Sample-path large deviation principle for a 2-D stochastic interacting vortex dynamics with singular kernel
Abstract
We consider a stochastic interacting vortex system of 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 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 norms along sample paths.
keywords:
[class=MSC]keywords:
,
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
Let be the two dimensional torus, i.e. the unit cube with metric , and be the Green function on , i.e.
(1.1) |
in which means Dirac measure with singularity at . We define a particle stochastic vortex dynamics by stochastic differential equations on 2D torus for any given :
(1.2) |
with random initial data , independent of the family of 2D-Brownian motions, where the Biot-Savart kernel is given by . Supposing the initial energy, i.e.
is finite, we are interested in the sample-path LDP of the empirical measure process when goes to .
The model without noise, i.e. , 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 is exactly a Hamiltonian system with the singular potential 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 of the corresponding model in converges to the vorticity form of two dimensional Navier-Stokes equations as .
Consider the incompressible Navier-Stokes equations on :
(1.3) | ||||
Its vorticity satisfies a continuity equation
(1.4) |
with a velocity field driven by the vorticity itself , in which for the case of . (1.4) is called the voricity form of two-dimensional Navier-Stokes equations.
Osada [42, 43] investigated the well posedness of (1.2) in and obtained a propagation of chaos for the equation (1.4) for large . For general positive , using the cutoff kernels 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 large and without cutoff when the distributions of initial data are regular enough.
Back to , the well posedness for this SDEs with the singular kernel is firstly studied in [15]. In this case, as a classical result, are smooth even function on and
where is a bounded correction to periodize on . In our paper, we need a refined result that
(1.5) | ||||
for certain smooth function satisfying
and . The method of [24] also works and the propagation of chaos holds for (1.4) on . The slight difference is that instead of in the case of in which is the solution of Navier Stokes equation (1.3) on .
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 .
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 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 norm over time, which is our key observation and main tools throughout this paper.
1.2 Notations and definitions
We write and for the collection of infinitely differentiable functions on and generalized functions on .
We use to represent pairing of linear space and its dual space, such as inner product, continuous function integral with respect to a measure, and so on.
Denote by the space of probability measures on . For we define the Wasserstein metric
(1.6) |
where
Then is a complete separable metric space. See Chapter 7 of [1] or Chapter 7 of [52] for properties of this metric. Given , for any we define
(1.7) |
Let be the set of all satisfying that there exists such that
We refer to smooth mollifier throughout this paper as a non-negative even function that is smooth and its integral over is one.
1.3 Statement of the main result
We introduce the state space
As shown by [15], for the case , the system (1.2) is well-posed for almost every , 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 as -valued random variables, independent of the family of 2D-Brownian motions. Suppose that , then there exists an unique strong solution to (2.1) and
Now we introduce the energy functional for empirical measures
where is the diagonal set. We impose the following superexponential finite energy condition for initial data.
Condition 1.1.
The removal of diagonal set is not convenient for estimations, so we also introduce the energy functional containing the diagonal set,
It’s always well defined since is bounded below. By Fatou’s Lemma, is lower semi-continuous functional under the weak topology, and we left more properties of into Appendix C.
Let be defined by
and for and , we define
To check is well defined, by Theorem 8.3.1 of [1], for , we can define for almost every . In addition, implies that for almost every . As a consequence of (1.5) and Young’s inequality, if , then , so
is well defined for any smooth function . Thus for almost every .
Our main result is
Theorem 1.2 (Large deviation principle).
Under the condition of Lemma 1.1, suppose , as random variables on , satisfies a large deviation principle with the rate function and Condition 1.1 holds. Then for each , the stochastic empirical process satisfies a sample-path large deviation principle in the space with a good rate function (usually called action)
(1.8) |
that is equivalent to the following two conditions:
1. (Large deviation upper bound) For any closed set in ,
2. (Large deviation lower bound) For any open set in ,
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 as the modification, which implies finite energy and finite integral of norm of the distributions over time.
Besides the modification by , the proof for the lower bound still needs several new inequalities, which can bound the singular terms such as by the integral of 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 norms along sample path naturally emerges. We can not expect the norms to be bounded from above since the initial norm can be infinite, but the integral of 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 .
As stated in [5], the rate function with modification makes it harder to prove the upper bound, requiring a prior estimation of . However, always takes on the empirical measure , even if their limit as has finite . Thus we make use of the convolution method with a sequence of smooth mollifiers. By taking specific mollifiers converging to Dirac measure with a suitable slow rate, we yield an energy dissipation structure for assuming is finite, which gives the control for . 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
where the constant is to make . is a non-negative smooth function on with a compact support. Then let be the specific smooth mollifiers generated by and defined by
where . Note that the slow convergent rate of 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 that we will show later.
Condition 1.2.
Theorem 1.3 (Large deviation principle II).
To conclude the proof, we will also show that finite is stronger than finite , 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 under convolution and give the prior estimation of 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 are determined by the stochastic differential equations
(2.1) |
By Ito’s formula, for each smooth function on , we have
To write the generator of , we need to write the singular term as a 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 is the diagonal set, noting
(2.2) | ||||
is only dependent on and and linear of In view of (1.5), is a bounded function so that for each ,
(2.3) |
is bounded. Hence the final expression of (2.2) is always well defined and can be extended to any . Now we introduce the symmetrization operator satisfying
(2.4) |
As stated above, (2.4) is always well-defined and
(2.5) |
In fact is an extension for If as stated before Theorem 1.2, . Then for each ,
Hence
(2.6) |
With this notation, after further calculations, we can write the generator for as
(2.7) |
for each such that
Then we define
(2.8) | ||||
in which as well as is defined in distribution sense and the variational derivative
is a smooth function on .
At least formally by Ito’s formula, for such that , the exponential form
is always considered to be a martingale. In fact, the proof of [33, 14] only use the case when and To adapt their proof to the more general initial data, we shall use a bigger domain , 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 , 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 there exists a compact set such that
(2.9) |
Proof.
Formally, there are three steps to prove the large deviation upper bound for compact sets. First, by Ito’s formula, for each , and ,
is a non-negative martingale, hence for each measurable set ,
(2.11) | ||||
where is the indicator function and
(2.12) | ||||
Second, obtain the limit of the right-hand-side of (2.11) when goes to infinity (write it as ). Taking the supremum over , and then we have
Finally, use Lemma in [32] to exchange the supremum and infimum to obtain the upper bound for any compact set , i.e.
However, the exchange of the supremum and infimum requires that is lower semi-continuous under the topology of . Hence we do need the singular term is continuous under weak topology. However, that is not true. Fortunately, is continuous under weak topology with finite energy, stated in the following lemma.
Lemma 2.2.
If , and (or ), then for each such that and , we have
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,
(2.13) |
Proposition 2.4 (Upper bound for compact sets).
Proof.
Take in Proposition 2.1. For any measurable set , we define
By Lemma 2.3, as
(2.15) |
In view of (2.11),
On the one hand, for , define and it’s easy to obtain that for certain constant . On the other hand, by Varadhan’s lemma,
exists. Hence,
(2.16) |
where
Now we want to prove that
(2.17) | ||||
It is trivial if there exists such that for . Otherwise, pick approximating the infimum in (2.16) for each satisfying , i.e.
Since and is lower semi-continuous, we can find , a limiting point of , with . Then by Lemma 2.2 and dominated convergence theorem,
which combining with (2.16) implies
(2.18) | ||||
Use the fact is continuous when , 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 ,
Taking , we obtain
(2.19) | ||||
where the expression inside infimum is lower semi-continuous with respect to . Taking the infimum with respect to and in the RHS of (2.19) over , 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 , by (2.15), we have
(2.20) | ||||
With the prior estimation , we can obtain the upper bound is, as shown in the lemma below, actually equal to in (1.8).
Lemma 2.5 (Variational representation of the rate function).
If such that and , then . Conversely, if with , then =
Finally, we obtain the large deviation upper bound.
Proposition 2.6 (Upper bound).
Under the condition of Theorem 1.3, for each closed set in
(2.24) |
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 , take
By Proposition 5.12 of [31], it’s a martingale. Thus we can apply Girsanov formula. Let be the augmented filtration given by initial data and the Brownian motion (see (2.3) in Section 5.2 of [31]). For , let Let and then is a -dimensional Brownian motion under . In addition, we have
(2.25) |
Formally, the empirical distribution of (2.25) would converge to the solution of
(2.26) |
Hence a definition of weak solution is required for characterizing the mean-field limit.
Definition 2.1.
For , we say is a weak solution of (2.26) if and for each and ,
Then we will prove the law of large numbers under .
Lemma 2.7.
Let with and . Then there exists an unique weak solution of (2.26) such that and
where is a -dependent constant. In addition, for each , if a sequence satisfies , and , then
Remark.
It might happen that for small , but it would not happen when big enough (see Lemma 5.3). Hence, we allow to be defined only for large .
Generally speaking, this result would imply that if the initial data converging to in weak topology, then the limit of relative entropy of with respect to under the scaling would provide a large deviation lower bound, i.e.
for any open set containing , in which
Moreover, at least formally, by taking the infimum for we will obtain exactly the rate function, i.e.
(2.27) |
Define
In our paper, for regular enough, i.e. for certain , the infimum in (2.27) is always taken in the case that .
Since our initial data is more general, we need a stronger result stated below involving the stability for initial values.
Lemma 2.8.
Let with and . For any there exists such that for each sequence satisfying and ,
(2.28) |
where In addition, if , then there exists a constant dependent on such that
(2.29) |
As one may expect, the set of ”nice” trajectories, consisting of all for , plays a key role in proof of lower bound of LDP. We want to prove each with finite rate function can be approximated by a series of in the sense stated below.
Lemma 2.9 (Density of nice trajectory).
For each with , define If with , there exists such that
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 in , which is stated below.
Proposition 2.10 (Lower bound).
Under the condition of Theorem 1.3, for any and ,
Proof.
If , the result is trivial. Otherwise, and . By Condition 1.2, we can pick big enough such that
(2.30) |
By LDP of initial data, for each ,
which in combination with (2.30) leads to
(2.31) |
Let and assume By Lemma 2.9, we can select such that and
(2.32) |
We claim that for each ,
(2.33) |
Then by (2.32) we have
and arrive at the lower bound.
So we just need to verify (2.33). For any small enough such that , take in Lemma 2.8 such that (2.28) and (2.29) holds replacing by . Note that
(2.34) | ||||
By (2.31), there exists such that when for . Hence for , take such that and
(2.35) |
In view of (2.31), (2.34) and (2.35),
(2.36) | ||||
Since ,
By Jensen’s inequality,
By Lemma 2.8, the second expression on the RHS of the above inequality converges to . The first one is equal to
(2.37) |
3 Prior energy estimation and crucial inequalities
This section investigates some estimations related to the energy functional along with . 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
and
(3.1) |
where . For convenience, we take . In fact, the estimations in this section hold for each supported on and satisfying the Condition 3.1 below.
Condition 3.1.
(1) is a non-negative smooth even function.
(2) For each .
(3)
(4) There exists a constant such that
Define , which also are smooth mollifiers. It is straightforward to show that for certain function satisfying Condition 3.1 and being supported on , and
(3.2) |
3.2 Convolution estimations
Lemma 3.2.
There exists a constant such that for ,
Proof.
Lemma 3.3.
For any , there exists a sequence such that for each , satisfying ,
Proof.
Step 1.
First, we firstly prove that for each , there exists a constant not dependent on and such that for ,
(3.8) | ||||
Noticing that if and
together with (2.4), (3.7) and Lemma 3.2, it’s enough to show there exists a constant such that for each ,
(3.9) |
Since can be seen as a periodic function, without loss of generality, we assume , so that . Let . Then
(3.10) | ||||
where we used (3.6) to obtain
As mentioned in [28], , i.e. there exist such that . Therefore,
So
Since , for we have
and thus by (1.5),
Therefore (3.10) can deduce (3.9) and we arrive at (3.8).
Step 2.
Take in (3.8). Noting that , for , we have
By (C.9),
Taking
then
Through a simple calculation, we can check
∎
Remark.
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 such that for each ,
Proof.
Let be a radial function such that for and for We define for , and We start with proving that for there exist constants , such that
(3.11) |
By (1.5), we can find such that for each
(3.12) |
So there exists such that
Also we have
Let . Then we have
Turn to the proof of the desired inequalities. A consequence of (3.11) along with Young’s inequality implies that there exists a constant , such that
and
If , we take and have
If , by Young’s inequality,
in which due to (1.5). So we can pick such that
To obtain the second inequality, noting that and by young’s inequality, we have
Hence by Holder’s inequality,
∎
Corollary 3.5.
For each , , smooth mollifier and measurable function satisfying , one has
where is the constant in Lemma 3.4.
Proof.
3.3 Proof of Lemma 2.3
We need a generalization of the Doob submartingale inequality.
Lemma 3.6.
If is a positive continuous local martingale, then for each
Proof.
Since is a positive local martingale, it’s a supermartingale. Let . Then is a non-negative supermartingale. Hence
∎
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 such that for each sequence satisfying and
(3.13) | ||||
for any .
Proof.
Define , . By Ito’s formula and using the fact , we have
After further calculation, for any
(3.14) | ||||
is a positive continuous martingale. By Lemma 3.6, for each with
(3.15) | ||||
By Jensen inequality
Then by Lemma 3.4, there exists such that
(3.16) |
Combining (3.16) with (3.15) and taking such that , we have
For each , define as a subset of by
By Lemma 3.3, there exists a sequence only dependent on such that if , then
(3.17) |
Take big enough such that and . Define
We claim that if and , then .
To prove it by contradiction, suppose . Since ,
(3.18) |
Noting , we have
(3.19) |
However, (3.18), (3.17) and (3.19) imply
which is a contradiction. Hence, for each with , there exists such that if then
Therefore,
Recall so . By the arbitrariness of 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 For and , define the norm
Define as the completion of under That is a Hilbert space with inner product
Then for any , there exists a function defined on such that
and
The inner product thus can be written as
Recall that
We claim that means can be seen as a bounded linear operator on . To obtain that, for define
By the definition of , for a test function , taking , we have
implying is a bounded operator on .
By Riesz representation theorem, if , there exists such that
and , i.e. for each
(4.1) | ||||
In addition, by (4.1), taking smooth test function approximating in the definition of , we finally have
(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.
Proof.
Since , as discussed in the previous subsection, there exists satisfying (4.2) and (4.1) holds for each . Since , it’s standard to see (4.3) holds for each by smooth approximation of truncated functions only supported on .
Take a smooth mollifier with compact support and define . Taking in (4.3), noting for almost every and by (2.6), we can check that is absolutely continuous with respect to and for a.e.-
(4.5) |
where .
Now we show is absolutely continuous with respect to . Notice that and are bounded by constants which only depend on , then for
Then by directly taking derivative and dominated convergence theorem, we have
By (4.5) and noticing as well as
(4.6) | ||||
To use (4.6) to show (4.4), we just need to prove that
(4.7) |
(4.8) | ||||
and the convergence of rest terms can be proved by Fatou’s Lemma and Jensen’s inequality.
Note that for a.e. since . By Lemma 3.4, for a.e. , so that by property of smooth mollifiers,
(4.9) |
By Holder’s inequality, for a.e. ,
(4.10) | ||||
where we used the fact that So far we have proved that
Apply Lemma 3.4 along with Jensen’s inequality to (4.10),
Since , by dominated convergence theorem, (4.7) follows.
Turing to (4.8), by Holder’s inequality,
(4.11) | ||||
The condition along with (4.2) implies for a.e. . Thus by (4.9) the point-wise convergence of holds. Finally, by Lemma 3.4 and Jensen’s inequality, the RHS of (4.11) is bounded by which is integrable. By dominated convergence theorem, we reach (4.8). ∎
Let the entropy functional be defined by
and the Fisher information functional be defined by
Lemma 4.2.
Proof.
We start with proving (4.12). Still using the smooth mollifiers in Lemma 4.1, let . By (4.5),
(4.14) | ||||
Due to mean value inequality,
(4.15) | ||||
where we used Lemma 8.1.10 of [1] to obtain the last inequality. Since , we have and
(4.16) | ||||
By Jensen’s inequality
(4.17) |
Applying (4.15), (4.16) (4.17) to (4.14), we have
Let
By Gronwall inequality
Note that
(4.18) | ||||
From Lemma 8.1.10 of [1] combined with Lemma 3.4 and Jensen’s inequality [1], (4.18) can be controlled by . In addition, for satisfying by Fatou’s Lemma and Lemma 8.1.10 of [1], the first term converge to . By property of mollifiers, the other two terms also converge to . Hence (4.18) converge to for a.e . Taking , by dominated convergence theorem we have and by Fatou’s lemma we arrive at
4.3 Proof of Lemma 2.5
Proof.
Suppose , and . Take in Lemma 4.1. Let and then in distribution sense,
By Lemma 8.3.1 of [1], to show , we just need to prove
(4.19) |
By Lemma A.1, Proposition A.2 and Lemma 3.4,
Then (4.19) holds by Holders inequality and Lemma 4.2.
To prove the second conclusion, note that if , by Lemma A.1,
By Lemma D.34 of [20], for almost every , there exists such that
and
By definition of and Cauchy-Schwarz inequality,
∎
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 .
Lemma 5.1.
Given with . Suppose and are two weak solutions of (2.26) for and . Then
Proof.
Write Take smooth mollifiers 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
(5.1) | ||||
By Lemma 3.4 and Jensen’s inequality,
Similarly,
Since and are finite, applying dominated convergence theorem to (5.1), we have
(5.2) | ||||
Noticing that we have
Since writing ,
Hence,
(5.3) | ||||
Apply Gagliardo–Nirenberg interpolation inequality to (5.3),
(5.4) | ||||
where we used the fact
On the other hand, by Cauchy–Schwarz inequality and (C.5),
(5.5) | ||||
Combining (5.2) with (5.4) and (5.5), we have
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 , and only dependent on such that for each sequence satisfying and
(5.6) | ||||
Proof.
Recall By Ito’s formula, as calculation in Lemma 3.7, for any
(5.7) | ||||
is a positive continuous martingale. By Lemma 3.6, for each with
By Corollary 3.5, there exists dependent on and , such that
Combining this inequality with (3.16) and taking such that , we have
The rest is similar to the proof of Lemma 3.7. ∎
Then we verify that there exists a sequence of satisfying the conditions of Lemma 5.2.
Lemma 5.3.
For with , there exists such that
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 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
First we consider the case that Recall and write
In view of (2.4), we just need to show
By Tietze extension theorem, we can construct a continuous function such that
for and
By Lemma C.3 for , there exists such that for ,
Similarly,
By lower semi-continuity of in , . Hence we can take small enough such that
(5.10) | ||||
Since is a continuous function, and in narrow topology,
(5.11) |
Combining (5.10) and (5.11), we have
Let and we conclude the proof for the case .
For the general case, let
Note that
Take , then we have
By Lemma C.3, take and we complete the proof. ∎
Proof of Lemma 2.7.
Given and and By Ito’s formula, for ,
(5.12) | ||||
Note that by (2.5)
(5.13) |
Then by [29] and Theorems 8.6 and 8.8 in Chapter 3 of [18], we conclude that is tight in .
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 . By the Skorohod representation theorem, we can construct a canonical probability space (write for the corresponding probability measures) on which random variables are defined, with the property that has the same law as for such that
(5.14) |
By Lemma 5.2 and Borel-Cantelli Lemma,
Then by the lower semi-continuities of under the topology of ,
(5.15) |
By Lemma 2.2,
By (5.12) and dominated convergence theorem, for each
(5.16) | ||||
and
(5.17) | ||||
By Fatou’s Lemma and (5.12),
(5.18) | ||||
Since , we have . Then by (5.15), (5.17), (5.16) and (5.18), is almost surely a weak solution of (2.26) for in Definition 2.1.
Proof of Lemma 2.8.
We start with proving that there exists such that for any and ,
(5.19) |
To prove it by contradiction, suppose that it’s not true. Then, there exists so that for any we can find and such that
Taking subsequence if necessary, assume , where depend on . Then by Lemma 2.7,
Hence we obtain
(5.20) |
For , we write satisfying (5.20). Since we have the estimation
with a similar argument as the proof of Lemma 2.7,
and is a weak solution of (2.26). By Lemma 5.1, . However, it contradicts (5.20), so we have proved (5.19).
5.4 Proof of Lemma 2.9
Lemma 5.4.
For each , is smooth on and
Proof.
Let . When , it’s a classical result that the solution of (2.26) can be obtained by (1.3) on . implies and . Actually, implies 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 as well as on , to prove this Lemma.
By Lemma 4.2, for each , take such that and
Similar with Lemma 3.2 of [24], for
Hence
Then by Theorem 2.5 of [24],
This meets the assumptions of the theorem of [7] (which improves Theorem B of [3]), so is a smooth classical solution on .
To show , note that for can be see as a solution of second-order parabolic equation
Since we have proved is smooth on and , we conclude the proof by strong maximum principle of parabolic equation. ∎
Lemma 5.5.
(5.21) |
There exists a constant , such that for each smooth mollifier ,
(5.22) |
Proof.
Proof of Lemma 2.9.
We split the proof into three steps.
Step 1. Construction.
Write as the heat kernel on , defined in Appendix B. Given and , define
One can easily check that . Writing , we construct below corresponding to such that
(5.23) |
For , take and (5.23) holds. Since we only care the weak solution, we just define , which would not bring any problems.
By Lemma 5.4, is smooth and , so for , is uniform elliptic. Hence (5.23) can be seen as a second-order elliptic equation for . We expect all the coefficients in (5.23) as an equation for is regular (i.e. their derivatives of arbitrary order are uniform bounded), which ensures an unique weak solution such that is uniform bounded. However, may not be regular when is approaching . But we notice that is the heat kernel, which implies
Then (5.23) reduces to
(5.24) |
Pointwisely for , take as the weak solution of (5.24). Due to smoothness of , all the coefficients in (5.24) is regular, so is uniform bounded and for , (5.23) holds.
For because of the convolution by , is uniform elliptic. Regard (5.23) as a second-order elliptic equations for and take as its weak solution. For ,
(5.25) | ||||
which is regular, and by (4.5), is also regular in , so is uniform bounded when .
Therefore (5.23) holds for the constructed pairs and is uniform bounded. As a consequence, and is a weak solution of (2.26) for
Step 2. Verify that the limit of , when taking first and then letting , is bounded above by
Without loss of generality, we assume
Write . By definition,
(5.26) | ||||
For , we have
By Proportion A.2 and Lemma 3.4,
Since the result of Lemma 5.5 also holds for the solution of heat equation,
Noting is continuous under weak topology, by Fatou’s Lemma and continuity of (as a consequence of Lemma 5.5),
By (5.25),
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,
which is, in view of Lemma 5.5, bounded by
Also by Fatou’s Lemma,
By Lemma A.3, for each
Combining these estimations, we have
Note that implies so by Lemma 3.4 and dominated convergence theorem, the second term vanishes when . The limit of third term is also non-positive by Fatou’s Lemma. Thus
Due to the arbitrariness of we conclude that
Step 3. Convergence.
It’s easy to check for each , To conclude the proof, we need strength this point-wise convergence into
which can be obtained by the compactness result below. ∎
Lemma 5.6.
Appendix A Weighted Sobolev space
In this section, we give some property of for . Recall that
Lemma A.1.
Suppose then
Proof.
By definition of and the inequality ,
∎
Recall
and
Proposition A.2.
For
Proof.
Lemma A.3.
Let be an arbitrary smooth mollifier. Then for each
Proof.
By Jensen’s inequality, for each ,
Using
Hence,
Taking the supremum for we conclude the result. ∎
Appendix B Heat kernel on torus
We define the heat kernel on ,
(B.1) |
for
Theorem B.1.
(1) For any
is a well defined smooth mollifier.
(2)
(3) For and , satisfies the heat equation:
(B.2) |
and .
Proof.
(1) Based on the theory of uniformly convergent series, is a well defined smooth function. Since the integral of every term in (B.1) is zero except the case , By Poisson summation formula (e.g. see Theorem 3.1.17 of [25]),
(B.3) |
Hence We conclude that is a smooth mollifier.
(2) It follows by direct calculation.
(3) Since , for each . By (B.3),
Hence is continuous under weak topology at . By (2), we have .
For , since each item in the sum of (B.1) for satisfies (B.2), also satisfies (B.2), so does .
∎
Appendix C Property of energy functional
By (1.5), is bounded from below, so
is well-defined for any probability measure . By Young’s inequality, for any finite signed measure on , , is in .
Lemma C.1.
Suppose are probability measures. we have
(1) has a weak derivative and
(C.1) |
(C.2) |
(2) If is a smooth mollifier,
(C.3) |
Proof.
We start with (1). (C.1) is standard and follows from Fubini’s lemma and integration
by parts against test functions 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),
Noting that , by Fubini’s lemma and integration by parts
∎
Proposition C.2.
If and , then
(C.4) |
(C.5) |
In particular, If , then
(C.6) |
where we admit if
Proof.
We first prove (C.4).
At the end of this section, we give some estimates for with
Lemma C.3.
There exists a constant , such that for each with and ,
(C.8) |
(C.9) |
In addition, given a smooth mollifier and a non-negative sequence , let . Then for each there exists such that for each and ,
(C.10) |
Proof.
[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 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 . \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 -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