
ρ(r)2u(r)4(JnJn𝒩ρ)(r)𝒩ρ(r)40\displaystyle\leq\|\rho(r)\|_{2}\|u(r)\|_{4}\|(J_{n}*J_{n}*\nabla\mathcal{N}*\rho)(r)-\nabla\mathcal{N}*\rho(r)\|_{4}\to 0

where we used the fact that (𝒩ρ)(r,x)u(r,x)=0.(\nabla\mathcal{N}*\rho)(r,x)\cdot u(r,x)=0. So far we have proved that

limn𝕋2(JnJn𝒩ρ)(r,x)u(r,x)ρ(r,x)𝑑x=0 for a.e. r.\lim_{n\to\infty}\int_{\mathbb{T}^{2}}(J_{n}*J_{n}*\nabla\mathcal{N}*\rho)(r,x)\cdot u(r,x)\rho(r,x)dx=0\text{ for a.e. }r.

Apply Lemma 3.4 along with Jensen’s inequality to (4.10),

|𝕋2(JnJn𝒩ρ)(r,x)u(r,x)ρ(r,x)𝑑x|Cρ(r)22.\left|\int_{\mathbb{T}^{2}}(J_{n}*J_{n}*\nabla\mathcal{N}*\rho)(r,x)\cdot u(r,x)\rho(r,x)dx\right|\leq C\|\rho(r)\|_{2}^{2}.

Since QT(ρ)<Q_{T}(\rho)<\infty, by dominated convergence theorem, (4.7) follows.

Turing to (4.8), by Holder’s inequality,

|𝕋2(JnJn𝒩ρ𝒩ρ)(r,x)^p(r,x)ρ(r,x)𝑑x|\displaystyle\bigg{|}\int_{\mathbb{T}^{2}}(J_{n}*J_{n}*\nabla\mathcal{N}*\rho-\nabla\mathcal{N}*\rho)(r,x)\cdot\hat{\nabla}p(r,x)\rho(r,x)dx\bigg{|} (4.11)
(𝕋2|^p(r,x)|2ρ(r,x)𝑑x)12ρ(r)212(JnJn𝒩ρ𝒩ρ)(r)4.\displaystyle\leq\left(\int_{\mathbb{T}^{2}}|\hat{\nabla}p(r,x)|^{2}\rho(r,x)dx\right)^{\frac{1}{2}}\|\rho(r)\|_{2}^{\frac{1}{2}}\|(J_{n}*J_{n}*\nabla\mathcal{N}*\rho-\nabla\mathcal{N}*\rho)(r)\|_{4}.

The condition 𝔸¯T(ρ)<\overline{\mathbb{A}}_{T}(\rho)<\infty along with (4.2) implies 𝕋2|^p(r,x)|2ρ(r,x)𝑑x<\int_{\mathbb{T}^{2}}|\hat{\nabla}p(r,x)|^{2}\rho(r,x)dx<\infty for a.e. rr. Thus by (4.9) the point-wise convergence of 𝕋2(JnJn𝒩ρ)(r,x)^p(r,x)ρ(r,x)𝑑x\int_{\mathbb{T}^{2}}(J_{n}*J_{n}*\nabla\mathcal{N}*\rho)(r,x)\cdot\hat{\nabla}p(r,x)\rho(r,x)dx holds. Finally, by Lemma 3.4 and Jensen’s inequality, the RHS of (4.11) is bounded by 𝕋2|^p(r,x)|2ρ(r,x)𝑑x+Cρ(r)22,\int_{\mathbb{T}^{2}}|\hat{\nabla}p(r,x)|^{2}\rho(r,x)dx+C\|\rho(r)\|_{2}^{2}, which is integrable. By dominated convergence theorem, we reach (4.8). ∎

Let the entropy functional be defined by

S(γ)={𝕋2γ(x)logγ(x)𝑑x,if γ(dx)=γ(x)dx,,otherwise,S(\gamma)=\left\{\begin{aligned} &\int_{\mathbb{T}^{2}}\gamma(x)\log\gamma(x)dx,&\text{if }\gamma(dx)=\gamma(x)dx,\\ &\infty,&\text{otherwise},\end{aligned}\right.

and the Fisher information functional be defined by

I(γ)={𝕋2|γ(x)|2γ(x)𝑑x,if γ(dx)=γ(x)dx and γL1(𝕋2),,otherwise.I(\gamma)=\left\{\begin{aligned} &\int_{\mathbb{T}^{2}}\frac{|\nabla\gamma(x)|^{2}}{\gamma(x)}dx,&\text{if }\gamma(dx)=\gamma(x)dx\text{ and }\nabla\gamma\in L^{1}(\mathbb{T}^{2}),\\ &\infty,&otherwise.\end{aligned}\right.
Lemma 4.2.

If ρ\rho satisfies assumptions in Lemma 4.1, then for each t(0,T]t\in(0,T],

S(ρ(t))2𝔸¯T(ρ)+log(2QT(ρ)νt+1).S(\rho(t))\leq 2\overline{\mathbb{A}}_{T}(\rho)+\log\left(\frac{2Q_{T}(\rho)}{\nu t}+1\right). (4.12)

In addition, for each α>0\alpha>0,

0TtαI(ρ(t))𝑑t<.\int_{0}^{T}t^{\alpha}I(\rho(t))dt<\infty. (4.13)
Proof.

We start with proving (4.12). Still using the smooth mollifiers JnJ_{n} in Lemma 4.1, let u=𝒦ρ,ρ^n=Jnρnu=\mathcal{K}*\rho,\hat{\rho}_{n}=J_{n}*\rho_{n}. By (4.5),

teS(ρ^n(t))=0tr[reS(ρ^n(r))]dr\displaystyle te^{S(\hat{\rho}_{n}(t))}=\int_{0}^{t}\partial_{r}\left[re^{S(\hat{\rho}_{n}(r))}\right]dr (4.14)
=0teS(ρ^n(r))𝑑r+0treS(ρ^n(r))1+log(ρ^n(r)),tρ^n(r)𝑑r\displaystyle=\int_{0}^{t}e^{S(\hat{\rho}_{n}(r))}dr+\int_{0}^{t}re^{S(\hat{\rho}_{n}(r))}\left\langle 1+\log(\hat{\rho}_{n}(r)),\partial_{t}\hat{\rho}_{n}(r)\right\rangle dr
=0teS(ρ^n(r))𝑑rν0treS(ρ^n(r))I(ρ^n(r))𝑑r\displaystyle=\int_{0}^{t}e^{S(\hat{\rho}_{n}(r))}dr-\nu\int_{0}^{t}re^{S(\hat{\rho}_{n}(r))}I(\hat{\rho}_{n}(r))dr
+0t𝕋2reS(ρ^n(r))(Jnρ)(r,x)(Jnρ)(r,x)(Jn(ρ^p))(r,x)𝑑x𝑑r\displaystyle+\int_{0}^{t}\int_{\mathbb{T}^{2}}re^{S(\hat{\rho}_{n}(r))}\frac{(\nabla J_{n}*\rho)(r,x)}{(J_{n}*\rho)(r,x)}\cdot(J_{n}*(\rho\hat{\nabla}p))(r,x)dxdr
+0t𝕋2reS(ρ^n(r))(Jnρ)(r,x)(Jnρ)(r,x)(Jn(ρu))(r,x)𝑑x𝑑r.\displaystyle+\int_{0}^{t}\int_{\mathbb{T}^{2}}re^{S(\hat{\rho}_{n}(r))}\frac{(\nabla J_{n}*\rho)(r,x)}{(J_{n}*\rho)(r,x)}\cdot(J_{n}*(\rho u))(r,x)dxdr.

Due to mean value inequality,

ν24I(ρ^n(r))+ν2𝕋2(Jnρ)(r,x)(Jnρ)(r,x)(Jn(ρ^p))(r,x)𝑑x\displaystyle-\frac{\nu^{2}}{4}I(\hat{\rho}_{n}(r))+\frac{\nu}{2}\int_{\mathbb{T}^{2}}\frac{(\nabla J_{n}*\rho)(r,x)}{(J_{n}*\rho)(r,x)}\cdot(J_{n}*(\rho\hat{\nabla}p))(r,x)dx (4.15)
14𝕋2|Jn(ρ^p)(r,x)(Jnρ)(r,x)|2(Jnρ)(r,x)𝑑x14𝕋2|^p(r,x)|2ρ(r,x)𝑑x,\displaystyle\leq\frac{1}{4}\int_{\mathbb{T}^{2}}\left|\frac{J_{n}*(\rho\hat{\nabla}p)(r,x)}{(J_{n}*\rho)(r,x)}\right|^{2}(J_{n}*\rho)(r,x)dx\leq\frac{1}{4}\int_{\mathbb{T}^{2}}|\hat{\nabla}p(r,x)|^{2}\rho(r,x)dx,

where we used Lemma 8.1.10 of [1] to obtain the last inequality. Since divu=0{\rm div}\,u=0, we have 𝕋2(Jnρ)(r,x)u(r,x)𝑑x=0\int_{\mathbb{T}^{2}}(\nabla J_{n}*\rho)(r,x)\cdot u(r,x)dx=0 and

ν2𝕋2(Jnρ)(r,x)(Jnρ)(r,x)(Jn(ρu))(r,x)𝑑x\displaystyle\frac{\nu}{2}\int_{\mathbb{T}^{2}}\frac{(\nabla J_{n}*\rho)(r,x)}{(J_{n}*\rho)(r,x)}\cdot(J_{n}*(\rho u))(r,x)dx (4.16)
=ν2𝕋2(Jnρ)(r,x)(Jnρ)(r,x)((Jn(ρu))(r,x)(Jnρ)(r,x)u(r,x))(Jnρ)(r,x)𝑑x\displaystyle=\frac{\nu}{2}\int_{\mathbb{T}^{2}}\frac{(\nabla J_{n}*\rho)(r,x)}{(J_{n}*\rho)(r,x)}\cdot\left(\frac{(J_{n}*(\rho u))(r,x)}{(J_{n}*\rho)(r,x)}-u(r,x)\right)(J_{n}*\rho)(r,x)dx
14𝕋2|Jn(ρu)(r,x)(Jnρ)(r,x)u(r,x)|2(Jnρ)(r,x)𝑑x+ν24I(ρ^n(r)).\displaystyle\leq\frac{1}{4}\int_{\mathbb{T}^{2}}\left|\frac{J_{n}*(\rho u)(r,x)}{(J_{n}*\rho)(r,x)}-u(r,x)\right|^{2}(J_{n}*\rho)(r,x)dx+\frac{\nu^{2}}{4}I(\hat{\rho}_{n}(r)).

By Jensen’s inequality

S(Jnρ)logJnρ22=log(Jnρ122+1).S(J_{n}*\rho)\leq\log\|J_{n}*\rho\|_{2}^{2}=\log(\|J_{n}*\rho-1\|_{2}^{2}+1). (4.17)

Applying (4.15), (4.16) (4.17) to (4.14), we have

ν2teS(ρ^n(t))QT(ρ^n)+ν2t\displaystyle\frac{\nu}{2}te^{S(\hat{\rho}_{n}(t))}\leq Q_{T}(\hat{\rho}_{n})+\frac{\nu}{2}t
+140treS(ρ^n(r))𝕋2[|Jn(ρu)(r,x)(Jnρ)(r,x)u(r,x)|2(Jnρ)(r,x)\displaystyle+\frac{1}{4}\int_{0}^{t}re^{S(\hat{\rho}_{n}(r))}\int_{\mathbb{T}^{2}}\bigg{[}\left|\frac{J_{n}*(\rho u)(r,x)}{(J_{n}*\rho)(r,x)}-u(r,x)\right|^{2}(J_{n}*\rho)(r,x)
+|^p(r,x)|2ρ(r,x)]dxdr.\displaystyle+|\hat{\nabla}p(r,x)|^{2}\rho(r,x)\bigg{]}dxdr.

Let

εn=0T𝕋2|Jn(ρu)(r,x)(Jnρ)(r,x)u(r,x)|2(Jnρ)(r,x)𝑑x𝑑r.\varepsilon_{n}=\int_{0}^{T}\int_{\mathbb{T}^{2}}\left|\frac{J_{n}*(\rho u)(r,x)}{(J_{n}*\rho)(r,x)}-u(r,x)\right|^{2}(J_{n}*\rho)(r,x)dxdr.

By Gronwall inequality

teS(ρ^n(t))(t+2νQT(ρ))e2𝔸¯T(ρ)+12νεn.te^{S(\hat{\rho}_{n}(t))}\leq\left(t+\frac{2}{\nu}Q_{T}(\rho)\right)e^{2\overline{\mathbb{A}}_{T}(\rho)+\frac{1}{2\nu}\varepsilon_{n}}.

Note that

𝕋2|Jn(ρu)(r,x)(Jnρ)(r,x)u(r,x)|2(Jnρ)(r,x)𝑑x\displaystyle\int_{\mathbb{T}^{2}}\left|\frac{J_{n}*(\rho u)(r,x)}{(J_{n}*\rho)(r,x)}-u(r,x)\right|^{2}(J_{n}*\rho)(r,x)dx (4.18)
=𝕋2|Jn(ρu)(r,x)(Jnρ)(r,x)|2(Jnρ)(r,x)𝑑x+𝕋2|u(r,x)|2(Jnρ)(r,x)𝑑x\displaystyle=\int_{\mathbb{T}^{2}}\left|\frac{J_{n}*(\rho u)(r,x)}{(J_{n}*\rho)(r,x)}\right|^{2}(J_{n}*\rho)(r,x)dx+\int_{\mathbb{T}^{2}}\left|u(r,x)\right|^{2}(J_{n}*\rho)(r,x)dx
2𝕋2Jn(ρu)(r,x)u(r,x)𝑑x.\displaystyle-2\int_{\mathbb{T}^{2}}J_{n}*(\rho u)(r,x)\cdot u(r,x)dx.

From Lemma 8.1.10 of [1] combined with Lemma 3.4 and Jensen’s inequality [1], (4.18) can be controlled by ρ(r)22\|\rho(r)\|_{2}^{2}. In addition, for rr satisfying ρ(r)L2(𝕋2),\rho(r)\in L^{2}(\mathbb{T}^{2}), by Fatou’s Lemma and Lemma 8.1.10 of [1], the first term converge to 𝕋2|u(r,x)|2ρ(r,x)𝑑x\int_{\mathbb{T}^{2}}|u(r,x)|^{2}\rho(r,x)dx. By property of mollifiers, the other two terms also converge to 𝕋2|u(r,x)|2ρ(r,x)𝑑x\int_{\mathbb{T}^{2}}|u(r,x)|^{2}\rho(r,x)dx. Hence (4.18) converge to 0 for a.e rr. Taking nn\to\infty, by dominated convergence theorem we have εn0,\varepsilon_{n}\to 0, and by Fatou’s lemma we arrive at

S(ρ(t))2𝔸¯T(ρ)+log(1+2QT(ρ)νt).S(\rho(t))\leq 2\overline{\mathbb{A}}_{T}(\rho)+\log\left(1+\frac{2Q_{T}(\rho)}{\nu t}\right).

Turning to (4.13), by (4.5), for α>0,\alpha>0,

tαS(ρ^n(t))=0trα1+log(ρ^n(r)),tρ^n(r)𝑑r+0tαrα1S(ρ^n(r))𝑑r\displaystyle t^{\alpha}S(\hat{\rho}_{n}(t))=\int_{0}^{t}r^{\alpha}\left\langle 1+\log(\hat{\rho}_{n}(r)),\partial_{t}\hat{\rho}_{n}(r)\right\rangle dr+\int_{0}^{t}\alpha r^{\alpha-1}S(\hat{\rho}_{n}(r))dr
=0tαrα1S(ρ^n(r))ν0trαI(ρ^n(r))𝑑r\displaystyle=\int_{0}^{t}\alpha r^{\alpha-1}S(\hat{\rho}_{n}(r))-\nu\int_{0}^{t}r^{\alpha}I(\hat{\rho}_{n}(r))dr
+0t𝕋2rα(Jnρ)(r,x)(Jnρ)(r,x)(Jn(ρ^p))(r,x)𝑑x𝑑r\displaystyle+\int_{0}^{t}\int_{\mathbb{T}^{2}}r^{\alpha}\frac{(\nabla J_{n}*\rho)(r,x)}{(J_{n}*\rho)(r,x)}\cdot(J_{n}*(\rho\hat{\nabla}p))(r,x)dxdr
+0t𝕋2rα(Jnρ)(r,x)(Jnρ)(r,x)(Jn(ρu))(r,x)𝑑x𝑑r.\displaystyle+\int_{0}^{t}\int_{\mathbb{T}^{2}}r^{\alpha}\frac{(\nabla J_{n}*\rho)(r,x)}{(J_{n}*\rho)(r,x)}\cdot(J_{n}*(\rho u))(r,x)dxdr.

Similarly with (4.15), (4.16), we have

ν4I(ρ^n(r))+𝕋2(Jnρ)(r,x)(Jnρ)(r,x)(Jn(ρ^p))(r,x)𝑑x\displaystyle-\frac{\nu}{4}I(\hat{\rho}_{n}(r))+\int_{\mathbb{T}^{2}}\frac{(\nabla J_{n}*\rho)(r,x)}{(J_{n}*\rho)(r,x)}\cdot(J_{n}*(\rho\hat{\nabla}p))(r,x)dx
1ν𝕋2|^p(r,x)|2ρ(r,x)𝑑x;\displaystyle\leq\frac{1}{\nu}\int_{\mathbb{T}^{2}}|\hat{\nabla}p(r,x)|^{2}\rho(r,x)dx;
ν4I(ρ^n(r))+𝕋2(Jnρ)(r,x)(Jnρ)(r,x)(Jn(ρu))(r,x)𝑑x\displaystyle-\frac{\nu}{4}I(\hat{\rho}_{n}(r))+\int_{\mathbb{T}^{2}}\frac{(\nabla J_{n}*\rho)(r,x)}{(J_{n}*\rho)(r,x)}\cdot(J_{n}*(\rho u))(r,x)dx
1ν𝕋2|Jn(ρu)(r,x)(Jnρ)(r,x)u(r,x)|2(Jnρ)(r,x)𝑑x=εnν.\displaystyle\leq\frac{1}{\nu}\int_{\mathbb{T}^{2}}\left|\frac{J_{n}*(\rho u)(r,x)}{(J_{n}*\rho)(r,x)}-u(r,x)\right|^{2}(J_{n}*\rho)(r,x)dx=\frac{\varepsilon_{n}}{\nu}.

Hence,

0tαS(ρ^n(t))0tαrα1S(ρ^n(r))𝑑rν20trαI(ρ^n(r))𝑑r+tα(4𝔸¯T(ρ)+εnν).0\leq t^{\alpha}S(\hat{\rho}_{n}(t))\leq\int_{0}^{t}\alpha r^{\alpha-1}S(\hat{\rho}_{n}(r))dr-\frac{\nu}{2}\int_{0}^{t}r^{\alpha}I(\hat{\rho}_{n}(r))dr+t^{\alpha}\left(4\overline{\mathbb{A}}_{T}(\rho)+\frac{\varepsilon_{n}}{\nu}\right).

Then by Jensen’s inequality and Fatou’s Lemma,

00tαrα1S(ρ(r))𝑑rν20trαI(ρ(r))𝑑r+4tα𝔸¯T(ρ).0\leq\int_{0}^{t}\alpha r^{\alpha-1}S(\rho(r))dr-\frac{\nu}{2}\int_{0}^{t}r^{\alpha}I(\rho(r))dr+4t^{\alpha}\overline{\mathbb{A}}_{T}(\rho).

The result is now an immediate consequence of (4.12). ∎

4.3 Proof of Lemma 2.5

Proof.

Suppose ρC([0,T];𝒫(𝕋2))\rho\in C([0,T];\mathcal{P}(\mathbb{T}^{2})), QT(ρ)<Q_{T}(\rho)<\infty and 𝔸¯T(ρ)<\overline{\mathbb{A}}_{T}(\rho)<\infty. Take ^p\hat{\nabla}p in Lemma 4.1. Let u=𝒦ρu=\mathcal{K}*\rho and then in distribution sense,

tρ+div(ρu)νΔρ+div(ρ^p)=0.\partial_{t}\rho+{\rm div}\,(\rho u)-\nu\Delta\rho+{\rm div}\,(\rho\hat{\nabla}p)=0.

By Lemma 8.3.1 of [1], to show ρAC((0,T);𝒫(𝕋2))\rho\in AC((0,T);\mathcal{P}(\mathbb{T}^{2})), we just need to prove

0TνΔρ(t)div(ρ(t)^p(t))div(ρ(t)u(t))1,ρ(t)𝑑t<.\int_{0}^{T}\|\nu\Delta\rho(t)-{\rm div}\,(\rho(t)\hat{\nabla}p(t))-{\rm div}\,(\rho(t)u(t))\|_{-1,\rho(t)}dt<\infty. (4.19)

By Lemma A.1, Proposition A.2 and Lemma 3.4,

0TνΔρ(t)div(ρ(t)^p(t))div(ρ(t)u(t))1,ρ(t)𝑑t\displaystyle\int_{0}^{T}\|\nu\Delta\rho(t)-{\rm div}\,(\rho(t)\hat{\nabla}p(t))-{\rm div}\,(\rho(t)u(t))\|_{-1,\rho(t)}dt
0T(ρ(t)2+I12(ρ(t))+(𝕋2|^p(t)|2𝑑ρ(t))12)𝑑t.\displaystyle\leq\int_{0}^{T}\bigg{(}\|\rho(t)\|_{2}+I^{\frac{1}{2}}(\rho(t))+\left(\int_{\mathbb{T}^{2}}|\hat{\nabla}p(t)|^{2}d\rho(t)\right)^{\frac{1}{2}}\bigg{)}dt.

Then (4.19) holds by Holders inequality and Lemma 4.2.
To prove the second conclusion, note that if ρAC((0,T);𝒫(𝕋2))\rho\in AC((0,T);\mathcal{P}(\mathbb{T}^{2})), by Lemma A.1,

𝔸¯T(ρ)=14ν0T𝕋2|^p(t)|2𝑑ρ(t)𝑑t\displaystyle\overline{\mathbb{A}}_{T}(\rho)=\frac{1}{4\nu}\int_{0}^{T}\int_{\mathbb{T}^{2}}|\hat{\nabla}p(t)|^{2}d\rho(t)dt
14ν0Ttρ(t)νΔρ(t)+div(ρ(t)u(t))1,ρ(t)2𝑑t=𝔸T(ρ).\displaystyle\geq\frac{1}{4\nu}\int_{0}^{T}\|\partial_{t}\rho(t)-\nu\Delta\rho(t)+{\rm div}\,(\rho(t)u(t))\|^{2}_{-1,\rho(t)}dt=\mathbb{A}_{T}(\rho).

By Lemma D.34 of [20], for almost every t[0,T]t\in[0,T], there exists ~p(t)\tilde{\nabla}p(t) such that

tρ(t)νΔρ(t)+div(ρ(t)u(t)),φ=𝕋2φ~p(t)𝑑ρ(t),φC(𝕋2),\left\langle\partial_{t}\rho(t)-\nu\Delta\rho(t)+{\rm div}\,(\rho(t)u(t)),\varphi\right\rangle=\int_{\mathbb{T}^{2}}\nabla\varphi\cdot\tilde{\nabla}p(t)d\rho(t),\quad\forall\varphi\in C^{\infty}(\mathbb{T}^{2}),

and

div(ρ(t)~p(t))1,ρ(t)2=𝕋2|~p(t)|2𝑑ρ(t).\|-{\rm div}\,(\rho(t)\tilde{\nabla}p(t))\|^{2}_{-1,\rho(t)}=\int_{\mathbb{T}^{2}}|\tilde{\nabla}p(t)|^{2}d\rho(t).

By definition of 𝔸¯T(ρ)\overline{\mathbb{A}}_{T}(\rho) and Cauchy-Schwarz inequality,

𝔸¯T(ρ)supϕ[(0T𝕋2|~p(t)|2dρ(t)dt)12(0T𝕋2|ϕ(t)|2dρ(t)dt)12\displaystyle\overline{\mathbb{A}}_{T}(\rho)\leq\sup_{\phi}\bigg{[}\bigg{(}\int_{0}^{T}\int_{\mathbb{T}^{2}}|\tilde{\nabla}p(t)|^{2}d\rho(t)dt\bigg{)}^{\frac{1}{2}}\bigg{(}\int_{0}^{T}\int_{\mathbb{T}^{2}}|\nabla\phi(t)|^{2}d\rho(t)dt\bigg{)}^{\frac{1}{2}}
ν𝕋2|ϕ(t)|2dρ(t)dt]14ν0T𝕋2|~p(t)|2dρ(t)dt𝔸T(ρ).\displaystyle-\nu\int_{\mathbb{T}^{2}}|\nabla\phi(t)|^{2}d\rho(t)dt\bigg{]}\leq\frac{1}{4\nu}\int_{0}^{T}\int_{\mathbb{T}^{2}}|\tilde{\nabla}p(t)|^{2}d\rho(t)dt\leq\mathbb{A}_{T}(\rho).

5 Perturbed dynamics and nice trajectory approximation

In this section, we establish the law of large number (LLN) for the perturbed systems (2.25) and prove the ”nice” trajectory approximation. Section 5.1 investigates the uniqueness of perturbed mean-field equation (2.26). Section 5.2 provides a prior energy estimation similar to Lemma 3.7 which is crucial for proving LLN. Section 5.3 presents the proof of Lemmas 2.2, 2.7 and 2.8. Finally, we reach Lemma 2.9 in Section 5.4.

5.1 Uniqueness of perturbed mean-field equation

We prove the uniqueness of weak solution of (2.26), the mean-field equation for the stochastic interacting models perturbed by vv.

Lemma 5.1.

Given γ𝒫(𝕋2)\gamma\in\mathcal{P}(\mathbb{T}^{2}) with e(γ)<e(\gamma)<\infty. Suppose ρ\rho and ρ\rho^{\prime} are two weak solutions of (2.26) for vL([0,T]×𝕋2;2)v\in L^{\infty}([0,T]\times\mathbb{T}^{2};\mathbb{R}^{2}) and ρ(0)=ρ(0)=γ\rho(0)=\rho^{\prime}(0)=\gamma. Then ρ=ρ.\rho^{\prime}=\rho.

Proof.

Write u=𝒦ρ,u=\mathcal{K}*\rho, u=𝒦ρ.u^{\prime}=\mathcal{K}*\rho^{\prime}. Take smooth mollifiers JnJ_{n} defined in Lemma 4.1. Then by (C.5) and with the same argument as the steps 2&3 in the proof of Lemma 4.1, we obtain

12Jn(u(t)u(t))22=ν0tJn(ρ(s)ρ(s))22𝑑s\displaystyle\frac{1}{2}\|J_{n}*(u(t)-u^{\prime}(t))\|_{2}^{2}=-\nu\int_{0}^{t}\|J_{n}*(\rho(s)-\rho^{\prime}(s))\|^{2}_{2}ds (5.1)
0t(ρ(s)ρ(s))v(s),JnJn𝒩(ρ(s)ρ(s))𝑑s\displaystyle-\int_{0}^{t}\left\langle(\rho(s)-\rho^{\prime}(s))v(s),J_{n}*J_{n}*\mathcal{N}*(\rho(s)-\rho^{\prime}(s))\right\rangle ds
0tρ(s)u(s)ρ(s)u(s),JnJn𝒩(ρ(s)ρ(s))𝑑s\displaystyle-\int_{0}^{t}\left\langle\rho(s)u(s)-\rho^{\prime}(s)u^{\prime}(s),J_{n}*J_{n}*\nabla\mathcal{N}*(\rho(s)-\rho^{\prime}(s))\right\rangle ds

By Lemma 3.4 and Jensen’s inequality,

|ρ(t)u(t)ρ(t)u(t),JnJn𝒩(ρ(t)ρ(t))|\displaystyle\big{|}\left\langle\rho(t)u(t)-\rho^{\prime}(t)u^{\prime}(t),J_{n}*J_{n}*\nabla\mathcal{N}*(\rho(t)-\rho^{\prime}(t))\right\rangle\big{|}
4sup(ρ~,u~){ρ(t),ρ(t)}×{u(t),u(t)}𝕋2u~2𝑑ρ~C(ρ(t)122+ρ(t)122).\displaystyle\leq 4\sup_{(\tilde{\rho},\tilde{u})\in\{\rho(t),\rho^{\prime}(t)\}\times\{u(t),u^{\prime}(t)\}}\int_{\mathbb{T}^{2}}\tilde{u}^{2}d\tilde{\rho}\leq C\left(\|\rho(t)-1\|_{2}^{2}+\|\rho^{\prime}(t)-1\|_{2}^{2}\right).

Similarly,

|(ρ(t)ρ(t))v(t),JnJn𝒩(ρ(t)ρ(t))|\displaystyle\big{|}\left\langle(\rho(t)-\rho^{\prime}(t))v(t),J_{n}*J_{n}*\mathcal{N}*(\rho(t)-\rho^{\prime}(t))\right\rangle\big{|}
Cv(ρ(t)2+ρ(t)2)(u(t)2+u(t)2)).\displaystyle\leq C_{v}\left(\|\rho(t)\|_{2}+\|\rho^{\prime}(t)\|_{2}\right)\left(\|u(t)\|_{2}+\|u^{\prime}(t)\|_{2})\right).

Since QT(ρ)Q_{T}(\rho) and QT(ρ)Q_{T}(\rho^{\prime}) are finite, applying dominated convergence theorem to (5.1), we have

12u(t)u(t)22=ν0tρ(s)ρ(s)22𝑑s\displaystyle\frac{1}{2}\|u(t)-u^{\prime}(t)\|_{2}^{2}=-\nu\int_{0}^{t}\|\rho(s)-\rho^{\prime}(s)\|^{2}_{2}ds (5.2)
0t(ρ(s)ρ(s))v(s),𝒩(ρ(s)ρ(s))𝑑s\displaystyle-\int_{0}^{t}\left\langle(\rho(s)-\rho^{\prime}(s))v(s),\nabla\mathcal{N}*(\rho(s)-\rho^{\prime}(s))\right\rangle ds
0tρ(s)u(s)ρ(s)u(s),𝒩(ρ(s)ρ(s))𝑑s\displaystyle-\int_{0}^{t}\left\langle\rho(s)u(s)-\rho^{\prime}(s)u^{\prime}(s),\nabla\mathcal{N}*(\rho(s)-\rho^{\prime}(s))\right\rangle ds

Noticing that div(ρ(t)u(t))=curl(u(t)u(t)),{\rm div}\,(\rho(t)u(t))={\rm curl}\,(u(t)\cdot\nabla u(t)), we have

|0tρ(s)u(s)ρ(s)u(s),𝒩(ρρ)(s)𝑑s|\displaystyle\bigg{|}\int_{0}^{t}\left\langle\rho(s)u(s)-\rho^{\prime}(s)u^{\prime}(s),\nabla\mathcal{N}*(\rho-\rho^{\prime})(s)\right\rangle ds\bigg{|}
=|0t(uu)(s)(uu)(s),u(s)u(s)𝑑s|.\displaystyle=\bigg{|}\int_{0}^{t}\left\langle(u\cdot\nabla u)(s)-(u^{\prime}\cdot\nabla u^{\prime})(s),u(s)-u^{\prime}(s)\right\rangle ds\bigg{|}.

Since divu(t)=0,{\rm div}\,u^{\prime}(t)=0, writing w=u(t)u(t)w=u(t)-u^{\prime}(t),

u(t)(uu)(t),(uu)(t)=12|w|2,u(t)=0.\left\langle u^{\prime}(t)\cdot\nabla(u-u^{\prime})(t),(u-u^{\prime})(t)\right\rangle=\frac{1}{2}\left\langle\nabla|w|^{2},u^{\prime}(t)\right\rangle=0.

Hence,

|0t(uu)(s)(uu)(s),u(s)u(s)𝑑s|\displaystyle\bigg{|}\int_{0}^{t}\left\langle(u\cdot\nabla u)(s)-(u^{\prime}\cdot\nabla u^{\prime})(s),u(s)-u^{\prime}(s)\right\rangle ds\bigg{|} (5.3)
=|0t(uu)(s)u(s),u(s)u(s)𝑑s|\displaystyle=\bigg{|}\int_{0}^{t}\left\langle(u-u^{\prime})(s)\cdot\nabla u(s),u(s)-u^{\prime}(s)\right\rangle ds\bigg{|}
0tu(s)u(s)42u(s)2𝑑s.\displaystyle\leq\int_{0}^{t}\|u(s)-u^{\prime}(s)\|_{4}^{2}\|\nabla u(s)\|_{2}ds.

Apply Gagliardo–Nirenberg interpolation inequality to (5.3),

|0tρ(s)u(s)ρ(s)u(s),𝒩(ρρ)(s)𝑑s|\displaystyle\bigg{|}\int_{0}^{t}\left\langle\rho(s)u(s)-\rho^{\prime}(s)u^{\prime}(s),\nabla\mathcal{N}*(\rho-\rho^{\prime})(s)\right\rangle ds\bigg{|} (5.4)
0tu(s)u(s)2u(s)u(s)2u(s)2𝑑s,\displaystyle\leq\int_{0}^{t}\|u(s)-u^{\prime}(s)\|_{2}\|\nabla u(s)-\nabla u^{\prime}(s)\|_{2}\|\nabla u(s)\|_{2}ds,
40tu(s)u(s)2ρ(s)ρ(s)2ρ(s)12𝑑s,\displaystyle\leq 4\int_{0}^{t}\|u(s)-u^{\prime}(s)\|_{2}\|\rho(s)-\rho^{\prime}(s)\|_{2}\|\rho(s)-1\|_{2}ds,

where we used the fact

u(t)222i,j1,2ij(𝒩ρ)222Δ(𝒩ρ)22=2ρ122.\|\nabla u(t)\|^{2}_{2}\leq 2\sum_{i,j\in{1,2}}\|\partial_{ij}(-\mathcal{N}*\rho)\|_{2}^{2}\leq 2\|\Delta(-\mathcal{N}*\rho)\|_{2}^{2}=2\|\rho-1\|_{2}^{2}.

On the other hand, by Cauchy–Schwarz inequality and (C.5),

|0t(ρ(s)ρ(s)v(s)),𝒩(ρ(s)ρ(s))𝑑s|\displaystyle\bigg{|}\int_{0}^{t}\left\langle(\rho(s)-\rho^{\prime}(s)v(s)),\nabla\mathcal{N}*(\rho(s)-\rho^{\prime}(s))\right\rangle ds\bigg{|} (5.5)
Cv0tu(s)u(s)2ρ(s)ρ(s)2𝑑s,\displaystyle\leq C_{v}\int_{0}^{t}\|u(s)-u^{\prime}(s)\|_{2}\|\rho(s)-\rho^{\prime}(s)\|_{2}ds,

Combining (5.2) with (5.4) and (5.5), we have

12u(t)u(t)22ν0tρ(s)ρ(s)22𝑑s\displaystyle\frac{1}{2}\|u(t)-u^{\prime}(t)\|_{2}^{2}\leq-\nu\int_{0}^{t}\|\rho(s)-\rho^{\prime}(s)\|^{2}_{2}ds
+C0tu(s)u(s)2ρ(s)ρ(s)2(ρ(s)12+1)𝑑s\displaystyle+C\int_{0}^{t}\|u(s)-u^{\prime}(s)\|_{2}\|\rho(s)-\rho^{\prime}(s)\|_{2}(\|\rho(s)-1\|_{2}+1)ds
0tC24νu(s)u(s)22(ρ(s)12+1)2𝑑s,\displaystyle\leq\int_{0}^{t}\frac{C^{2}}{4\nu}\|u(s)-u^{\prime}(s)\|_{2}^{2}(\|\rho(s)-1\|_{2}+1)^{2}ds,

and the uniqueness follows from Gronwall’s inequality. ∎

5.2 Prior energy estimation of perturbed dynamics

To prove the law of large number for perturbed dynamics, we also need a prior energy estimation, whose proof is similar to that of Lemma 3.7.

Lemma 5.2.

There exists constants λ>0\lambda>0, and CvC_{v} only dependent on v\|v\|_{\infty} such that for each sequence (ηn)(\eta_{n}) satisfying ηn𝒳n\eta_{n}\in\mathcal{X}_{n} and e(ζnηn)R,e(\zeta_{n}*\eta_{n})\leq R,

lim supn\displaystyle\limsup_{n\to\infty} 1nlogPηnv{sup0tT(e(ζnρn(t))+ν20tζnρn(s)122𝑑s)>l}\displaystyle\frac{1}{n}\log P^{v}_{\eta_{n}}\bigg{\{}\sup_{0\leq t\leq T}\bigg{(}e(\zeta_{n}*\rho_{n}(t))+\frac{\nu}{2}\int_{0}^{t}\|\zeta_{n}*\rho_{n}(s)-1\|_{2}^{2}ds\bigg{)}>l\bigg{\}} (5.6)
λ(lRCvT).\displaystyle\leq-\lambda(l-R-C_{v}T).
Proof.

Recall 𝒩n=Gn𝒩,ωn=νn(Gn(0)1).\mathcal{N}_{n}=G_{n}*\mathcal{N},\omega_{n}=\frac{\nu}{n}(G_{n}(0)-1). By Ito’s formula, as calculation in Lemma 3.7, for any λ>0,\lambda>0,

exp{nλ[e(ζnρn(t))e(ζnρn(0))+ν0tζnρn(s)122ds\displaystyle\exp\bigg{\{}n\lambda\bigg{[}e(\zeta_{n}*\rho_{n}(t))-e(\zeta_{n}*\rho_{n}(0))+\nu\int_{0}^{t}\|\zeta_{n}*\rho_{n}(s)-1\|_{2}^{2}ds (5.7)
0t𝒩nρn(s),𝐑(ρn(s))𝑑sλν0t|𝒩nρn(s)|2,ρn(s)𝑑sωnt\displaystyle-\int_{0}^{t}\left\langle\nabla\mathcal{N}_{n}*\rho_{n}(s),\mathbf{R}(\rho_{n}(s))\right\rangle ds-\lambda\nu\int_{0}^{t}\left\langle|\nabla\mathcal{N}_{n}*\rho_{n}(s)|^{2},\rho_{n}(s)\right\rangle ds-\omega_{n}t
0t𝒩nρn(s)v(s),ρn(s)ds]}\displaystyle-\int_{0}^{t}\left\langle\nabla\mathcal{N}_{n}*\rho_{n}(s)\cdot v(s),\rho_{n}(s)\right\rangle ds\bigg{]}\bigg{\}}

is a positive continuous martingale. By Lemma 3.6, for each ηn𝒳n\eta_{n}\in\mathcal{X}_{n} with e(ζnηn)R,e(\zeta_{n}*\eta_{n})\leq R,

Pηnv{sup0<tT[e(ζnρn(t))0t𝒩nρn(s),𝐑(ρn(s))ds\displaystyle P^{v}_{\eta_{n}}\bigg{\{}\sup_{0<t\leq T}\bigg{[}e(\zeta_{n}*\rho_{n}(t))-\int_{0}^{t}\left\langle\nabla\mathcal{N}_{n}*\rho_{n}(s),\mathbf{R}(\rho_{n}(s))\right\rangle ds
+ν0tζnρn(s)122𝑑sνλ0t𝕋2|𝒩nρn(s)|2ρn(s,dx)𝑑s\displaystyle+\nu\int_{0}^{t}\|\zeta_{n}*\rho_{n}(s)-1\|_{2}^{2}ds-\nu\lambda\int_{0}^{t}\int_{\mathbb{T}^{2}}|\nabla\mathcal{N}_{n}*\rho_{n}(s)|^{2}\rho_{n}(s,dx)ds
0t𝒩nρn(s)v(s),ρn(s)ds]>l}enλ(lRωnT).\displaystyle-\int_{0}^{t}\left\langle\nabla\mathcal{N}_{n}*\rho_{n}(s)\cdot v(s),\rho_{n}(s)\right\rangle ds\bigg{]}>l\bigg{\}}\leq e^{-n\lambda(l-R-\omega_{n}T)}.

By Corollary 3.5, there exists Cv>0C_{v}>0 dependent on ν\nu and v\|v\|_{\infty}, such that

𝒩nρn(t)v(t),ρn(t)Cv+ν6ζnρn(t)122.\left\langle\nabla\mathcal{N}_{n}*\rho_{n}(t)\cdot v(t),\rho_{n}(t)\right\rangle\leq C_{v}+\frac{\nu}{6}\|\zeta_{n}*\rho_{n}(t)-1\|_{2}^{2}.

Combining this inequality with (3.16) and taking λ\lambda such that C2λ=16C_{2}\lambda=\frac{1}{6}, we have

Pηnv{sup0<tT[e(ζnρn(t))0t𝒩nρn(s),𝐑(ρn(s))ds\displaystyle P^{v}_{\eta_{n}}\bigg{\{}\sup_{0<t\leq T}\bigg{[}e(\zeta_{n}*\rho_{n}(t))-\int_{0}^{t}\left\langle\nabla\mathcal{N}_{n}*\rho_{n}(s),\mathbf{R}(\rho_{n}(s))\right\rangle ds
+2ν30tζnρn(s)122ds]>l}enλ(lRωnTCvT).\displaystyle+\frac{2\nu}{3}\int_{0}^{t}\|\zeta_{n}*\rho_{n}(s)-1\|_{2}^{2}ds\bigg{]}>l\bigg{\}}\leq e^{-n\lambda(l-R-\omega_{n}T-C_{v}T)}.

The rest is similar to the proof of Lemma 3.7. ∎

Then we verify that there exists a sequence of (γn)n1(\gamma_{n})_{n\geq 1} satisfying the conditions of Lemma 5.2.

Lemma 5.3.

For γ𝒫(𝕋2)\gamma\in\mathcal{P}(\mathbb{T}^{2}) with e(γ)<e(\gamma)<\infty, there exists γn𝒳n\gamma_{n}\in\mathcal{X}_{n} such that

limne(ζnγn)=e(γ),limnd(γn,γ)=0.\lim_{n\to\infty}e(\zeta_{n}*\gamma_{n})=e(\gamma),\quad\lim_{n\to\infty}d(\gamma_{n},\gamma)=0.
Proof.

Take a sequence of i.i.d. random variables (Yi)i1(Y_{i})_{i\geq 1} with the law γ\gamma.
Let ηn:=1ni=1nδYi\eta_{n}:=\frac{1}{n}\sum_{i=1}^{n}\delta_{Y_{i}}. Write 𝒩n=Gn𝒩.\mathcal{N}_{n}=G_{n}*\mathcal{N}. By (C.5) we have

𝔼[ζn𝒦ηn𝒦γ22]=𝔼[1n21i,jn𝒩n(YiYj)]+𝒦γ22\displaystyle\mathbb{E}\left[\|\zeta_{n}*\mathcal{K}*\eta_{n}-\mathcal{K}*\gamma\|_{2}^{2}\right]=\mathbb{E}\left[\frac{1}{n^{2}}\sum_{1\leq i,j\leq n}\mathcal{N}_{n}(Y_{i}-Y_{j})\right]+\|\mathcal{K}*\gamma\|_{2}^{2} (5.8)
2𝔼[ζn𝒩ηn,γ]\displaystyle-2\mathbb{E}\left[\left\langle\zeta_{n}*\mathcal{N}*\eta_{n},\gamma\right\rangle\right]
=1n2ij𝔼𝒩n(YiYj)+1n𝒩n(0)+𝒦γ22\displaystyle=\frac{1}{n^{2}}\sum_{i\neq j}\mathbb{E}\mathcal{N}_{n}(Y_{i}-Y_{j})+\frac{1}{n}\mathcal{N}_{n}(0)+\|\mathcal{K}*\gamma\|_{2}^{2}
i=1n2n𝕋2𝔼(ζn𝒩)(xYi)γ(dx)\displaystyle-\sum_{i=1}^{n}\frac{2}{n}\int_{\mathbb{T}^{2}}\mathbb{E}(\zeta_{n}*\mathcal{N})(x-Y_{i})\gamma(dx)
=n1n𝒦ζnγ22+𝒦γ22+1n𝒩n(0)\displaystyle=\frac{n-1}{n}\|\mathcal{K}*\zeta_{n}*\gamma\|_{2}^{2}+\|\mathcal{K}*\gamma\|_{2}^{2}+\frac{1}{n}\mathcal{N}_{n}(0)
2(ζn𝒩)γ,γ.\displaystyle-2\left\langle(\zeta_{n}*\mathcal{N})*\gamma,\gamma\right\rangle.

Note that limn|1n𝒩n(0)|=limnmn2G𝒩1n=0.\lim_{n\to\infty}\left|\frac{1}{n}\mathcal{N}_{n}(0)\right|=\lim_{n\to\infty}\frac{m_{n}^{2}\|G\|_{\infty}\|\mathcal{N}\|_{1}}{n}=0. By Jensen’s inequality and Fatou’s Lemma, along with (C.6), we have

lim supn𝔼[𝒦(ζnηnγ)22]2𝒦γ222𝒩γ,γ=0.\limsup_{n\to\infty}\mathbb{E}\left[\|\mathcal{K}*(\zeta_{n}*\eta_{n}-\gamma)\|_{2}^{2}\right]\leq 2\|\mathcal{K}*\gamma\|_{2}^{2}-2\left\langle\mathcal{N}*\gamma,\gamma\right\rangle=0. (5.9)

Pick γn𝒳n\gamma_{n}\in\mathcal{X}_{n}, such that 0ζn𝒦γn𝒦γ22𝔼[ζn𝒦ηn𝒦γ22]0\leq\|\zeta_{n}*\mathcal{K}*\gamma_{n}-\mathcal{K}*\gamma\|_{2}^{2}\leq\mathbb{E}\left[\|\zeta_{n}*\mathcal{K}*\eta_{n}-\mathcal{K}*\gamma\|_{2}^{2}\right]. In view of (5.9) and (C.6), limne(γn)=e(γ),\lim_{n\to\infty}e(\gamma_{n})=e(\gamma), and 𝒩γn𝒩γ\nabla\mathcal{N}*\gamma_{n}\to\nabla\mathcal{N}*\gamma in L2(𝕋2)L^{2}(\mathbb{T}^{2}). Hence for each φC(𝕋2)\varphi\in C^{\infty}(\mathbb{T}^{2}),

γnγ,ϕ=γnγ,Δ𝒩ϕ=𝒩γn𝒩γ,ϕ0,\left\langle\gamma_{n}-\gamma,\phi\right\rangle=\left\langle\gamma_{n}-\gamma,-\Delta\mathcal{N}*\phi\right\rangle=\left\langle\nabla\mathcal{N}*\gamma_{n}-\nabla\mathcal{N}*\gamma,\nabla\phi\right\rangle\to 0,

implying limnd(γn,γ)=0.\lim_{n\to\infty}d(\gamma_{n},\gamma)=0.

5.3 Proof of Lemmas 2.2, 2.7 and 2.8

Before proving the law of large number, we give the proof for the fact that 𝐑\mathbf{R} is continuous with finite energy, which is Lemma 2.2.

Proof of Lemma 2.2.

By Jensen’s inequality, we only need to consider the case that supn1e(ζnγ)<.\sup_{n\geq 1}e(\zeta_{n}*\gamma)<\infty.

First we consider the case that φn=φ.\varphi_{n}=\varphi. Recall w(x,y)=r(x,y)𝒦(xy)w(x,y)=r(x,y)\mathcal{K}(x-y) and write

f(x,y):={12φ(x)φ(y)r(x,y)w(xy),if xy;0,if x=y.f(x,y):=\left\{\begin{aligned} &\frac{1}{2}\frac{\varphi(x)-\varphi(y)}{r(x,y)}w(x-y),&\text{if }x\neq y;\\ &0,&\text{if }x=y.\end{aligned}\right.

In view of (2.4), we just need to show

(𝕋2)2f(x,y)γn(dx)γn(dy)(𝕋2)2f(x,y)γ(dx)γ(dy).\int_{(\mathbb{T}^{2})^{2}}f(x,y)\gamma_{n}(dx)\gamma_{n}(dy)\to\int_{(\mathbb{T}^{2})^{2}}f(x,y)\gamma(dx)\gamma(dy).

By Tietze extension theorem, we can construct a continuous function fδf_{\delta} such that fδ(x,y)=f(x,y),f_{\delta}(x,y)=f(x,y), for r(x,y)δr(x,y)\geq\delta and sup|fδ|sup|f|.\sup|f_{\delta}|\leq\sup|f|.
By Lemma C.3 for δ<12\delta<\frac{1}{2}, there exists nδn_{\delta} such that for nnδn\geq n_{\delta},

|(𝕋2)2(ffδ)(x,y)γn(dx)γn(dy)|2sup|f|r(x,y)<δγn(dx)γn(dy)\displaystyle\bigg{|}\int_{(\mathbb{T}^{2})^{2}}(f-f_{\delta})(x,y)\gamma_{n}(dx)\gamma_{n}(dy)\bigg{|}\leq 2\sup|f|\int_{r(x,y)<\delta}\gamma_{n}(dx)\gamma_{n}(dy)
4sup|f|4πe(ζnγn)+C𝒩logδ.\displaystyle\leq 4\sup|f|\frac{4\pi e(\zeta_{n}*\gamma_{n})+C_{\mathcal{N}}}{-\log\delta}.

Similarly,

|(𝕋2)2(ffδ)(x,y)γ(dx)γ(dy)|2sup|f|r(x,y)<δγ(dx)γ(dy)\displaystyle\bigg{|}\int_{(\mathbb{T}^{2})^{2}}(f-f_{\delta})(x,y)\gamma(dx)\gamma(dy)\bigg{|}\leq 2\sup|f|\int_{r(x,y)<\delta}\gamma(dx)\gamma(dy)
2sup|f|4πe(γ)+C𝒩logδ.\displaystyle\leq 2\sup|f|\frac{4\pi e(\gamma)+C_{\mathcal{N}}}{-\log\delta}.

By lower semi-continuity of ee in 𝒫(𝕋2)\mathcal{P}(\mathbb{T}^{2}), e(γ)supne(ζnγn)<e(\gamma)\leq\sup_{n}e(\zeta_{n}*\gamma_{n})<\infty. Hence we can take δ\delta small enough such that

lim supn|(𝕋2)2(ffδ)(x,y)γn(dx)γn(dy)|<ε3,\displaystyle\limsup_{n\to\infty}\bigg{|}\int_{(\mathbb{T}^{2})^{2}}(f-f_{\delta})(x,y)\gamma_{n}(dx)\gamma_{n}(dy)\bigg{|}<\frac{\varepsilon}{3}, (5.10)
|(𝕋2)2(ffδ)(x,y)γ(dx)γ(dy)|<ε3.\displaystyle\bigg{|}\int_{(\mathbb{T}^{2})^{2}}(f-f_{\delta})(x,y)\gamma(dx)\gamma(dy)\bigg{|}<\frac{\varepsilon}{3}.

Since fδ(x,y)f_{\delta}(x,y) is a continuous function, and γnγnγγ\gamma_{n}\otimes\gamma_{n}\to\gamma\otimes\gamma in narrow topology,

limn(𝕋2)2fδ(x,y)γn(dx)γn(dy)=(𝕋2)2fδ(x,y)γ(dx)γ(dy).{}\lim_{n\to\infty}\int_{(\mathbb{T}^{2})^{2}}f_{\delta}(x,y)\gamma_{n}(dx)\gamma_{n}(dy)=\int_{(\mathbb{T}^{2})^{2}}f_{\delta}(x,y)\gamma(dx)\gamma(dy). (5.11)

Combining (5.10) and (5.11), we have

lim supn|(𝕋2)2f(x,y)γn(dx)γn(dy)(𝕋2)2f(x,y)γ(dx)γ(dy)|<23ε.\limsup_{n\to\infty}\left|\int_{(\mathbb{T}^{2})^{2}}f(x,y)\gamma_{n}(dx)\gamma_{n}(dy)-\int_{(\mathbb{T}^{2})^{2}}f(x,y)\gamma(dx)\gamma(dy)\right|<\frac{2}{3}\varepsilon.

Let ε0\varepsilon\to 0 and we conclude the proof for the case φn=φ\varphi_{n}=\varphi.

For the general case, let

fn(x,y):={12φn(x)φn(y)r(x,y)w(xy),if xy;0,if x=y.f_{n}(x,y):=\left\{\begin{aligned} &\frac{1}{2}\frac{\varphi_{n}(x)-\varphi_{n}(y)}{r(x,y)}w(x-y),&\text{if }x\neq y;\\ &0,&\text{if }x=y.\end{aligned}\right.

Note that

|φn,𝐑(γn)φ,𝐑(γn)|supr(x,y)δ|f(x,y)fn(x,y)|\displaystyle|\left\langle\varphi_{n},\mathbf{R}(\gamma_{n})\right\rangle-\left\langle\varphi,\mathbf{R}(\gamma_{n})\right\rangle|\leq\sup_{r(x,y)\geq\delta}|f(x,y)-f_{n}(x,y)|
+supr(x,y)<δ|f(x,y)+fn(x,y)|r(x,y)<δγn(dx)γn(dy).\displaystyle+\sup_{r(x,y)<\delta}|f(x,y)+f_{n}(x,y)|\int_{r(x,y)<\delta}\gamma_{n}(dx)\gamma_{n}(dy).

Take nn\to\infty, then we have

lim supn|φn,𝐑(γn)φ,𝐑(γn)|\displaystyle\limsup_{n\to\infty}|\left\langle\varphi_{n},\mathbf{R}(\gamma_{n})\right\rangle-\left\langle\varphi,\mathbf{R}(\gamma_{n})\right\rangle|
12(φ+φn)supn1r(x,y)<δγn(dx)γn(dy).\displaystyle\leq\frac{1}{2}\left(\|\nabla\varphi\|_{\infty}+\|\nabla\varphi_{n}\|_{\infty}\right)\sup_{n\geq 1}\int_{r(x,y)<\delta}\gamma_{n}(dx)\gamma_{n}(dy).

By Lemma C.3, take δ0\delta\downarrow 0 and we complete the proof. ∎

Proof of Lemma 2.7.

Given γn𝒳n\gamma_{n}\in\mathcal{X}_{n} and e(ζnγn)Re(\zeta_{n}*\gamma_{n})\leq R and limnd(γn,γ)=0.\lim_{n\to\infty}d(\gamma_{n},\gamma)=0. By Ito’s formula, for ϕC(𝕋2)\phi\in C^{\infty}(\mathbb{T}^{2}),

dϕ,ρn(t)=νΔϕ,ρn(t)dt+ϕ,𝐑(ρn(t))dt+v(t)ϕ,ρn(t)dt\displaystyle d\left\langle\phi,\rho_{n}(t)\right\rangle=\left\langle\nu\Delta\phi,\rho_{n}(t)\right\rangle dt+\left\langle\nabla\phi,\mathbf{R}(\rho_{n}(t))\right\rangle dt+\left\langle v(t)\cdot\nabla\phi,\rho_{n}(t)\right\rangle dt (5.12)
+2νni=1nϕ(Xi)dWi(t).\displaystyle+\frac{\sqrt{2\nu}}{n}\sum_{i=1}^{n}\nabla\phi(X_{i})dW_{i}(t).

Note that by (2.5)

𝔼γnv[|ϕ,ρn(t+h)ρn(t)|]C(2ϕ+vϕ)h.\mathbb{E}^{v}_{\gamma_{n}}[|\left\langle\phi,\rho_{n}(t+h)-\rho_{n}(t)\right\rangle|]\leq C\left(\|\nabla^{2}\phi\|_{\infty}+\|v\|_{\infty}\|\nabla\phi\|_{\infty}\right)h. (5.13)

Then by [29] and Theorems 8.6 and 8.8 in Chapter 3 of [18], we conclude that ρn\rho_{n} is tight in C([0,T];𝒫(𝕋2))C([0,T];\mathcal{P}(\mathbb{T}^{2})).

By Pohorov theorem, the family of probability laws is relatively compact in the topology of weak convergence of probability measures. We select a convergent subsequence indexed by nkn_{k}. By the Skorohod representation theorem, we can construct a canonical probability space (write P^\hat{P} for the corresponding probability measures) on which random variables ρ,ρ^n\rho,\hat{\rho}_{n} are defined, with the property that ρ^k\hat{\rho}_{k} has the same law as ρnk\rho_{n_{k}} for k=1,2,k=1,2,\cdots such that

ρ^kρa.s.P^ in C([0,T];𝒫(𝕋2)) as k.\hat{\rho}_{k}\to\rho\quad a.s.~{}\hat{P}\text{ in }C([0,T];\mathcal{P}(\mathbb{T}^{2}))\text{ as }k\to\infty. (5.14)

By Lemma 5.2 and Borel-Cantelli Lemma,

lim supkQT(ζnkρ^k)R+CvTa.s.P^.\limsup_{k\to\infty}Q_{T}(\zeta_{n_{k}}*\hat{\rho}_{k})\leq R+C_{v}T\quad a.s.~{}\hat{P}.

Then by the lower semi-continuities of QTQ_{T} under the topology of C([0,T];𝒫(𝕋2))C([0,T];\mathcal{P}(\mathbb{T}^{2})),

QT(ρ)R+CvT,a.s.P^.Q_{T}(\rho)\leq R+C_{v}T,\quad a.s.~{}\hat{P}. (5.15)

By Lemma 2.2,

limkϕ(t),𝐑(ρ^k(t))=ϕ(t),𝐑(ρ(t))a.s.P^.\lim_{k\to\infty}\left\langle\nabla\phi(t),\mathbf{R}(\hat{\rho}_{k}(t))\right\rangle=\left\langle\nabla\phi(t),\mathbf{R}(\rho(t))\right\rangle\quad a.s.\;\hat{P}.

By (5.12) and dominated convergence theorem, for each ϕC([0,T]×𝕋2),\phi\in C^{\infty}([0,T]\times\mathbb{T}^{2}),

limk[ϕ(t),ρ^k(t)ϕ(s),ρ^k(s)stϕ(r),𝐑(ρ^k(r))dr\displaystyle\lim_{k\to\infty}\bigg{[}\left\langle\phi(t),\hat{\rho}_{k}(t)\right\rangle-\left\langle\phi(s),\hat{\rho}_{k}(s)\right\rangle-\int_{s}^{t}\left\langle\nabla\phi(r),\mathbf{R}(\hat{\rho}_{k}(r))\right\rangle dr (5.16)
sttϕ(r)+ϕ(r)v(r)+νΔϕ(r),ρ^k(r)dr]\displaystyle-\int_{s}^{t}\left\langle\partial_{t}\phi(r)+\nabla\phi(r)\cdot v(r)+\nu\Delta\phi(r),\hat{\rho}_{k}(r)\right\rangle dr\bigg{]}
=ϕ(t),ρ(t)ϕ(s),ρ(s)sttϕ(r)+ϕ(r)v(r)+νΔϕ(r),ρ(r)𝑑r\displaystyle=\left\langle\phi(t),\rho(t)\right\rangle-\left\langle\phi(s),\rho(s)\right\rangle-\int_{s}^{t}\left\langle\partial_{t}\phi(r)+\nabla\phi(r)\cdot v(r)+\nu\Delta\phi(r),\rho(r)\right\rangle dr
stϕ(r),𝐑(ρ(r))𝑑ra.s.P^,\displaystyle-\int_{s}^{t}\left\langle\nabla\phi(r),\mathbf{R}(\rho(r))\right\rangle dr\quad a.s.\;\hat{P},

and

𝔼^[ϕ(t),ρ(t)ϕ(s),ρ(s)sttϕ(r)+ϕ(r)v(r)+νΔϕ(r),ρ(r)dr\displaystyle\hat{\mathbb{E}}\bigg{[}\left\langle\phi(t),\rho(t)\right\rangle-\left\langle\phi(s),\rho(s)\right\rangle-\int_{s}^{t}\left\langle\partial_{t}\phi(r)+\nabla\phi(r)\cdot v(r)+\nu\Delta\phi(r),\rho(r)\right\rangle dr (5.17)
stϕ(r),𝐑(ρ(r))dr]=0.\displaystyle-\int_{s}^{t}\left\langle\nabla\phi(r),\mathbf{R}(\rho(r))\right\rangle dr\bigg{]}=0.

By Fatou’s Lemma and (5.12),

Var[ϕ(t),ρ(t)ϕ(s),ρ(s)sttϕ(r)+ϕ(r)v(r)+νΔϕ(r),ρ(r)dr\displaystyle\text{Var}\bigg{[}\left\langle\phi(t),\rho(t)\right\rangle-\left\langle\phi(s),\rho(s)\right\rangle-\int_{s}^{t}\left\langle\partial_{t}\phi(r)+\nabla\phi(r)\cdot v(r)+\nu\Delta\phi(r),\rho(r)\right\rangle dr (5.18)
stϕ(r),𝐑(ρ(r))dr]\displaystyle-\int_{s}^{t}\left\langle\nabla\phi(r),\mathbf{R}(\rho(r))\right\rangle dr\bigg{]}
lim infkVar[sttϕ(r)+ϕ(r)v(r)+νΔϕ(r),ρ^k(r)dr\displaystyle\leq\liminf_{k\to\infty}\text{Var}\bigg{[}-\int_{s}^{t}\left\langle\partial_{t}\phi(r)+\nabla\phi(r)\cdot v(r)+\nu\Delta\phi(r),\hat{\rho}_{k}(r)\right\rangle dr
+ϕ(t),ρ^k(t)ϕ(s),ρ^k(s)stϕ(r),𝐑(ρ^k(r))dr]\displaystyle+\left\langle\phi(t),\hat{\rho}_{k}(t)\right\rangle-\left\langle\phi(s),\hat{\rho}_{k}(s)\right\rangle-\int_{s}^{t}\left\langle\nabla\phi(r),\mathbf{R}(\hat{\rho}_{k}(r))\right\rangle dr\bigg{]}
limk2νnkϕ2=0.\displaystyle\leq\lim_{k\to\infty}\frac{2\nu}{n_{k}}\|\nabla\phi\|_{\infty}^{2}=0.

Since limnd(ρn(0),γ)=limnd(γn,γ)=0\lim_{n\to\infty}d(\rho_{n}(0),\gamma)=\lim_{n\to\infty}d(\gamma_{n},\gamma)=0, we have ρ(0)=γ\rho(0)=\gamma. Then by (5.15), (5.17), (5.16) and (5.18), ρ\rho is almost surely a weak solution of (2.26) for vv in Definition 2.1.

By Lemma 5.3, the sequence γn\gamma_{n} does exists for R>e(γ)R>e(\gamma). According to the uniqueness, i.e. Lemma 5.1, ρ\rho is not stochastic and we have determined an unique weak solution ργ,v\rho^{\gamma,v}. Finally we conclude the proof by the relatively compactness of law of ρn\rho_{n}. ∎

Proof of Lemma 2.8.

We start with proving that there exists δ(0,ε)\delta\in(0,\varepsilon) such that for any γnBδ(γ)𝒳n\gamma_{n}\in B_{\delta}(\gamma)\cap\mathcal{X}_{n} and e(ζnγn)Re(\zeta_{n}*\gamma_{n})\leq R,

limnPγnv(sup0tTd(ρn(t),ργ,v(t))>ε)=0.\lim_{n\to\infty}P^{v}_{\gamma_{n}}\left(\sup_{0\leq t\leq T}d(\rho_{n}(t),\rho^{\gamma,v}(t))>\varepsilon\right)=0. (5.19)

To prove it by contradiction, suppose that it’s not true. Then, there exists ε>0,R>0,\varepsilon>0,R>0, so that for any δ\delta we can find γnBδ(γ)𝒳n\gamma_{n}\in B_{\delta}(\gamma)\cap\mathcal{X}_{n} and e(ζnγn)Re(\zeta_{n}*\gamma_{n})\leq R such that

lim supnPγnv(sup0tTd(ρn(t),ργ,v(t))>ε)>0.\limsup_{n\to\infty}P^{v}_{\gamma_{n}}\left(\sup_{0\leq t\leq T}d(\rho_{n}(t),\rho^{\gamma,v}(t))>\varepsilon\right)>0.

Taking subsequence if necessary, assume limnd(γn,γ)=0\lim_{n\to\infty}d(\gamma_{n},\gamma^{\prime})=0, where γ=γ(δ)\gamma^{\prime}=\gamma^{\prime}(\delta) depend on δ\delta. Then by Lemma 2.7,

limnPγnv(sup0tTd(ρn(t),ργ,v(t))>ε2)=0.\lim_{n\to\infty}P^{v}_{\gamma_{n}}\left(\sup_{0\leq t\leq T}d(\rho_{n}(t),\rho^{\gamma^{\prime},v}(t))>\frac{\varepsilon}{2}\right)=0.

Hence we obtain

sup0tTd(ργ,v(t),ργ,v(t))ε2.\sup_{0\leq t\leq T}d\left(\rho^{\gamma,v}(t),\rho^{\gamma^{\prime},v}(t)\right)\geq\frac{\varepsilon}{2}. (5.20)

For δ=1n\delta=\frac{1}{n}, we write ηn=γ(1n)B¯1n(γ)\eta_{n}=\gamma^{\prime}(\frac{1}{n})\in\overline{B}_{\frac{1}{n}}(\gamma) satisfying (5.20). Since we have the estimation

QT(ρηn,v)R+CvT,Q_{T}(\rho^{\eta_{n},v})\leq R+C_{v}T,

with a similar argument as the proof of Lemma 2.7,

ρηn,vρ in C([0,T];𝒫(𝕋2)),\rho^{\eta_{n},v}\to\rho\text{ in }C([0,T];\mathcal{P}(\mathbb{T}^{2})),

and ρ\rho is a weak solution of (2.26). By Lemma 5.1, ρ=ργ,v\rho=\rho^{\gamma,v}. However, it contradicts (5.20), so we have proved (5.19).

Now turn to (2.29). By Lemma 1.14 of [52],

|𝔼γnv[0T𝕋2|v(t)|2𝑑ρn(t)𝑑t]0T𝕋2|v(t)|2𝑑ργ,v(t)𝑑t|εTsup0tTv2(t)Lip\displaystyle\left|\mathbb{E}^{v}_{\gamma_{n}}\left[\int_{0}^{T}\int_{\mathbb{T}^{2}}|v(t)|^{2}d\rho_{n}(t)dt\right]-\int_{0}^{T}\int_{\mathbb{T}^{2}}|v(t)|^{2}d\rho^{\gamma,v}(t)dt\right|\leq\varepsilon T\sup_{0\leq t\leq T}\|v^{2}(t)\|_{\text{Lip}}
+Tv2P(sup0tTd(ρn(t),ργ,v(t))>ε).\displaystyle+T\|v\|_{\infty}^{2}P\left(\sup_{0\leq t\leq T}d(\rho_{n}(t),\rho^{\gamma,v}(t))>\varepsilon\right).

By definition, 14ν0T𝕋2|v(t)|2𝑑ργ,v(t)𝑑t=14ν0T𝕋2|p(t)|2𝑑ργ,v(t)𝑑t=𝔸T(ργ,v),\frac{1}{4\nu}\int_{0}^{T}\int_{\mathbb{T}^{2}}|v(t)|^{2}d\rho^{\gamma,v}(t)dt=\frac{1}{4\nu}\int_{0}^{T}\int_{\mathbb{T}^{2}}|\nabla p(t)|^{2}d\rho^{\gamma,v}(t)dt=\mathbb{A}_{T}(\rho^{\gamma,v}), and then the conclusion follows by (5.19). ∎

5.4 Proof of Lemma 2.9

To prove Lemma 2.9, we need certain properties of the special case with v=0v=0 in (2.26).

Lemma 5.4.

For each ε>0\varepsilon>0, ργ,0\rho^{\gamma,0} is smooth on ([ε,T]×𝕋2)([\varepsilon,T]\times\mathbb{T}^{2}) and infxργ,0(ε,x)>0.\inf_{x}\rho^{\gamma,0}(\varepsilon,x)>0.

Proof.

Let u(t):=𝒦ργ,0u(t):=\mathcal{K}*\rho^{\gamma,0}. When v=0v=0, it’s a classical result that the solution of (2.26) can be obtained by (1.3) on 𝕋2\mathbb{T}^{2}. QT(ργ,0)<Q_{T}(\rho^{\gamma,0})<\infty implies u(t)L2(𝕋2)u(t)\in L^{2}(\mathbb{T}^{2}) and ργ,0L2([0,T];L2(𝕋2))\rho^{\gamma,0}\in L^{2}([0,T];L^{2}(\mathbb{T}^{2})). Actually, QT(ργ,0)<Q_{T}(\rho^{\gamma,0})<\infty implies ργ,0\rho^{\gamma,0} is the Leray solution [35] for two dimensional Navier Stokes equation, which is very regular. We will use the results in [24] and [3, 7], which work on 2\mathbb{R}^{2} as well as on 𝕋2\mathbb{T}^{2}, to prove this Lemma.

By Lemma 4.2, for each TT, take ε1(0,ε)\varepsilon_{1}\in(0,\varepsilon) such that S(ργ,0(ε1))<S(\rho^{\gamma,0}(\varepsilon_{1}))<\infty and

ε1TI(ργ,0(t))𝑑t<.\int_{\varepsilon_{1}}^{T}I(\rho^{\gamma,0}(t))dt<\infty.

Similar with Lemma 3.2 of [24], for q[1,2),q\in[1,2),

ργ,0(t)qCqI(ργ,0(t))3/21/q.\|\nabla\rho^{\gamma,0}(t)\|_{q}\leq C_{q}I(\rho^{\gamma,0}(t))^{3/2-1/q}.

Hence

ργ,0L2q/(3q2)((ε1,T);Lq(𝕋2)).\nabla\rho^{\gamma,0}\in L^{2q/(3q-2)}((\varepsilon_{1},T);L^{q}(\mathbb{T}^{2})).

Then by Theorem 2.5 of [24],

ργ,0C([ε1,);L1(𝕋2)C((ε1,);L(𝕋2)).\rho^{\gamma,0}\in C([\varepsilon_{1},\infty);L^{1}(\mathbb{T}^{2})\cap C((\varepsilon_{1},\infty);L^{\infty}(\mathbb{T}^{2})).

This meets the assumptions of the theorem of [7] (which improves Theorem B of [3]), so ργ,0\rho^{\gamma,0} is a smooth classical solution on (ε1,T]×𝕋2(\varepsilon_{1},T]\times\mathbb{T}^{2}.

To show ργ,0(ε)>0\rho^{\gamma,0}(\varepsilon)>0, note that for ργ,0\rho^{\gamma,0} can be see as a solution of second-order parabolic equation

tρνΔρ+ρu=0.\partial_{t}\rho-\nu\Delta\rho+\nabla\rho\cdot u=0.

Since we have proved ργ,0\rho^{\gamma,0} is smooth on ([ε2,T]×𝕋2)\left(\left[\frac{\varepsilon}{2},T\right]\times\mathbb{T}^{2}\right) and ργ,0(ε2)0\rho^{\gamma,0}(\frac{\varepsilon}{2})\geq 0, we conclude the proof by strong maximum principle of parabolic equation. ∎

Lemma 5.5.
e(γ)e(ργ,0(t))=ν0tργ,0(s)122𝑑s;e(\gamma)-e(\rho^{\gamma,0}(t))=\nu\int_{0}^{t}\|\rho^{\gamma,0}(s)-1\|_{2}^{2}ds; (5.21)

There exists a constant C>0C>0, such that for each smooth mollifier JJ,

0tI(Jργ,0(s))𝑑s2ν(S(Jγ)S(Jργ,0(t)))+C(e(γ)e(ργ,0(t))).\int_{0}^{t}I(J*\rho^{\gamma,0}(s))ds\leq\frac{2}{\nu}(S(J*\gamma)-S(J*\rho^{\gamma,0}(t)))+C(e(\gamma)-e(\rho^{\gamma,0}(t))). (5.22)
Proof.

(5.21) follows by Lemma 4.1. Write u(t)=𝒦ργ,0(t)u(t)=\mathcal{K}*\rho^{\gamma,0}(t). By (4.5),

S(Jργ,0(t))S(Jγ)=ν0tI(Jργ,0(s))𝑑s\displaystyle S(J*\rho^{\gamma,0}(t))-S(J*\gamma)=-\nu\int_{0}^{t}I(J*\rho^{\gamma,0}(s))ds
+0t𝕋2(Jργ,0)(t,x)(Jργ,0)(t,x)J(ργ,0u)(t,x)(Jργ,0)(t,x)(Jργ,0)(t,x)𝑑x𝑑t.\displaystyle+\int_{0}^{t}\int_{\mathbb{T}^{2}}\frac{(\nabla J*\rho^{\gamma,0})(t,x)}{(J*\rho^{\gamma,0})(t,x)}\cdot\frac{J*(\rho^{\gamma,0}u)(t,x)}{(J*\rho^{\gamma,0})(t,x)}(J*\rho^{\gamma,0})(t,x)dxdt.

Using the inequality abν2a2+12νb2ab\leq\frac{\nu}{2}a^{2}+\frac{1}{2\nu}b^{2} along with Lemma 8.1.10 of [1], we have

S(Jργ,0(t))S(Jγ)ν20tI(Jργ,0(s))𝑑s\displaystyle S(J*\rho^{\gamma,0}(t))-S(J*\gamma)\leq-\frac{\nu}{2}\int_{0}^{t}I(J*\rho^{\gamma,0}(s))ds
+12ν0t𝕋2|J(ργ,0u)(t,x)(Jργ,0)(t,x)|2(Jργ,0)(t,x)𝑑x𝑑t\displaystyle+\frac{1}{2\nu}\int_{0}^{t}\int_{\mathbb{T}^{2}}\bigg{|}\frac{J*(\rho^{\gamma,0}u)(t,x)}{(J*\rho^{\gamma,0})(t,x)}\bigg{|}^{2}(J*\rho^{\gamma,0})(t,x)dxdt
ν20tI(Jργ,0(s))𝑑s+12ν0t𝕋2|u(t,x)|2ργ,0(t,x)𝑑x𝑑t.\displaystyle\leq-\frac{\nu}{2}\int_{0}^{t}I(J*\rho^{\gamma,0}(s))ds+\frac{1}{2\nu}\int_{0}^{t}\int_{\mathbb{T}^{2}}|u(t,x)|^{2}\rho^{\gamma,0}(t,x)dxdt.

Then (5.22) follows Lemma 3.4 and (5.21). ∎

Proof of Lemma 2.9.

We split the proof into three steps.
Step 1. Construction.
Write Φt\Phi_{t} as the heat kernel on 𝕋2\mathbb{T}^{2}, defined in Appendix B. Given t1,t2>0t_{1},t_{2}>0 and 2t1+t2T2t_{1}+t_{2}\leq T, define

ρ~(t):=ρ~t1,t2(t)={ργ,0(t), if 0t<t1,Φν(tt1)ργ,0(t1), if t1t<t1+t2,Φνt2ργ,0(2t1+t2t), if t1+t2t<2t1+t2Φνt2ρ(t2t1t2), if 2t1+t2tT.\tilde{\rho}(t):=\tilde{\rho}_{t_{1},t_{2}}(t)=\left\{\begin{aligned} &\rho^{\gamma,0}(t),&\text{ if }0\leq t<t_{1},\\ &\Phi_{\nu(t-t_{1})}*\rho^{\gamma,0}(t_{1}),&\text{ if }t_{1}\leq t<t_{1}+t_{2},\\ &\Phi_{\nu t_{2}}*\rho^{\gamma,0}(2t_{1}+t_{2}-t),&\text{ if }t_{1}+t_{2}\leq t<2t_{1}+t_{2}\\ &\Phi_{\nu t_{2}}*\rho(t-2t_{1}-t_{2}),&\text{ if }2t_{1}+t_{2}\leq t\leq T.\end{aligned}\right.

One can easily check that ρ~C([0,T];𝒫(𝕋2))\tilde{\rho}\in C([0,T];\mathcal{P}(\mathbb{T}^{2})). Writing u~=𝒦ρ~\tilde{u}=\mathcal{K}*\tilde{\rho}, we construct v(t)=p(t)v(t)=\nabla p(t) below corresponding to ρ~(t)\tilde{\rho}(t) such that

tρ~(t)νΔρ~(t)+div(ρ~(t)u~(t))+div(ρ~(t)p(t))=0.\partial_{t}\tilde{\rho}(t)-\nu\Delta\tilde{\rho}(t)+{\rm div}\,(\tilde{\rho}(t)\tilde{u}(t))+{\rm div}\,(\tilde{\rho}(t)\nabla p(t))=0. (5.23)

For t[0,t1]t\in[0,t_{1}], take p(t)=0p(t)=0 and (5.23) holds. Since we only care the weak solution, we just define p(t1+t2)=p(2t1+t2)=0p(t_{1}+t_{2})=p(2t_{1}+t_{2})=0, which would not bring any problems.
By Lemma 5.4, ρ~(t1)\tilde{\rho}(t_{1}) is smooth and infxρ~(t1,x)>0\inf_{x}\tilde{\rho}(t_{1},x)>0, so for t1<t<t1+t2t_{1}<t<t_{1}+t_{2}, ρ~(t)\tilde{\rho}(t) is uniform elliptic. Hence (5.23) can be seen as a second-order elliptic equation for p(t)p(t). We expect all the coefficients in (5.23) as an equation for p(t)p(t) is regular (i.e. their derivatives of arbitrary order are uniform bounded), which ensures an unique weak solution p(t)p(t) such that p(t)+2p(t)\|\nabla p(t)\|_{\infty}+\|\nabla^{2}p(t)\|_{\infty} is uniform bounded. However, tρ~(t)\partial_{t}\tilde{\rho}(t) may not be regular when tt is approaching t1t_{1}. But we notice that Φt\Phi_{t} is the heat kernel, which implies

tρ~(t)=νΔρ~(t),t(t1,t2).\partial_{t}\tilde{\rho}(t)=\nu\Delta\tilde{\rho}(t),\quad\forall t\in(t_{1},t_{2}).

Then (5.23) reduces to

div(ρ~(t)u~(t))+ρ~(t)p(t)+ρ~(t)Δp(t)=0,t(t1,t2).{\rm div}\,(\tilde{\rho}(t)\tilde{u}(t))+\nabla\tilde{\rho}(t)\cdot\nabla p(t)+\tilde{\rho}(t)\Delta p(t)=0,\quad\forall t\in(t_{1},t_{2}). (5.24)

Pointwisely for tt, take p(t)p(t) as the weak solution of (5.24). Due to smoothness of ρ~(t1)\tilde{\rho}(t_{1}), all the coefficients in (5.24) is regular, so p(t)+2p(t)\|\nabla p(t)\|_{\infty}+\|\nabla^{2}p(t)\|_{\infty} is uniform bounded and for t1<t<t1+t2t_{1}<t<t_{1}+t_{2}, (5.23) holds.

For tt1+t2,t\geq t_{1}+t_{2}, because of the convolution by Φνt2\Phi_{\nu t_{2}}, ρ~(t)\tilde{\rho}(t) is uniform elliptic. Regard (5.23) as a second-order elliptic equations for p(t)p(t) and take p(t)p(t) as its weak solution. For t(t1+t2,2t1+t2)t\in(t_{1}+t_{2},2t_{1}+t_{2}),

tρ~(t)=Φνt2tργ,0(2t1+t2t)\displaystyle\partial_{t}\tilde{\rho}(t)=-\Phi_{\nu t_{2}}*\partial_{t}\rho^{\gamma,0}(2t_{1}+t_{2}-t) (5.25)
=νΦνt2Δργ,0(2t1+t2t)+Φνt2div(𝐑(ργ,0(2t1+t2t)))\displaystyle=-\nu\Phi_{\nu t_{2}}*\Delta\rho^{\gamma,0}(2t_{1}+t_{2}-t)+\Phi_{\nu t_{2}}*{\rm div}\,(\mathbf{R}(\rho^{\gamma,0}(2t_{1}+t_{2}-t)))
=νΔρ~(t)+Φνt2div(𝐑(ργ,0(2t1+t2t))),\displaystyle=-\nu\Delta\tilde{\rho}(t)+\Phi_{\nu t_{2}}*{\rm div}\,(\mathbf{R}(\rho^{\gamma,0}(2t_{1}+t_{2}-t))),

which is regular, and by (4.5), tρ~(t)\partial_{t}\tilde{\rho}(t) is also regular in (2t1+t2,T)(2t_{1}+t_{2},T), so p(t)+2p(t)\|\nabla p(t)\|_{\infty}+\|\nabla^{2}p(t)\|_{\infty} is uniform bounded when tt1+t2t\geq t_{1}+t_{2}.

Therefore (5.23) holds for the constructed pairs (ρ~,p)(\tilde{\rho},p) and p(t)+2p(t)\|\nabla p(t)\|_{\infty}+\|\nabla^{2}p(t)\|_{\infty} is uniform bounded. As a consequence, p𝒥\nabla p\in\mathcal{J} and ρ~\tilde{\rho} is a weak solution of (2.26) for p.\nabla p.

Step 2. Verify that the limit of 𝔸T(ρ~)\mathbb{A}_{T}(\tilde{\rho}), when taking t10t_{1}\downarrow 0 first and then letting t20t_{2}\downarrow 0, is bounded above by 𝔸T(ρ).\mathbb{A}_{T}(\rho).
Without loss of generality, we assume 2t1+t2<T.2t_{1}+t_{2}<T. Write u~(t)=𝒦ρ~(t),u(t)=𝒦ρ(t)\tilde{u}(t)=\mathcal{K}*\tilde{\rho}(t),u(t)=\mathcal{K}*\rho(t). By definition,

4ν𝔸T(ρ~)=(t1t1+t2+t1+t22t1+t2+2t1+t2T)tρ~(t)νΔρ~(t)+div(ρ~(t)u~(t))1,ρ~(t)2dt\displaystyle 4\nu\mathbb{A}_{T}(\tilde{\rho})=\left(\int_{t_{1}}^{t_{1}+t_{2}}+\int_{t_{1}+t_{2}}^{2t_{1}+t_{2}}+\int_{2t_{1}+t_{2}}^{T}\right)\|\partial_{t}\tilde{\rho}(t)-\nu\Delta\tilde{\rho}(t)+{\rm div}\,(\tilde{\rho}(t)\tilde{u}(t))\|^{2}_{-1,\tilde{\rho}(t)}dt (5.26)
=I+II+III.\displaystyle=I+II+III.

For t(t1,t1+t2)t\in(t_{1},t_{1}+t_{2}), tρ~(t)=νΔρ~(t),\partial_{t}\tilde{\rho}(t)=\nu\Delta\tilde{\rho}(t), we have

I=t1t1+t2div(ρ~(t)u~(t)1,ρ~(t)2dt.I=\int_{t_{1}}^{t_{1}+t_{2}}\|{\rm div}\,(\tilde{\rho}(t)\tilde{u}(t)\|^{2}_{-1,\tilde{\rho}(t)}dt.

By Proportion A.2 and Lemma 3.4,

ICt1t1+t2ρ~(t)22𝑑t.I\leq C\int_{t_{1}}^{t_{1}+t_{2}}\|\tilde{\rho}(t)\|_{2}^{2}dt.

Since the result of Lemma 5.5 also holds for the solution of heat equation,

IC(e(ργ,0(t1))e(Φνt2ργ,0(t1))).I\leq C(e(\rho^{\gamma,0}(t_{1}))-e(\Phi_{\nu t_{2}}*\rho^{\gamma,0}(t_{1}))).

Noting ργ,0\rho^{\gamma,0} is continuous under weak topology, by Fatou’s Lemma and continuity of e(ργ,0)e(\rho^{\gamma,0}) (as a consequence of Lemma 5.5),

lim supt10+IC(e(γ)e(Φνt2γ)).\limsup_{t_{1}\to 0+}I\leq C(e(\gamma)-e(\Phi_{\nu t_{2}}*\gamma)).

By (5.25),

II=t1+t22t1+t22νΔρ~(t)+div(ρ~(t)u~(t))+Φνt2div(𝐑(ργ,0(2t1+t2t)))1,ρ~(t)2𝑑t\displaystyle II=\int_{t_{1}+t_{2}}^{2t_{1}+t_{2}}\|-2\nu\Delta\tilde{\rho}(t)+{\rm div}\,(\tilde{\rho}(t)\tilde{u}(t))+\Phi_{\nu t_{2}}*{\rm div}\,(\mathbf{R}(\rho^{\gamma,0}(2t_{1}+t_{2}-t)))\|_{-1,\tilde{\rho}(t)}^{2}dt
12t1+t22t1+t2νΔρ~(t)1,ρ~(t)2𝑑t+3t1+t22t1+t2div(ρ~(t)u~(t))1,ρ~(t)2𝑑t\displaystyle\leq 12\int_{t_{1}+t_{2}}^{2t_{1}+t_{2}}\|\nu\Delta\tilde{\rho}(t)\|_{-1,\tilde{\rho}(t)}^{2}dt+3\int_{t_{1}+t_{2}}^{2t_{1}+t_{2}}\|{\rm div}\,(\tilde{\rho}(t)\tilde{u}(t))\|^{2}_{-1,\tilde{\rho}(t)}dt
+3t1+t22t1+t2Φνt2div(𝐑(ργ,0(2t1+t2t)))1,ρ~(t)2𝑑t\displaystyle+3\int_{t_{1}+t_{2}}^{2t_{1}+t_{2}}\|\Phi_{\nu t_{2}}*{\rm div}\,\left(\mathbf{R}(\rho^{\gamma,0}(2t_{1}+t_{2}-t))\right)\|^{2}_{-1,\tilde{\rho}(t)}dt
12t1+t22t1+t2νΔρ~(t)1,ρ~(t)2𝑑t+3t1+t22t1+t2div(ρ~(t)u~(t))1,ρ~(t)2𝑑t\displaystyle\leq 12\int_{t_{1}+t_{2}}^{2t_{1}+t_{2}}\|\nu\Delta\tilde{\rho}(t)\|_{-1,\tilde{\rho}(t)}^{2}dt+3\int_{t_{1}+t_{2}}^{2t_{1}+t_{2}}\|{\rm div}\,(\tilde{\rho}(t)\tilde{u}(t))\|^{2}_{-1,\tilde{\rho}(t)}dt
+3t1+t22t1+t2div𝐑(ργ,0(2t1+t2t))1,ργ,0(2t1+t2t)2𝑑t,\displaystyle+3\int_{t_{1}+t_{2}}^{2t_{1}+t_{2}}\|{\rm div}\,\mathbf{R}(\rho^{\gamma,0}(2t_{1}+t_{2}-t))\|^{2}_{-1,\rho^{\gamma,0}(2t_{1}+t_{2}-t)}dt,

where we used Lemma A.3 to get the last inequality. Then by Proposition A.2 and Lemma 3.4 along with Jensen’s inequality,

II=12t1+t22t1+t2I(ρ~(t))𝑑t+3t1+t22t1+t2𝕋2|u~(t)|2𝑑ρ~(t)𝑑t\displaystyle II=12\int_{t_{1}+t_{2}}^{2t_{1}+t_{2}}I(\tilde{\rho}(t))dt+3\int_{t_{1}+t_{2}}^{2t_{1}+t_{2}}\int_{\mathbb{T}^{2}}|\tilde{u}(t)|^{2}d\tilde{\rho}(t)dt
+3t1+t22t1+t2𝕋2|𝒦ργ,0(2t1+t2t)|2𝑑ργ,0(2t1+t2t)𝑑t\displaystyle+3\int_{t_{1}+t_{2}}^{2t_{1}+t_{2}}\int_{\mathbb{T}^{2}}|\mathcal{K}*\rho^{\gamma,0}(2t_{1}+t_{2}-t)|^{2}d\rho^{\gamma,0}(2t_{1}+t_{2}-t)dt
12ν20t1I(ργ,0(t))𝑑t+C0t1ργ,0(t)122𝑑t,\displaystyle\leq 12\nu^{2}\int_{0}^{t_{1}}I(\rho^{\gamma,0}(t))dt+C\int_{0}^{t_{1}}\|\rho^{\gamma,0}(t)-1\|_{2}^{2}dt,

which is, in view of Lemma 5.5, bounded by

24ν(S(Φνt2γ)S(Φνt2ργ,0(t1)))+C(e(γ)e(ργ,0(t1))).24\nu(S(\Phi_{\nu t_{2}}*\gamma)-S(\Phi_{\nu t_{2}}*\rho^{\gamma,0}(t_{1})))+C(e(\gamma)-e(\rho^{\gamma,0}(t_{1}))).

Also by Fatou’s Lemma,

lim supt1II0.\limsup_{t_{1}\to\infty}II\leq 0.

By Lemma A.3, for each δ>0\delta>0

III(1δ)10T2t1t2tρ(t)νΔρ(t)+div(ρ(t)u(t))1,ρ(t)2𝑑t\displaystyle III\leq(1-\delta)^{-1}\int_{0}^{T-2t_{1}-t_{2}}\|\partial_{t}\rho(t)-\nu\Delta\rho(t)+{\rm div}\,(\rho(t)u(t))\|^{2}_{-1,\rho(t)}dt
+δ10T2t1t2𝕋2|Φνt2u(t)u(t)|2𝑑ρ(t)𝑑t.\displaystyle+\delta^{-1}\int_{0}^{T-2t_{1}-t_{2}}\int_{\mathbb{T}^{2}}|\Phi_{\nu t_{2}}*u(t)-u(t)|^{2}d\rho(t)dt.

Combining these estimations, we have

lim supt10+𝔸T(ρ~)(1δ)1𝔸T(ρ)+(4νδ)10T𝕋2|Φνt2u(t)u(t)|2𝑑ρ(t)𝑑t\displaystyle\limsup_{t_{1}\to 0+}\mathbb{A}_{T}(\tilde{\rho})\leq(1-\delta)^{-1}\mathbb{A}_{T}(\rho)+(4\nu\delta)^{-1}\int_{0}^{T}\int_{\mathbb{T}^{2}}|\Phi_{\nu t_{2}}*u(t)-u(t)|^{2}d\rho(t)dt
+C(e(γ)e(Φνt2γ)).\displaystyle+C(e(\gamma)-e(\Phi_{\nu t_{2}}*\gamma)).

Note that QT(ρ)<Q_{T}(\rho)<\infty implies 0Tρ(t)22𝑑t<,\int_{0}^{T}\|\rho(t)\|_{2}^{2}dt<\infty, so by Lemma 3.4 and dominated convergence theorem, the second term vanishes when t20t_{2}\downarrow 0. The limit of third term is also non-positive by Fatou’s Lemma. Thus

lim supt20+lim supt10+𝔸T(ρ~)(1δ)1𝔸T(ρ).\limsup_{t_{2}\to 0+}\limsup_{t_{1}\to 0+}\mathbb{A}_{T}(\tilde{\rho})\leq(1-\delta)^{-1}\mathbb{A}_{T}(\rho).

Due to the arbitrariness of δ\delta we conclude that

limt20+limt10+𝔸T(ρ~)𝔸T(ρ).\lim_{t_{2}\to 0+}\lim_{t_{1}\to 0+}\mathbb{A}_{T}(\tilde{\rho})\leq\mathbb{A}_{T}(\rho).

Step 3. Convergence.

It’s easy to check for each t[0,T]t\in[0,T], limt20+limt10+d(ρ~(t),ρ(t))=0.\lim_{t_{2}\to 0+}\lim_{t_{1}\to 0+}d(\tilde{\rho}(t),\rho(t))=0. To conclude the proof, we need strength this point-wise convergence into

limt20+limt10+sup0tTd(ρ~(t),ρ(t))=0,\lim_{t_{2}\to 0+}\lim_{t_{1}\to 0+}\sup_{0\leq t\leq T}d(\tilde{\rho}(t),\rho(t))=0,

which can be obtained by the compactness result below. ∎

Lemma 5.6.

If (ρn)n1(\rho_{n})_{n\geq 1} is a sequence of weak solutions of (2.26) for vn=pn𝒥v_{n}=\nabla p_{n}\in\mathcal{J} with

supn10T𝕋2vn2(t,x)ρ(t,x)𝑑x𝑑t<,\sup_{n\geq 1}\int_{0}^{T}\int_{\mathbb{T}^{2}}v_{n}^{2}(t,x)\rho(t,x)dxdt<\infty,

then (ρn)n1(\rho_{n})_{n\geq 1} is relatively compact set in C([0,T];𝒫(𝕋2)).C([0,T];\mathcal{P}(\mathbb{T}^{2})).

Proof.

Suppose

0T𝕋2vn2(t,x)ρ(t,x)𝑑x𝑑tK\int_{0}^{T}\int_{\mathbb{T}^{2}}v_{n}^{2}(t,x)\rho(t,x)dxdt\leq K

for certain K0.K\geq 0. Note that by (2.5) and Holder’s inequality

|ϕ,ρn(t+h)ρn(t)|C2ϕh+tt+h𝕋2vn(s,x)ϕ(s,x)ρn(s,x)𝑑x𝑑s\displaystyle|\left\langle\phi,\rho_{n}(t+h)-\rho_{n}(t)\right\rangle|\leq C\|\nabla^{2}\phi\|_{\infty}h+\int_{t}^{t+h}\int_{\mathbb{T}^{2}}v_{n}(s,x)\cdot\nabla\phi(s,x)\rho_{n}(s,x)dxds
Cϕh+(tt+h𝕋2|ϕ(s,x)|2ρ(s,x)𝑑x𝑑s)12K12\displaystyle\leq C_{\phi}h+\left(\int_{t}^{t+h}\int_{\mathbb{T}^{2}}|\nabla\phi(s,x)|^{2}\rho(s,x)dxds\right)^{\frac{1}{2}}K^{\frac{1}{2}}
Cϕ(h+h12K12).\displaystyle\leq C_{\phi}\left(h+h^{\frac{1}{2}}K^{\frac{1}{2}}\right).

Then we conclude the proof by [29] and Theorems 8.6 and 8.8 in Chapter 3 of [18]. ∎

Appendix A Weighted Sobolev space

In this section, we give some property of 1,μ\|\cdot\|_{-1,\mu} for μ𝒫(𝕋2)\mu\in\mathcal{P}(\mathbb{T}^{2}). Recall that

m1,μ2=supϕC(𝕋2){2ϕ,m𝕋2|ϕ(x)|2𝑑m}.\|m\|_{-1,\mu}^{2}=\sup_{\phi\in C^{\infty}(\mathbb{T}^{2})}\left\{2\left\langle\phi,m\right\rangle-\int_{\mathbb{T}^{2}}|\nabla\phi(x)|^{2}dm\right\}.
Lemma A.1.

Suppose 𝕋2|v|2𝑑μ<,\int_{\mathbb{T}^{2}}|v|^{2}d\mu<\infty, then

div(μv)1,μ2𝕋2|v|2𝑑μ.\|-{\rm div}\,(\mu v)\|_{-1,\mu}^{2}\leq\int_{\mathbb{T}^{2}}|v|^{2}d\mu.
Proof.

By definition of 1,μ\|\cdot\|_{-1,\mu} and the inequality 2aba2b22ab-a^{2}\leq b^{2},

div(μv)1,μ2=supϕC(𝕋2){2𝕋2vϕdμ𝕋2|ϕ|2𝑑μ}𝕋2|v|2𝑑μ.\|-{\rm div}\,(\mu v)\|^{2}_{-1,\mu}=\sup_{\phi\in C^{\infty}(\mathbb{T}^{2})}\left\{2\int_{\mathbb{T}^{2}}v\cdot\nabla\phi d\mu-\int_{\mathbb{T}^{2}}|\nabla\phi|^{2}d\mu\right\}\leq\int_{\mathbb{T}^{2}}|v|^{2}d\mu.

Recall

S(μ)={𝕋2μ(x)logμ(x)𝑑x,if μ(dx)=μ(x)dx,,otherwise,S(\mu)=\left\{\begin{aligned} &\int_{\mathbb{T}^{2}}\mu(x)\log\mu(x)dx,&\text{if }\mu(dx)=\mu(x)dx,\\ &\infty,&otherwise,\end{aligned}\right.

and

I(μ)={𝕋2|μ(x)|2μ(x)𝑑x,if μ(dx)=μ(x)dx and μL1(𝕋2),,otherwise.I(\mu)=\left\{\begin{aligned} &\int_{\mathbb{T}^{2}}\frac{|\nabla\mu(x)|^{2}}{\mu(x)}dx,&\text{if }\mu(dx)=\mu(x)dx\text{ and }\nabla\mu\in L^{1}(\mathbb{T}^{2}),\\ &\infty,&otherwise.\end{aligned}\right.
Proposition A.2.

For μ𝒫(𝕋2),\mu\in\mathcal{P}(\mathbb{T}^{2}),

Δμ1,μ2=I(μ) if S(μ)<;\|\Delta\mu\|^{2}_{-1,\mu}=I(\mu)\text{ if }S(\mu)<\infty;
div[μ(𝒦μ)]1,μ2𝕋2|𝒦μ|2𝑑μ if μL2(𝕋2).\|{\rm div}\,[\mu(\mathcal{K}*\mu)]\|^{2}_{-1,\mu}\leq\int_{\mathbb{T}^{2}}|\mathcal{K}*\mu|^{2}d\mu\text{ if }\mu\in L^{2}(\mathbb{T}^{2}).
Proof.

The first equality is proved by Theorem D.45. in [20] in 2\mathbb{R}^{2}, which can be adapted to 𝕋2\mathbb{T}^{2}. The second inequality follows from Lemma A.1 and Lemma 3.4. ∎

Lemma A.3.

Let JJ be an arbitrary smooth mollifier. Then for each δ>0,\delta>0,

Jm+div(Jμ(Jv))1,Jμ2(1δ)1m+div(μv)1,μ2+δ1𝕋2|Jvv|2𝑑μ.\|J*m+{\rm div}\,(J*\mu(J*v))\|^{2}_{-1,J*\mu}\leq(1-\delta)^{-1}\|m+{\rm div}\,(\mu v)\|^{2}_{-1,\mu}+\delta^{-1}\int_{\mathbb{T}^{2}}|J*v-v|^{2}d\mu.
Proof.

By Jensen’s inequality, for each φC(𝕋2)\varphi\in C^{\infty}(\mathbb{T}^{2}),

2Jm+div(Jμ(Jv)),φ𝕋2|φ|2d(Jμ)\displaystyle 2\left\langle J*m+{\rm div}\,(J*\mu(J*v)),\varphi\right\rangle-\int_{\mathbb{T}^{2}}|\nabla\varphi|^{2}d(J*\mu)
2m+div(μ(Jv)),Jφ𝕋2|Jφ|2𝑑μ\displaystyle\leq 2\left\langle m+{\rm div}\,(\mu(J*v)),J*\varphi\right\rangle-\int_{\mathbb{T}^{2}}|\nabla J*\varphi|^{2}d\mu
=2m+div(μv),Jφ+2div(μ(Jvv)),Jφ𝕋2|Jφ|2𝑑μ.\displaystyle=2\left\langle m+{\rm div}\,(\mu v),J*\varphi\right\rangle+2\left\langle{\rm div}\,(\mu(J*v-v)),J*\varphi\right\rangle-\int_{\mathbb{T}^{2}}|\nabla J*\varphi|^{2}d\mu.

Using 2abδa2+1δb2,2ab\leq\delta a^{2}+\frac{1}{\delta}b^{2},

2div(μ(Jvv)),Jφδ1𝕋2|Jvv|2𝑑μ+δ𝕋2|Jφ|2𝑑μ.2\left\langle{\rm div}\,(\mu(J*v-v)),J*\varphi\right\rangle\leq\delta^{-1}\int_{\mathbb{T}^{2}}|J*v-v|^{2}d\mu+\delta\int_{\mathbb{T}^{2}}|\nabla J*\varphi|^{2}d\mu.

Hence,

2Jm+div(Jμ(Jv)),φ𝕋2|φ|2d(Jμ)\displaystyle 2\left\langle J*m+{\rm div}\,(J*\mu(J*v)),\varphi\right\rangle-\int_{\mathbb{T}^{2}}|\nabla\varphi|^{2}d(J*\mu)
2m+div(μv),Jφ+δ1𝕋2|Jvv|2𝑑μ(1δ)𝕋2|Jφ|2𝑑μ\displaystyle\leq 2\left\langle m+{\rm div}\,(\mu v),J*\varphi\right\rangle+\delta^{-1}\int_{\mathbb{T}^{2}}|J*v-v|^{2}d\mu-(1-\delta)\int_{\mathbb{T}^{2}}|\nabla J*\varphi|^{2}d\mu
=(1δ)1(2m+div(μv),(1δ)Jφ𝕋2|(1δ)Jφ|2𝑑μ)\displaystyle=(1-\delta)^{-1}\left(2\left\langle m+{\rm div}\,(\mu v),(1-\delta)J*\varphi\right\rangle-\int_{\mathbb{T}^{2}}|(1-\delta)\nabla J*\varphi|^{2}d\mu\right)
+δ1𝕋2|Jvv|2𝑑μδ1𝕋2|Jvv|2𝑑μ+(1δ)1m+div(μv)1,μ2.\displaystyle+\delta^{-1}\int_{\mathbb{T}^{2}}|J*v-v|^{2}d\mu\leq\delta^{-1}\int_{\mathbb{T}^{2}}|J*v-v|^{2}d\mu+(1-\delta)^{-1}\|m+{\rm div}\,(\mu v)\|^{2}_{-1,\mu}.

Taking the supremum for φ\varphi we conclude the result. ∎

Appendix B Heat kernel on torus

We define the heat kernel on 𝕋2\mathbb{T}^{2},

Φ(t,x)=n,m=e4π2(n2+m2)texp(2πi(nx1+mx2)),x=(x1,x2),\Phi(t,x)=\sum_{n,m=-\infty}^{\infty}e^{-4\pi^{2}(n^{2}+m^{2})t}\exp(2\pi i(nx_{1}+mx_{2})),\quad x=(x_{1},x_{2}), (B.1)

for t>0.t>0.

Theorem B.1.

(1) For any t>0,t>0, Φt(x):=Φ(t,x)\Phi_{t}(x):=\Phi(t,x) is a well defined smooth mollifier.
(2) Φt(x)Φs(x)=Φt+s(x).\Phi_{t}(x)*\Phi_{s}(x)=\Phi_{t+s}(x).
(3) For γ𝒫(𝕋2)\gamma\in\mathcal{P}(\mathbb{T}^{2}) and ν>0\nu>0, ρ(t)=Φνtγ\rho(t)=\Phi_{\nu t}*\gamma satisfies the heat equation:

tρ(t,x)=νΔρ(t,x),t>0,\frac{\partial}{\partial t}\rho(t,x)=\nu\Delta\rho(t,x),\quad t>0, (B.2)

and ρC([0,T];𝒫(𝕋2))\rho\in C([0,T];\mathcal{P}(\mathbb{T}^{2})).

Proof.

(1) Based on the theory of uniformly convergent series, Φt(x,y)\Phi_{t}(x,y) is a well defined smooth function. Since the integral of every term in (B.1) is zero except the case (n,m)=(0,0)(n,m)=(0,0), 𝕋2Φt𝑑x=1.\int_{\mathbb{T}^{2}}\Phi_{t}dx=1. By Poisson summation formula (e.g. see Theorem 3.1.17 of [25]),

Φ(t,x)=n,m=14πte(x1n)2+(x2m)24t.\Phi(t,x)=\sum_{n,m=-\infty}^{\infty}\frac{1}{4\pi t}e^{-\frac{(x_{1}-n)^{2}+(x_{2}-m)^{2}}{4t}}. (B.3)

Hence Φ(t,x)>0.\Phi(t,x)>0. We conclude that Φt(x)\Phi_{t}(x) is a smooth mollifier.
(2) It follows by direct calculation.
(3) Since 𝕋2Φt(x)𝑑x=1\int_{\mathbb{T}^{2}}\Phi_{t}(x)dx=1, for each t0,t\geq 0, ρ(t)𝒫(𝕋2)\rho(t)\in\mathcal{P}(\mathbb{T}^{2}). By (B.3),

limt0+|x|>δΦt(x)𝑑x=0.\lim_{t\to 0+}\int_{|x|>\delta}\Phi_{t}(x)dx=0.

Hence ρ\rho is continuous under weak topology at 0. By (2), we have ρC([0,T];𝒫(𝕋2))\rho\in C([0,T];\mathcal{P}(\mathbb{T}^{2})).
For t>0t>0, since each item in the sum of (B.1) for νt\nu t satisfies (B.2), Φνt\Phi_{\nu t} also satisfies (B.2), so does ρ(t)\rho(t). ∎

Appendix C Property of energy functional

By (1.5), 𝒩\mathcal{N} is bounded from below, so

𝒩γ,γ=𝕋2𝒩(xy)γ(dx)γ(dy)\left\langle\mathcal{N}*\gamma,\gamma\right\rangle=\int_{\mathbb{T}^{2}}\mathcal{N}(x-y)\gamma(dx)\gamma(dy)

is well-defined for any probability measure γ\gamma. By Young’s inequality, for any finite signed measure η\eta on 𝕋2\mathbb{T}^{2}, 𝒩η\mathcal{N}*\eta,𝒩η\nabla\mathcal{N}*\eta is in L1(𝕋2)L^{1}(\mathbb{T}^{2}).

Lemma C.1.

Suppose γ,η\gamma,\eta are probability measures. we have
(1) 𝒩γ,𝒦γ\mathcal{N}*\gamma,\mathcal{K}*\gamma has a weak derivative and

(𝒩γ)=𝒩γ;\nabla(\mathcal{N}*\gamma)=\nabla\mathcal{N}*\gamma; (C.1)
div(𝒦γ)=0,curl(𝒦γ)=γ.{\rm div}\,(\mathcal{K}*\gamma)=0,\quad{\rm curl}\,(\mathcal{K}*\gamma)=\gamma. (C.2)

(2) If JJ is a smooth mollifier,

𝒩JJγ,η=𝒩Jγ,𝒩Jη.\left\langle\mathcal{N}*J*J*\gamma,\eta\right\rangle=\left\langle\nabla\mathcal{N}*J*\gamma,\nabla\mathcal{N}*J*\eta\right\rangle. (C.3)
Proof.

We start with (1). (C.1) is standard and follows from Fubini’s lemma and integration by parts against test functions ϕC(𝕋2)\phi\in C^{\infty}(\mathbb{T}^{2}) to show the equality in the distribution sense. (C.2) can be proved by a similar argument as the proof of Lemma 7.1 of [21].
Turing to (2), by (1.1),

Jη=Δ𝒩Jη+1.J*\eta=-\Delta\mathcal{N}*J*\eta+1.

Noting that 𝕋2𝒩𝑑x=0\int_{\mathbb{T}^{2}}\mathcal{N}dx=0, by Fubini’s lemma and integration by parts

𝒩JJγ,η=𝒩Jγ,Δ𝒩Jη=𝒩Jγ,𝒩Jη.\left\langle\mathcal{N}*J*J*\gamma,\eta\right\rangle=\left\langle\mathcal{N}*J*\gamma,-\Delta\mathcal{N}*J*\eta\right\rangle=\left\langle\nabla\mathcal{N}*J*\gamma,\nabla\mathcal{N}*J*\eta\right\rangle.

Proposition C.2.

If γ,η𝒫(𝕋2)\gamma,\eta\in\mathcal{P}(\mathbb{T}^{2}) and 𝒩γ,𝒩ηL2(𝕋2)\nabla\mathcal{N}*\gamma,\nabla\mathcal{N}*\eta\in L^{2}(\mathbb{T}^{2}), then

𝒩γ,η=𝒩γ,𝒩η;\left\langle\mathcal{N}*\gamma,\eta\right\rangle=\left\langle\nabla\mathcal{N}*\gamma,\nabla\mathcal{N}*\eta\right\rangle; (C.4)
𝒩(γη),γη=𝒩(γη)22=𝒦(γη)22.\left\langle\mathcal{N}*(\gamma-\eta),\gamma-\eta\right\rangle=\|\nabla\mathcal{N}*(\gamma-\eta)\|_{2}^{2}=\|\mathcal{K}*(\gamma-\eta)\|_{2}^{2}. (C.5)

In particular, If γ𝒫(𝕋2)\gamma\in\mathcal{P}(\mathbb{T}^{2}), then

𝒩γ,γ=𝒩γ22=𝒦γ22,\left\langle\mathcal{N}*\gamma,\gamma\right\rangle=\|\nabla\mathcal{N}*\gamma\|_{2}^{2}=\|\mathcal{K}*\gamma\|_{2}^{2}, (C.6)

where we admit =\infty=\infty if 𝒩γL2(𝕋2).\nabla\mathcal{N}*\gamma\notin L^{2}(\mathbb{T}^{2}).

Proof.

We first prove (C.4).

Take smooth mollifiers Gn=ζnζnG_{n}=\zeta_{n}*\zeta_{n} defined in Section 3.1. By (3.5),

Gn𝒩(x)𝒩(x)+Cmn2.G_{n}*\mathcal{N}(x)\leq\mathcal{N}(x)+\frac{C}{m_{n}^{2}}.

Combining with (C.3),

𝒩γ,η𝒩Gnγ,ηCmn2\displaystyle\left\langle\mathcal{N}*\gamma,\eta\right\rangle\geq\left\langle\mathcal{N}*G_{n}*\gamma,\eta\right\rangle-\frac{C}{m_{n}^{2}} (C.7)
=𝒩ζnγ,ζnηCmn2=ζn𝒩γ,ζn𝒩ηCmn2.\displaystyle=\left\langle\mathcal{N}*\zeta_{n}*\gamma,\zeta_{n}*\eta\right\rangle-\frac{C}{m_{n}^{2}}=\left\langle\zeta_{n}*\nabla\mathcal{N}*\gamma,\zeta_{n}*\nabla\mathcal{N}*\eta\right\rangle-\frac{C}{m_{n}^{2}}.

Since 𝒩γ,𝒩ηL2(𝕋2)\nabla\mathcal{N}*\gamma,\nabla\mathcal{N}*\eta\in L^{2}(\mathbb{T}^{2}), by property of smooth mollifiers, ζn𝒩γ𝒩γ,\zeta_{n}*\nabla\mathcal{N}*\gamma\to\nabla\mathcal{N}*\gamma, ζn𝒩η𝒩η\zeta_{n}*\nabla\mathcal{N}*\eta\to\nabla\mathcal{N}*\eta in L2(𝕋2)L^{2}(\mathbb{T}^{2}), so

𝒩γ,ηlim supnζn𝒩γ,ζn𝒩η=𝒩γ,𝒩η.\left\langle\mathcal{N}*\gamma,\eta\right\rangle\geq\limsup_{n\to\infty}\left\langle\zeta_{n}*\nabla\mathcal{N}*\gamma,\zeta_{n}*\nabla\mathcal{N}*\eta\right\rangle=\left\langle\nabla\mathcal{N}*\gamma,\nabla\mathcal{N}*\eta\right\rangle.

By Fatou’s lemma and (C.3),

lim infnζn𝒩γ,ζn𝒩η=lim infn𝒩Gnγ,η𝒩γ,η.\liminf_{n\to\infty}\left\langle\zeta_{n}*\nabla\mathcal{N}*\gamma,\zeta_{n}*\nabla\mathcal{N}*\eta\right\rangle=\liminf_{n\to\infty}\left\langle\mathcal{N}*G_{n}*\gamma,\eta\right\rangle\geq\left\langle\mathcal{N}*\gamma,\eta\right\rangle.

Hence,

𝒩γ,𝒩η=limnζn𝒩γ,ζn𝒩η𝒩γ,η.\left\langle\nabla\mathcal{N}*\gamma,\nabla\mathcal{N}*\eta\right\rangle=\lim_{n\to\infty}\left\langle\zeta_{n}*\nabla\mathcal{N}*\gamma,\zeta_{n}*\nabla\mathcal{N}*\eta\right\rangle\geq\left\langle\mathcal{N}*\gamma,\eta\right\rangle.

So (C.4) holds and (C.5) is a direct corollary.

Now we turn to proving (C.6). Take γ=η\gamma=\eta in (C.7) and let nn\to\infty and then we obtain

𝒩γ,γlim supnζn𝒩γ22.\left\langle\mathcal{N}*\gamma,\gamma\right\rangle\geq\limsup_{n\to\infty}\|\zeta_{n}*\nabla\mathcal{N}*\gamma\|_{2}^{2}.

By Fatou’s lemma,

𝒩γ,γ𝒩γ22.\left\langle\mathcal{N}*\gamma,\gamma\right\rangle\geq\|\nabla\mathcal{N}*\gamma\|_{2}^{2}.

Since ζn𝒩γL2(𝕋2)\zeta_{n}*\nabla\mathcal{N}*\gamma\in L^{2}(\mathbb{T}^{2}), by (C.4), ζn𝒩γ,ζnγ=𝒩ζnγ22\left\langle\zeta_{n}*\mathcal{N}*\gamma,\zeta_{n}*\gamma\right\rangle=\|\nabla\mathcal{N}*\zeta_{n}*\gamma\|_{2}^{2}. By Fatou’s lemma and Jensen’s inequality,

𝒩γ,γlim infnζn𝒩γ,ζnγ=lim infn𝒩ζnγ22𝒩γ22.\left\langle\mathcal{N}*\gamma,\gamma\right\rangle\leq\liminf_{n\to\infty}\left\langle\zeta_{n}*\mathcal{N}*\gamma,\zeta_{n}*\gamma\right\rangle=\liminf_{n\to\infty}\|\nabla\mathcal{N}*\zeta_{n}*\gamma\|_{2}^{2}\leq\|\nabla\mathcal{N}*\gamma\|_{2}^{2}.

At the end of this section, we give some estimates for γ𝒫(𝕋2)\gamma\in\mathcal{P}(\mathbb{T}^{2}) with e(γ)<.e(\gamma)<\infty.

Lemma C.3.

There exists a constant C𝒩C_{\mathcal{N}}, such that for each γ𝒫(𝕋2)\gamma\in\mathcal{P}(\mathbb{T}^{2}) with e(γ)<e(\gamma)<\infty and δ<12\delta<\frac{1}{2},

(γγ)({(x,y):r(x,y)δ})C𝒩+4πe(γ)logδ.(\gamma\otimes\gamma)(\{(x,y):r(x,y)\leq\delta\})\leq\frac{C_{\mathcal{N}}+4\pi e(\gamma)}{-\log\delta}. (C.8)
γ(B¯δ(x))(C𝒩+4πe(γ)log(2δ))12.\gamma(\overline{B}_{\delta}(x))\leq\bigg{(}\frac{C_{\mathcal{N}}+4\pi e(\gamma)}{-\log(2\delta)}\bigg{)}^{\frac{1}{2}}. (C.9)

In addition, given a smooth mollifier JJ and a non-negative sequence mnm_{n}\to\infty, let Jn(x):=mn2J(mn1x)J_{n}(x):=m_{n}^{2}J\left(m_{n}^{-1}x\right). Then for each δ<12\delta<\frac{1}{2} there exists nδn_{\delta} such that for each γ𝒫(𝕋2)\gamma\in\mathcal{P}(\mathbb{T}^{2}) and nnδn\geq n_{\delta},

(γγ)({(x,y):r(x,y)δ})2C𝒩+8πe(Jnγ)logδ.(\gamma\otimes\gamma)(\{(x,y):r(x,y)\leq\delta\})\leq\frac{2C_{\mathcal{N}}+8\pi e(J_{n}*\gamma)}{-\log\delta}. (C.10)
Proof.

By (1.5), take C𝒩C_{\mathcal{N}} such that C𝒩+2π𝒩(xy)logr(x,y).C_{\mathcal{N}}+2\pi\mathcal{N}(x-y)\geq-\log r(x,y).

C𝒩+4πe(γ)𝕋2logr(x,y)γ(dx)γ(dy)\displaystyle C_{\mathcal{N}}+4\pi e(\gamma)\geq-\int_{\mathbb{T}^{2}}\log r(x,y)\gamma(dx)\gamma(dy)
inf|x|δlog(|x|)(γγ)({(x,y):r(x,y)δ}).\displaystyle\geq-\inf_{|x|\leq\delta}\log(|x|)(\gamma\otimes\gamma)(\{(x,y):r(x,y)\leq\delta\}).

Hence (C.8) holds. (C.9) follows from

γ(B¯δ(x))=(γγ)12(B¯δ(x)×B¯δ(x))(γγ)12({(x,y):r(x,y)2δ}).\gamma(\overline{B}_{\delta}(x))=(\gamma\otimes\gamma)^{\frac{1}{2}}(\overline{B}_{\delta}(x)\times\overline{B}_{\delta}(x))\leq(\gamma\otimes\gamma)^{\frac{1}{2}}(\{(x,y):r(x,y)\leq 2\delta\}).

By the property of smooth mollifiers, there exists nδn_{\delta} such that for each nnδn\geq n_{\delta},

C𝒩+2π(JnJn𝒩)(xy){0,r(x,y)>δ,12logδ,r(x,y)<δ,C_{\mathcal{N}}+2\pi(J_{n}*J_{n}*\mathcal{N})(x-y)\geq\left\{\begin{aligned} &0,&r(x,y)>\delta,\\ &-\frac{1}{2}\log\delta,&r(x,y)<\delta,\end{aligned}\right.

and similarly we can prove (C.10). ∎

{acks}

[Acknowledgments] The authors would like to thank Prof. Jin Feng from University of Kansas for his generous help.

The second author was supported by Biomedical Pioneering Innovation Center, Peking University. {funding} The authors were supported by NSFC (No. 11971037).

References

  • [1] {bbook}[author] \bauthor\bsnmAmbrosio, \bfnmLuigi\binitsL., \bauthor\bsnmGigli, \bfnmNicola\binitsN. and \bauthor\bsnmSavaré, \bfnmGiuseppe\binitsG. (\byear2008). \btitleGradient flows: in metric spaces and in the space of probability measures. \bseriesLectures in Mathematics. ETH Zürich. \bpublisherBirkhäuser Verlag, Basel. \endbibitem
  • [2] {bbook}[author] \bauthor\bsnmAnderson, \bfnmGreg W\binitsG. W., \bauthor\bsnmGuionnet, \bfnmAlice\binitsA. and \bauthor\bsnmZeitouni, \bfnmOfer\binitsO. (\byear2010). \btitleAn introduction to random matrices. \bpublisherCambridge university press. \endbibitem
  • [3] {barticle}[author] \bauthor\bsnmBen-Artzi, \bfnmMatania\binitsM. (\byear1994). \btitleGlobal solutions of two-dimensional Navier-Stokes and Euler equations. \bjournalArchive for Rational Mechanics and Analysis \bvolume128 \bpages329-358. \endbibitem
  • [4] {barticle}[author] \bauthor\bsnmBerman, \bfnmRobert J\binitsR. J. (\byear2017). \btitleLarge deviations for Gibbs measures with singular Hamiltonians and emergence of Kähler–Einstein metrics. \bjournalCommunications in Mathematical Physics \bvolume354 \bpages1133-1172. \endbibitem
  • [5] {barticle}[author] \bauthor\bsnmBertini, \bfnmLorenzo\binitsL., \bauthor\bsnmLandim, \bfnmClaudio\binitsC. and \bauthor\bsnmMourragui, \bfnmMustapha\binitsM. (\byear2009). \btitleDynamical large deviations for the boundary driven weakly asymmetric exclusion process. \bjournalThe Annals of Probability \bvolume37 \bpages2357-2403. \endbibitem
  • [6] {barticle}[author] \bauthor\bsnmBresch, \bfnmDidier\binitsD., \bauthor\bsnmJabin, \bfnmPierre-Emmanuel\binitsP.-E. and \bauthor\bsnmWang, \bfnmZhenfu\binitsZ. (\byear2019). \btitleOn mean-field limits and quantitative estimates with a large class of singular kernels: Application to the Patlak–Keller–Segel model. \bjournalComptes Rendus Mathematique \bvolume357 \bpages708-720. \endbibitem
  • [7] {barticle}[author] \bauthor\bsnmBrezis, \bfnmHaïm\binitsH. (\byear1994). \btitleRemarks on the preceding paper by M. Ben-Artzi” Global solutions of two-dimensional Navier-Stokes and Euler equations”. \bjournalArchive for Rational Mechanics and Analysis \bvolume128 \bpages359-360. \endbibitem
  • [8] {barticle}[author] \bauthor\bsnmBudhiraja, \bfnmAmarjit\binitsA., \bauthor\bsnmDupuis, \bfnmPaul\binitsP. and \bauthor\bsnmFischer, \bfnmMarkus\binitsM. (\byear2012). \btitleLarge deviation properties of weakly interacting processes via weak convergence methods. \bjournalThe Annals of Probability \bvolume40 \bpages74-102. \endbibitem
  • [9] {barticle}[author] \bauthor\bsnmCaglioti, \bfnmEmanuele\binitsE., \bauthor\bsnmLions, \bfnmPierre-Louis\binitsP.-L., \bauthor\bsnmMarchioro, \bfnmCarlo\binitsC. and \bauthor\bsnmPulvirenti, \bfnmMario\binitsM. (\byear1992). \btitleA special class of stationary flows for two-dimensional Euler equations: a statistical mechanics description. \bjournalCommunications in Mathematical Physics \bvolume143 \bpages501-525. \endbibitem
  • [10] {barticle}[author] \bauthor\bsnmCaglioti, \bfnmEmanuele\binitsE., \bauthor\bsnmLions, \bfnmPierre-Louis\binitsP.-L., \bauthor\bsnmMarchioro, \bfnmCarlo\binitsC. and \bauthor\bsnmPulvirenti, \bfnmM\binitsM. (\byear1995). \btitleA special class of stationary flows for two-dimensional Euler equations: a statistical mechanics description. Part II. \bjournalCommunications in Mathematical Physics \bvolume174 \bpages229-260. \endbibitem
  • [11] {barticle}[author] \bauthor\bsnmDawson, \bfnmDonald A\binitsD. A. and \bauthor\bsnmGärtner, \bfnmJürgen\binitsJ. (\byear1987). \btitleLarge deviations from the McKean-Vlasov limit for weakly interacting diffusions. \bjournalStochastics: An International Journal of Probability Stochastic Processes \bvolume20 \bpages247-308. \endbibitem
  • [12] {barticle}[author] \bauthor\bsnmDelort, \bfnmJean-Marc\binitsJ.-M. (\byear1991). \btitleExistence de nappes de tourbillon en dimension deux. \bjournalJournal of the American Mathematical Society \bvolume4 \bpages553-586. \endbibitem
  • [13] {bbook}[author] \bauthor\bsnmDembod, \bfnmA\binitsA. and \bauthor\bsnmZeltouni, \bfnmO\binitsO. (\byear2010). \btitleLarge Deviations Techniques and Applications, \bedition2 ed. \bseriesStochastic Modelling and Applied Probability. \bpublisherSpringer, Berlin, Heidelberg. \endbibitem
  • [14] {barticle}[author] \bauthor\bsnmDonsker, \bfnmMD\binitsM. and \bauthor\bsnmVaradhan, \bfnmSRS\binitsS. (\byear1989). \btitleLarge deviations from a hydrodynamic scaling limit. \bjournalCommunications on Pure and Applied Mathematics \bvolume42 \bpages243-270. \endbibitem
  • [15] {barticle}[author] \bauthor\bsnmDürr, \bfnmD\binitsD. and \bauthor\bsnmPulvirenti, \bfnmM\binitsM. (\byear1982). \btitleOn the vortex flow in bounded domains. \bjournalCommunications in Mathematical Physics \bvolume85 \bpages265-273. \endbibitem
  • [16] {barticle}[author] \bauthor\bsnmDuerinckx, \bfnmMitia\binitsM. (\byear2016). \btitleMean-field limits for some Riesz interaction gradient flows. \bjournalSIAM Journal on Mathematical Analysis \bvolume48 \bpages2269-2300. \endbibitem
  • [17] {barticle}[author] \bauthor\bsnmDupuis, \bfnmPaul\binitsP., \bauthor\bsnmLaschos, \bfnmVaios\binitsV. and \bauthor\bsnmRamanan, \bfnmKavita\binitsK. (\byear2020). \btitleLarge deviations for configurations generated by Gibbs distributions with energy functionals consisting of singular interaction and weakly confining potentials. \bjournalElectronic Journal of Probability \bvolume25 \bpages1-41. \endbibitem
  • [18] {bbook}[author] \bauthor\bsnmEthier, \bfnmStewart N\binitsS. N. and \bauthor\bsnmKurtz, \bfnmThomas G\binitsT. G. (\byear2009). \btitleMarkov processes: characterization and convergence \bvolume282. \bpublisherJohn Wiley & Sons. \endbibitem
  • [19] {barticle}[author] \bauthor\bsnmEyink, \bfnmGL\binitsG. and \bauthor\bsnmSpohn, \bfnmH\binitsH. (\byear1993). \btitleNegative-temperature states and large-scale, long-lived vortices in two-dimensional turbulence. \bjournalJournal of Statistical Physics \bvolume70 \bpages833-886. \endbibitem
  • [20] {bbook}[author] \bauthor\bsnmFeng, \bfnmJin\binitsJ. and \bauthor\bsnmKurtz, \bfnmThomas G\binitsT. G. (\byear2006). \btitleLarge deviations for stochastic processes. \bpublisherAmerican Mathematical Soc. \endbibitem
  • [21] {barticle}[author] \bauthor\bsnmFeng, \bfnmJin\binitsJ. and \bauthor\bsnmŚwiech, \bfnmAndrzej\binitsA. (\byear2013). \btitleOptimal control for a mixed flow of Hamiltonian and gradient type in space of probability measures. \bjournalTransactions of the American Mathematical Society \bvolume365 \bpages3987-4039. \endbibitem
  • [22] {barticle}[author] \bauthor\bsnmFernandez, \bfnmBegoña\binitsB. and \bauthor\bsnmMéléard, \bfnmSylvie\binitsS. (\byear1997). \btitleA Hilbertian approach for fluctuations on the McKean-Vlasov model. \bjournalStochastic Processes and their Applications \bvolume71 \bpages33-53. \endbibitem
  • [23] {barticle}[author] \bauthor\bsnmFontbona, \bfnmJ\binitsJ. (\byear2004). \btitleUniqueness for a weak nonlinear evolution equation and large deviations for diffusing particles with electrostatic repulsion. \bjournalStochastic Processes and their Applications \bvolume112 \bpages119-144. \endbibitem
  • [24] {barticle}[author] \bauthor\bsnmFournier, \bfnmNicolas\binitsN., \bauthor\bsnmHauray, \bfnmMaxime\binitsM. and \bauthor\bsnmMischler, \bfnmStéphane\binitsS. (\byear2014). \btitlePropagation of chaos for the 2D viscous vortex model. \bjournalJournal of the European Mathematical Society \bvolume16 \bpages1423-1466. \endbibitem
  • [25] {bbook}[author] \bauthor\bsnmGrafakos, \bfnmLoukas\binitsL. (\byear2014). \btitleClassical Fourier Analysis, \bedition3 ed. \bseriesGraduate Texts in Mathematics. \bpublisherSpringer-Verlag, New York. \endbibitem
  • [26] {barticle}[author] \bauthor\bsnmHardy, \bfnmAdrien\binitsA. (\byear2012). \btitleA note on large deviations for 2D Coulomb gas with weakly confining potential. \bjournalElectronic Communications in Probability \bvolume17 \bpages1-12. \endbibitem
  • [27] {barticle}[author] \bauthor\bsnmHelmholtz, \bfnmH von\binitsH. v. (\byear1867). \btitleLXIII. On Integrals of the hydrodynamical equations, which express vortex-motion. \bjournalThe London, Edinburgh, Dublin Philosophical Magazine Journal of Science \bvolume33 \bpages485-512. \endbibitem
  • [28] {barticle}[author] \bauthor\bsnmJabin, \bfnmPierre-Emmanuel\binitsP.-E. and \bauthor\bsnmWang, \bfnmZhenfu\binitsZ. (\byear2018). \btitleQuantitative estimates of propagation of chaos for stochastic systems with W1,W^{-1,\infty} kernels. \bjournalInventiones mathematicae \bvolume214 \bpages523-591. \endbibitem
  • [29] {binproceedings}[author] \bauthor\bsnmJakubowski, \bfnmAdam\binitsA. \btitleOn the Skorokhod topology. In \bbooktitleAnnales de l’IHP Probabilités et statistiques \bvolume22 \bpages263-285. \endbibitem
  • [30] {barticle}[author] \bauthor\bsnmJoyce, \bfnmGlenn\binitsG. and \bauthor\bsnmMontgomery, \bfnmDavid\binitsD. (\byear1973). \btitleNegative temperature states for the two-dimensional guiding-centre plasma. \bjournalJournal of Plasma Physics \bvolume10 \bpages107-121. \endbibitem
  • [31] {bbook}[author] \bauthor\bsnmKaratzas, \bfnmIoannis\binitsI. and \bauthor\bsnmShreve, \bfnmSteven\binitsS. (\byear1998). \btitleBrownian motion and stochastic calculus, \bedition2 ed. \bseriesGraduate Texts in Mathematics. \bpublisherSpringer, New York. \endbibitem
  • [32] {bbook}[author] \bauthor\bsnmKipnis, \bfnmClaude\binitsC. and \bauthor\bsnmLandim, \bfnmClaudio\binitsC. (\byear1999). \btitleScaling limits of interacting particle systems. \bseriesGrundlehren der mathematischen Wissenschaften. \bpublisherSpringer-Verlag, Berlin, Heidelberg. \endbibitem
  • [33] {barticle}[author] \bauthor\bsnmKipnis, \bfnmC\binitsC., \bauthor\bsnmOlla, \bfnmS\binitsS. and \bauthor\bsnmVaradhan, \bfnmSRS\binitsS. (\byear1989). \btitleHydrodynamics and large deviation for simple exclusion processes. \bjournalCommunications on Pure and Applied Mathematics \bvolume42 \bpages115-137. \endbibitem
  • [34] {bbook}[author] \bauthor\bsnmKirchhoff, \bfnmGustav\binitsG. and \bauthor\bsnmHensel, \bfnmKurt\binitsK. (\byear1883). \btitleVorlesungen über mathematische Physik \bvolume1. \bpublisherDruck und Verlag von BG Teubner. \endbibitem
  • [35] {barticle}[author] \bauthor\bsnmLeray, \bfnmJean\binitsJ. (\byear1933). \btitleÉtude de diverses équations intégrales non linéaires et de quelques problèmes que pose l’hydrodynamique. \bjournalJournal de Mathématiques Pures et Appliquées \bvolume12 \bpages1-82. \endbibitem
  • [36] {barticle}[author] \bauthor\bsnmLiu, \bfnmWei\binitsW. and \bauthor\bsnmWu, \bfnmLiming\binitsL. (\byear2020). \btitleLarge deviations for empirical measures of mean-field Gibbs measures. \bjournalStochastic Processes and their Applications \bvolume130 \bpages503-520. \endbibitem
  • [37] {barticle}[author] \bauthor\bsnmMarchioro, \bfnmC\binitsC. and \bauthor\bsnmPulvirenti, \bfnmM\binitsM. (\byear1982). \btitleHydrodynamics in two dimensions and vortex theory. \bjournalCommunications in Mathematical Physics \bvolume84 \bpages483-503. \endbibitem
  • [38] {barticle}[author] \bauthor\bsnmMcKean, \bfnmHenry P\binitsH. P. (\byear1967). \btitlePropagation of chaos for a class of non-linear parabolic equations. \bjournalStochastic Differential Equations \bpages41-57. \endbibitem
  • [39] {barticle}[author] \bauthor\bsnmMéléard, \bfnmSylvie\binitsS. (\byear2000). \btitleA trajectorial proof of the vortex method for the two-dimensional Navier-Stokes equation. \bjournalAnnals of Applied Probability \bvolume10 \bpages1197-1211. \endbibitem
  • [40] {bbook}[author] \bauthor\bsnmOlivieri, \bfnmEnzo\binitsE. and \bauthor\bsnmVares, \bfnmMaria Eulália\binitsM. E. (\byear2005). \btitleLarge deviations and metastability \bvolume100. \bpublisherCambridge University Press. \endbibitem
  • [41] {barticle}[author] \bauthor\bsnmOnsager, \bfnmLars\binitsL. (\byear1949). \btitleStatistical hydrodynamics. \bjournalIl Nuovo Cimento (1943-1954) \bvolume6 \bpages279-287. \endbibitem
  • [42] {barticle}[author] \bauthor\bsnmOsada, \bfnmHirofumi\binitsH. (\byear1985). \btitleA stochastic differential equation arising from the vortex problem. \bjournalProceedings of the Japan Academy, Series A, Mathematical Sciences \bvolume61 \bpages333-336. \endbibitem
  • [43] {barticle}[author] \bauthor\bsnmOsada, \bfnmHirofumi\binitsH. (\byear1986). \btitlePropagation of chaos for the two dimensional Navier-Stokes equation. \bjournalProceedings of the Japan Academy, Series A, Mathematical Sciences \bvolume62 \bpages8-11. \endbibitem
  • [44] {barticle}[author] \bauthor\bsnmQuastel, \bfnmJeremy\binitsJ. (\byear1995). \btitleLarge deviations from a hydrodynamic scaling limit for a nongradient system. \bjournalThe Annals of Probability \bpages724-742. \endbibitem
  • [45] {barticle}[author] \bauthor\bsnmQuastel, \bfnmJeremy\binitsJ., \bauthor\bsnmRezakhanlou, \bfnmF\binitsF. and \bauthor\bsnmVaradhan, \bfnmSRS\binitsS. (\byear1999). \btitleLarge deviations for the symmetric simple exclusion process in dimensions d3d\geq 3. \bjournalProbability Theory and Related Fields \bvolume113 \bpages1-84. \endbibitem
  • [46] {barticle}[author] \bauthor\bsnmSchochet, \bfnmSteven\binitsS. (\byear1996). \btitleThe point‐vortex method for periodic weak solutions of the 2‐D Euler equations. \bjournalCommunications on Pure and Applied Mathematics \bvolume49 \bpages911-965. \endbibitem
  • [47] {barticle}[author] \bauthor\bsnmSerfaty, \bfnmSylvia\binitsS. (\byear2020). \btitleMean field limit for Coulomb-type flows. \bjournalDuke Mathematical Journal \bvolume169 \bpages2887-2935. \endbibitem
  • [48] {barticle}[author] \bauthor\bsnmTakanobu, \bfnmSatoshi\binitsS. (\byear1985). \btitleOn the existence and uniqueness of SDE describing an nn-particle system interacting via a singular potential. \bjournalProceedings of the Japan Academy, Series A, Mathematical Sciences \bvolume61 \bpages287-290. \endbibitem
  • [49] {binbook}[author] \bauthor\bsnmTanaka, \bfnmHiroshi\binitsH. (\byear1984). \btitleLimit theorems for certain diffusion processes with interaction In \bbooktitleNorth-Holland Mathematical Library \bvolume32 \bpages469-488. \bpublisherElsevier. \endbibitem
  • [50] {barticle}[author] \bauthor\bsnmTanaka, \bfnmHiroshi\binitsH. and \bauthor\bsnmHitsuda, \bfnmMasuyuki\binitsM. (\byear1981). \btitleCentral limit theorem for a simple diffusion model of interacting particles. \bjournalHiroshima Mathematical Journal \bvolume11 \bpages415-423. \endbibitem
  • [51] {bbook}[author] \bauthor\bsnmTemam, \bfnmRoger\binitsR. (\byear1995). \btitleNavier–Stokes equations and nonlinear functional analysis. \bpublisherSIAM. \endbibitem
  • [52] {bbook}[author] \bauthor\bsnmVillani, \bfnmCédric\binitsC. (\byear2003). \btitleTopics in optimal transportation. \bpublisherAmerican Mathematical Soc. \endbibitem
  • [53] {barticle}[author] \bauthor\bsnmWang, \bfnmZhenfu\binitsZ., \bauthor\bsnmZhao, \bfnmXianliang\binitsX. and \bauthor\bsnmZhu, \bfnmRongchan\binitsR. (\byear2021). \btitleGaussian fluctuations for interacting particle systems with singular kernels. \bjournalarXiv preprint arXiv:2105.13201. \endbibitem