Concentration bounds for stochastic systems with singular kernels
Abstract.
This note is concerned with weakly interacting stochastic particle systems with possibly singular pairwise interactions. In this setting, we observe a connection between entropic propagation of chaos and exponential concentration bounds for the empirical measure of the system. In particular, we establish a variational upper bound for the probability of a certain rare event, and then use this upper bound to show that “controlled” entropic propagation of chaos implies an exponential concentration bound for the empirical measure. This connection allows us to infer concentration bounds for a class of singular stochastic systems through a simple adaptation of the arguments developed in [JW18].
1. Introduction
1.1. Stochastic particle systems
Let and be a vector field defined on the dimensional flat torus . We consider the system of particles described by the dynamics
(1.1) |
where the are independent standard Wiener processes defined on a standard filtered probability space . We work on a finite time horizon , and we assume that the initial positions of the particles are i.i.d., i.e.
Thus the data used to define our particle system consists of the time horizon , the maps , and the initial distribution . Under these conditions, the law 111Here and throughout the note we use the notation to indicate a curve , and we abuse notation by writing for the density of when it exists. of the particles is formally described by the Liouiville equation
(1.2) |
where . When and are smooth (or at least Lipschitz), it is well known that the asymptotic behaviour of the particle system (1.1) is described by the non-local Fokker-Planck equation
(1.3) |
More precisely, it is known that in an appropriate sense we have
The first statement above confirms that is the mean field limit of the empirical measures , while the second statement is referred to as propagation of chaos. Quantitative versions of these statements, as well as finer results like central limit theorems, large deviations principles, and concentration bounds for the empirical measures have also been obtained for regular . We do not make any attempt to summarize this literature, but refer to survey articles like [JW17, CD22] for an introduction. Extending such results to more singular kernels , which are often met in applications, is a very active area of research. We mention in particular the recent flurry of activity around singular kernels which derive from Riesz potentials (see e.g. [Ser20, BJW20, RS23, CdCRS23] and the references therein) and some recent efforts aimed at singular attractive kernels (see e.g. [BJW23, CdCRS23]). More relevant to the present note are the slightly less recent contributions of [JW16] and [JW18], where quantitative propagation of chaos was established by estimating the relative entropy between a solution of the Liouiville equation (1.2) and the tensor product of the solution of (1.3). More precisely, [JW18] established the following quantitative version of “entropic propagation of chaos”:
(1.4) |
for some constant independent of , where denotes the rescaled relative entropy
We note that (1.4) is established under a general initial condition for (1.2), but if then the first term on the right-hand side of (1.4) vanishes, leading to , which by a classical subadditivity property of relative entropy gives , where denotes the -particle marginal of , and in particular as for each fixed , which is what we mean by entropic propagation of chaos. We note also that establishing (1.4) is not the only way to obtain quantitative entropic propagation of chaos - we refer to [Lac23] for a more detailed discussion and for an approach which leads to optimal bounds on .
1.2. Our results
Our goal in this note is to point out a connection between entropic propagation of chaos and concentration inequalities for the empirical measure of the particle system (1.1). In particular, for , we are interested in bounds of the form
(1.5) |
where denotes the -Wasserstein distance. That is, we want to show that the empirical measure for the particle system admits concentration bounds of the same type available for i.i.d. samples, as established in [FG15]. When written in terms of the Liouville equations (which is often necessary for technical reasons when is singular), (1.5) becomes
(1.6) |
Such concentration inequalities were obtained for regular interactions by several authors. We highlight in particular [BGV05], where concentration inequalities for the system (1.1) were inferred from concentration inequalities for i.i.d. random variables via the so-called “synchronous coupling” method originally due to McKean [McK69] and popularized by Sznitman [Szn91], and [DLR18], where a more general concentration of measure result was obtained by exploiting certain (uniform in ) functional inequalities for the law of . When is singular, the approaches in [BGV05] and [DLR18] break down, and the only estimate similar to (1.5) we are aware of is Proposition 5.3 of [HC23], where the structure of the repulsive Riesz interactions is leveraged to get bounds somewhat similar to (1.5).
To explain our main idea, we must introduce the “controlled Liouville equation”:
(1.7) |
where is a measurable map which we view as a control. For the so-called kernels of [JW18], one can (as is explained in more detail in the proof of Proposition 1.6) easily generalize the original entropy estimate (1.4) of Jabin and Wang to get an estimate of the form222The factor of in (1.8) is purely aesthetic, and is included to make (1.8) more consistent with Proposition 3.1 below.
(1.8) |
for any solution to (1.7). We emphasize that here may be different from . We think of (1.8) as a “controlled” or “perturbed” version of the entropic propagation of chaos estimate (1.4) of Jabin and Wang. Our main observation is that an estimate of the form (1.8) in fact implies a concentration bound of the form (1.5). For technical reasons, we first make this observation precise through the following a-priori estimate, which requires and to be bounded.
Theorem 1.1.
There are constants depending only on and such that the following holds: Suppose that and are bounded, the initial condition has a bounded density, and that there exists a solution to (1.3) and a constant such that (1.8) holds for all and such that is an entropy solution of (1.7) as in Definition 2.2 below. Then the concentration bound (1.5) holds for each , with .
Remark 1.2.
Remark 1.3.
While is required to be bounded for technical reasons, the key point of the a-priori estimate is that depends on only through the constant appearing in (1.8), and not on . Similarly, there is no explicit dependence of our a-priori estimate on , so if one establishes (1.4) with a constant independent of (which can be done for the 2-D stochastic vortex model with the techniques of [GBM24]) then one gets a uniform in time concentration bound.
Remark 1.4.
As mentioned above, we can verify that (1.8) holds under roughly the same conditions as in [JW18]. Thus we make use of the following assumption (and refer to the notation section below for the definition of ):
Assumption 1.5.
For some , , we have , , , , and
Proposition 1.6.
Let Assumption 1.5 hold. Then there is a unique classical solution to (1.3), and for each there exists an admissible entropy solution to (1.2) in the sense of Definition 2.4. Moreover, for each there is a constant which depends only on , , and , such that any admissible entropy solution of (1.2) satisfies (1.6).
Remark 1.7.
Regarding possible extensions: one can easily extend Theorem 1.1 by replacing by , but the concentration bound will then depend on the (exponential) moments of the limit . We focus on the periodic case because in this case we can adapt the arguments of [JW18] to obtain (1.4). However, the recent preprint [FW23] shows that the program of [FW23] can be carried out in the whole space for the 2-D viscous vortex model, so the (non-periodic analogue of) our Theorem 1.1 can be combined with the argument in [FW23] to obtain an anologue of Proposition 1.6 in this setting. Likewise, one can easily adapt Theorem 1.1 to the kinetic setting, in which case the “controlled Liouiville equation” will involve a control only in the velocity (as is easily seen from an inspection of the proof of Lemma 3.1, in particular the application of Girsanov’s Theorem). By combining such a result with the techniques of [JW16], one can easily derive an anologue of Proposition 1.6 for kinetic particle systems with bounded forces, i.e. in the same setting as [JW16].
2. Notation and Preliminaries
2.1. Notation and some definitions
Throughout the paper is a positive real number and is the dimensional flat torus. is the convolution between the functions . If a function is sufficiently regular, then we denote by the partial derivative with respect to and the partial derivative with respect to (the -th coordinate) for , respectively. For , we use the symbol (resp. for the space of functions such that (resp. ). For , and , we denote by , the Sobolev ( weak derivatives in ) and Hölder spaces ( derivatives which are -Hölder continuous) on , respectively. Also, we write for the standard parabolic Hölder spaces on , e.g. will be the space of functions such that , , exist, and are -Hölder in and -Hölder in . Similarly, we will use for the standard parabolic Hölder spaces, e.g. will indicate the space of functions such that , , are in .
(resp. ) is the space of probability measures over (resp. a Polish space ), and is the set of probability measures which admit a density with respect to the Lebesgue measure. For , the Wasserstein space is denoted by and its metric is . Given a curve , we write for the space of -valued -square integrable functions over . The space of -values Borel measures over with finite total variation is denoted by . For any , is the uniform distribution on ; we will write when there is no confusion. The relative entropy of two probability measures is defined as follows
For simplicity, we write . For solutions of the Liouville equation (1.2) we will be using the symbol or to denote an element of and the symbol for the corresponding density, which we can view as an element of . We use the analogous notation for , a solution of the “perturbed” Liouville equation (1.7). We will be using the notation for the Fisher information and, for , .
We now recall the precise definition of the space .
Definition 2.1.
(i) A function with belongs to if and only if there exists a vector field such that . We denote
(ii) A vector field with belongs to if and only if there exists a matrix field such that in the sense that . We denote
As in [JW18], since (and possibly ) are not smooth, in order to get controlled entropy bounds between and we must work with entropy solutions.
Definition 2.2.
Suppose that and , so that there exists a vector field such that . A continuous map is an entropy solution to (1.7) on the time interval if solves (1.7) in the sense of distributions, exists in the weak sense for a.e and for each ,
(2.1) |
To indicate the dependence on and we call such a solution an -entropy solution. When and , we simply say that is an entropy solution of (1.2).
Remark 2.3.
(i) If , then we have the definition of entropy solution introduced in [JW18].
(ii) If satisfies (1.7) in the classical sense, then it is also an entropy solution.
(iii) It follows that any entropy solution is also an -entropy solution, for any bounded vector field , where is a vector field such that .
(iv) By virtue of Proposition 1 from [JW18]333Actually, regularity in time is not addressed in Proposition 1 of [JW18], but it is straightforward to check using the assumptions on that any entropy solution in the sense of Jabin and Wang admits a version in ., if and are bounded and , then there exists a entropy solution.
For technical reasons, we at times need to work specifically with entropy solutions of (1.2) which arise via a suitable mollification procedure. In particular, we fix throughout the paper a standard mollifier on , and we define , . Then we make the following definition.
2.2. Preliminary Results
In this subsection we state and prove some results that will be useful in the paper. The first is a refinement of a compactness result borrowed from [Dau23, Proposition 1.2] for solutions to the Fokker-Planck equation:
(2.2) |
where and .
Proposition 2.5.
Assume that, for all , solves the Fokker-Planck equation (2.2) starting from and satisfies the uniform energy estimate
(2.3) |
for some constant independent of . Then, for any , up to taking a subsequence, converges in to some . The curve is in , is absolutely continuous with respect to , for any such that it holds that
(2.4) |
and solves (2.2) starting from .
Proof.
Let . For the total variation we have
therefore . We can therefore use Banach-Alaoglu theorem and, due to standard estimates for the Fokker-Planck equation, we can deduce that, for any , converges in to some element . In fact, elementary arguments yield that the convergence also holds also in .
It is straightforward, because of the convergence, that satisfies the Fokker-Planck equation starting from . By Theorem 2.34 of [AFP00] we discover that is absolutely continuous with respect to and that (2.4) holds. The bound that (2.4) provides (when and ), yields , due to standard estimates for the Fokker-Planck equation. ∎
Remark 2.6.
We are also going to need the following two elementary lemmas, the first of which can be inferred [DE11, Proposition 2.4.2] (notice that the boundness assumption on there is not necessary) and the second of which is standard.
Lemma 2.7.
Let be a Polish space, and a function such that . Then, the following equality holds
(2.5) |
In particular, if is any Borel set, then
(2.6) |
Lemma 2.8.
Let be a sequence of probability measures over with densities converging weakly in to the probability measure . We also suppose that is an increasing sequence of subsets of converging to a set . Then, .
3. Proofs of the main results
3.1. Proof of Theorem 1.1
To prove Theorem 1.1, we are going to need the following variational lower bound for -entropy solutions.
Proposition 3.1.
Proof.
We give the proof only when for simplicity, because the proof with is the same but more notationally cumbersome.
Step 1 ( is smooth and terminal condition is mollified). Assume that is smooth. We consider the function with if and if , and, for , let be a smooth function such that on and if .
Since is assumed to be smooth, (1.1) is strongly uniquely solvable, therefore we can consider its solution and its law (under ) in the path space . By using Lemma 2.7 and Girsanov’s Theorem as in the proof of Theorem 4.1 of [BD98]444The only difference is that we have a random initial condition, which leads to the additional term in the right-hand side of (3.1), and the fact that we work on the torus, which is no problem because a version of (3.1) on the torus easily follows from the corresponding version on . , we have
(3.2) |
where the second infimum is taken over all pairs consisting of a square-integrable adapted -valued process and a continuous process satisfying
Furthermore, by the mimicking theorem [BS13, Corollary 3.7], for any such , there exists a measurable function and a process such that is a weak solution of
, and for any , . Moreover, it is clear that for any such , its law must be a weak solution of (1.2). Combining the last observations, we deduce that
(3.3) |
where the infimum is over pairs such that is a weak solution of (1.7).
Our next goal is to argue that we can restrict the infimum in (3.3) to smooth solutions of (1.7). For this, we note that by considering the infimum in (3.3) first with respect to and then with respect to , we see that we have
(3.4) |
with being the value function of a standard stochastic control problem, and in particular , where solves
(3.5) |
with terminal conditions .
Moreover, the theory of stochastic control also tells us that the optimal feedback in (3.3) is independent of , and takes the form . By parabolic regularity and the smoothness of and , and are smooth. In particular, is continuous, so for any , the minimization problem (3.4) admits a smooth minimizer . If is the weak solution of (1.7) driven by and and starting from , we deduce that is a smooth -minimizer for the infimum in (3.3). In particular, this shows that (3.3) remains true when the infimum in the right-hand side is restricted to pairs such that is smooth and is a classical (hence entropy) solution of (1.7).
Step 2 ( is smooth):
The goal in this step is to take in (3.3). By the previous step, we can find for each a pair such that is smooth, is a classical solution of (1.7) and
(3.6) |
Using (2.1) and applying Cauchy-Schwartz, it follows that
(3.7) |
However, converges to , so by the bounded convergence theorem, the left hand side of (3.6) converges and it is, therefore, bounded. Since , and are nonnegative, this implies that the terms on the right hand side of (3.6) are also uniformly bounded (independently of ). Since is uniformly bounded, due to the upper bound of , we also get that is uniformly bounded. Combining these facts with , we find that there exist constants with , which are independent of , such that (3.1) becomes
This clearly implies that , , are also uniformly bounded.
We can, now, apply Proposition 2.5 in various ways. By virtue of the uniform boundness of , , we get that as , up to subsequences
(3.8) | ||||
(3.9) |
In addition, by (3.1), we derive that is uniformly bounded independently of , so is uniformly bounded. The Vallée–Poussin theorem [BR07, Theorem 4.5.9] implies that is uniformly integrable over , hence by the Dunford-Pettis theorem [BR07, Theorem 4.7.18], the convergence in (3.8) also holds weakly (again up to subsequences):
(3.10) | ||||
(3.11) |
We now pass to the limit in (3.6). By the weak convergence of , the weak lower semicontinuity of the relative entropy, Proposition 2.5 and the nonnegativity of we deduce
(3.12) |
hence in order to prove (3.1), it suffices to show that is an admissible candidate for the infimum on its right hand side.
We will start by showing that or, equivalently, . Indeed, the family of sets is increasing and converges to as . We have by Markov’s inequality , thus because is uniformly bounded. But since is increasing and converges weakly in to , Lemma 2.8 and the last inequality imply , which is what we wanted.
We now prove that is an entropy solution. It is straightforward to check that the limit satisfies (1.7) in the weak sense. Thus, it is an admissible candidate for the first infimum in (3.3), hence
We pass to the limit as and by the weak lower semi-continuity of the entropy, Proposition 2.5 and the fact that is supported on , we get
We deduce that
(3.13) | ||||
(3.14) |
Since is a smooth -entropy solution, (2.1) holds for every and can be rewritten as
(3.15) |
We observe that since is uniformly bounded, by passing to a further subsequence if necessary, we get weakly in . Due to the lower semi-continuity of the entropy, Proposition 2.5 and the remark after Proposition 2.5, we get that the of the right hand side of (3.15) is at least
(3.16) |
On the other hand, because of the convergences (3.9), (3.13) and (3.14), the left hand side of (3.15) converges to
(3.17) |
so that satisfies (2.1) for each .
Step 3 (). Assume now that . We consider to be a family of mollifications of . For any , by step 2, (3.1) holds when . We denote by the right-hand side of (3.1) with replaced by ; so that . We wish to show that . Of course if , this is trivial, so we may assume that .
Note that since , the set is bounded. Set
Then, for an -minimizer for we have
On the one hand, since is uniformly bounded, this implies
(3.18) |
We note that due to the fact that is an -entropy solution, and (3.18), an argument as in Step 1 yields
(3.19) |
On the other hand, by Remark 2.3, is an -entropy solution, therefore
(3.20) |
To finish the proof, we will show that the integral terms in (3.20) converge to as . Indeed, for any and we have
Therefore, there exists a constant depending on the bound provided by (3.19) and such that
Now , so in , hence by passing to the limit we discover
and since was arbitrary, we get
(3.21) |
We also have by Cauchy-Schwartz
(3.22) |
Now we send to in (3.20) and use (3.21), (3.22) to to conclude that
∎
Proof of Theorem 1.1.
The first step will be to reinterpret [FG15, Theorem 2] in a convenient way. If we specialize this result to the torus, we find that for each there is a constant depending only on and such that whenever are i.i.d. -valued random variables with common law , we have In other words,
Recalling the variational formula (2.6) from Lemma 2.7, we find that the implication
(3.23) |
holds for , where .
We are now going to combine (3.23) with Proposition 3.1 to complete the proof. Suppose that we have a pair and such that is an entropy solution of (1.7). Suppose further that
holds for some . By assumption, it follows that
Some simple algebra shows that if , then (here we use )
In particular, setting (i.e. with , and applying (3.23), we find that we have the implication
In light of Proposition 3.1, this implies , or in other words ∎
3.2. Proof of Proposition 1.6
We begin by establishing the well-posedness of (1.3).
Proposition 3.2.
Proof.
The main challenge is to obtain appropriate a-priori estimates, so we explain this point in detail and then quickly sketch the existence and uniqueness part. Thus we assume for the moment that in addition to Assumption 1.5, is smooth, so that we have a unique classical solution . In what follows, can increase from line to line but can depend freely on the constants indicated in the statement of the proposition, and dependence on other parameters will be clearly indicated, e.g. indicates a constant which can depend on as well as the constants appearing in the statement of the proposition. Moreover, we let be a vector field with and . First, we have by integration parts and Young’s inequality
and so in particular , from which, by Cauchy-Schwartz, it follows that . Rewriting (1.3) in non-divergence form as
(3.24) |
a standard argument using the maximum principle shows that for each , we have
and likewise
Thus we have , and . From here, we view (3.24) as a perturbation of the heat equation, applying the standard Calderon-Zygmund estimates and then the Gagliardo-Nirenberg interpolation inequality to get for any ,
Thus for any , . Choosing a large enough , we get by Sobolev embeddings , and then by again viewing (3.24) as a perturbation of the heat equation, we get by the Schauder estimates
Thus we have established that when is smooth, the unique classical solution of (1.3) satisfies the estimates stated in the proposition. From this a-priori estimate, a standard mollification and compactness argument can be used to obtain the existence of a solution satisfying the desired bounds when is not smooth but Assumption 1.5 is in force. Uniqueness of classical solutions, meanwhile, can be easily proved in a number of ways, e.g. given two smooth solutions and one can compute and conclude via Grownall’s inequality. We omit the details. ∎
The proof of Proposition 1.6 also requires the following lemma, which is an easy extension of the main quantitative estimate of [JW18] to the “controlled” Liouivlle equation (1.7).
Lemma 3.3.
Proof.
The proof follows closely the proof of [JW18, Theorem 1], and so we report only the main difference. The first step is to mimic [JW18, Lemma 2] (here it is crucial that we work with an entropy solution ) to get
Young’s inequality immediately gives
(3.26) |
Notice that up to the last term and the factor appearing in the penultimate term, this is the same inequality appearing in [JW18, Lemma 2]. We now follow exactly the proof of [JW18, Theorem 1], applying Lemmas 3 and 4 of [JW18] (which are easily seen to apply here despite the fact that satisfies the “perturbed” Liouville equation (1.7) rather than the original Liouville equation (1.2)) to bound the second and third terms appearing on the right-hand side of (3.2). This results in the bound
with depending only on the constants indicated in the lemma. An application of Gronwall’s inequality completes the proof. ∎
Proof of Proposition 1.6.
The existence of an admissible entropy solution is proved already in [JW18, Proposition 1]. Let be the unique classical solutions of (1.2) with replacing , and be the unique classical solution of (1.3) with replacing . By Theorem 1.1 and Lemma 3.3, we have
(3.27) |
with depending only on the constants stated in the proposition (here we use the fact that and ). Let be the sequence appearing in the definition of admissible entropy solution. Notice that by the uniqueness part of Proposition 3.2, we must have for each fixed . Notice also that for large enough, we will have , so that
which, after replacing by , completes the proof. ∎
References
- [AFP00] Luigi Ambrosio, Nicola Fusco, and Diego Pallara. Functions of bounded variation and free discontinuity problems. Oxford university press, 2000.
- [BD98] Michelle Boué and Paul Dupuis. A variational representation for certain functionals of brownian motion. The Annals of Probability, 26(4):1641–1659, 1998.
- [BGV05] François Bolley, Arnaud Guillin, and Cédric Villani. Quantitative concentration inequalities for empirical measures on non-compact spaces. Probability Theory and Related Fields, 137:541–593, 2005.
- [BJW20] Didier Bresch, Pierre-Emmanuel Jabin, and Zhenfu Wang. Modulated free energy and mean field limit. Séminaire Laurent Schwartz — EDP et applications, pages 1–22, 2019-2020. talk:2.
- [BJW23] Didier Bresch, Pierre-Emmanuel Jabin, and Zhenfu Wang. Mean field limit and quantitative estimates with singular attractive kernels. Duke Mathematical Journal, 172(13):2591 – 2641, 2023.
- [BR07] Vladimir Igorevich Bogachev and Maria Aparecida Soares Ruas. Measure theory, volume 1. Springer, 2007.
- [BS13] Gerard Brunick and Steven Shreve. Mimicking an itô process by a solution of a stochastic differential equation. 2013.
- [CD22] Louis-Pierre Chaintron and Antoine Diez. Propagation of chaos: A review of models, methods and applications. i. models and methods, 2022.
- [CdCRS23] Antonin Chodron de Courcel, Matthew Rosenzweig, and Sylvia Serfaty. The attractive log gas: stability, uniqueness, and propagation of chaos, 2023.
- [CdCRS23] Antonin Chodron de Courcel, Matthew Rosenzweig, and Sylvia Serfaty. Sharp uniform-in-time mean-field convergence for singular periodic riesz flows. Ann. Inst. H. Poincaré C Anal. Non Linéaire, 223.
- [Dau23] Samuel Daudin. Optimal control of the fokker-planck equation under state constraints in the wasserstein space. Journal de Mathématiques Pures et Appliquées, 175:37–75, 2023.
- [DE11] Paul Dupuis and Richard S Ellis. A weak convergence approach to the theory of large deviations. John Wiley & Sons, 2011.
- [DLR18] François Delarue, Daniel Lacker, and Kavita Ramanan. From the master equation to mean field game limit theory: Large deviations and concentration of measure. The Annals of Probability, 2018.
- [FG15] Nicolas Fournier and Arnaud Guillin. On the rate of convergence in Wasserstein distance of the empirical measure. Probab. Theory Related Fields, 162(3-4):707–738, 2015.
- [FW23] Xuanrui Feng and Zhenfu Wang. uantitative propagation of chaos for 2d viscous vortex model on the whole space, 2023.
- [GBM24] Arnaud Guillin, Pierre Bris, and Pierre Monmarché. Uniform in time propagation of chaos for the 2d vortex model and other singular stochastic systems. Journal of the European Mathematical Society, pages 1–28, 01 2024.
- [HC23] Elias Hess-Childs. Large deviation principles for singular Riesz-type diffusive flows, 2023.
- [JW16] Pierre-Emmanuel Jabin and Zhenfu Wang. Mean field limit and propagation of chaos for vlasov systems with bounded forces. Journal of Functional Analysis, 271(12):3588–3627, 2016.
- [JW17] Pierre-Emmanuel Jabin and Zhenfu Wang. Mean Field Limit for Stochastic Particle Systems, volume 1. 04 2017.
- [JW18] Pierre-Emmanuel Jabin and Zhenfu Wang. Quantitative estimates of propagation of chaos for stochastic systems with kernels. Inventiones mathematicae, 214:523–591, 2018.
- [Lac23] Daniel Lacker. Hierarchies, entropy, and quantitative propagation of chaos for mean field diffusions. Probability and Mathematical Physics, 4(2):377––432, 2023.
- [McK69] Henry P. McKean. Propagation of chaos for a class of non-linear parabolic equations. In A. K. Aziz, editor, Lecture Series in Differential Equations, Volume 2, pages 177––194, Berlin, Heidelberg, 1969. Van Nostrand Reinhold Company.
- [RS23] Matthew Rosenzweig and Sylvia Serfaty. Global-in-time mean-field convergence for singular Riesz-type diffusive flows. The Annals of Applied Probability, 33(2):954 – 998, 2023.
- [Ser20] Sylvia Serfaty. Mean field limit for Coulomb-type flows. Duke Mathematical Journal, 169(15):2887 – 2935, 2020.
- [Szn91] Alain-Sol Sznitman. Topics in propagation of chaos. In Paul-Louis Hennequin, editor, Ecole d’Eté de Probabilités de Saint-Flour XIX — 1989, pages 165–251, Berlin, Heidelberg, 1991. Springer Berlin Heidelberg.