Quantitative relative entropy estimates on the whole space for convolution interaction forces
Abstract.
Quantitative estimates are derived, on the whole space, for the relative entropy between the joint law of random interacting particles and the tensorized law at the limiting systeme. The developed method combines the relative entropy method under the moderated interaction scaling introduced by Oeschläger, and the propagation of chaos in probability. The result includes the case that the interaction force does not need to be a potential field. Furthermore, if the interaction force is a potential field, with a convolutional structure, then the developed estimate also provides the modulated energy estimates. Moreover, we demonstrate propagation of chaos for large stochastic systems of interacting particles and discuss possible applications to various interacting particle systems, including the Coulomb interaction case.
Key words: diffusion-aggregation equation, interacting particle systems, McKean–Vlasov equations, non-linear non-local PDE, propagation of chaos, relative entropy, modulated energy.
MSC 2010 Classification: 35D30, 35Q70, 60K35.
1. Introduction
In this article we study -particle systems given by stochastic differential equations (SDEs) of the form
starting from i.i.d. initial data . Such interacting systems arise naturally in various areas of science and engineering, including physics, chemistry, biology, ecology, and social sciences. For instance, they represent the behavior of ion channels, chemotaxis [KS70, HP09, Hor04], angiogenesis on the microscopic level and swarm movement [TBL06], flocking [HL09], opinion dynamics [Lor07, Hos20], cancer invasion [DTGC14] on the macroscopic level. The macroscopic level is often described through the evolution of the density of particles/individuals known to satisfy an aggregation-diffusion equation, which in general is a non-local, non-linear partial differential equation (PDE). Transitioning from microscopic models to continuum descriptions, i.e. approaches infinity, entails to explore the mean-field limit, see e.g. [Szn91, CCH14, JW16, Jab14]. It consists of demonstrating the convergence of the empirical measure for all , where is defined as
Various topologies are considered for the convergence, such as weak convergence, convergence in Wasserstein distance, convergence in terms of the Boltzmann entropy, and convergence in terms of the Fisher information. A comprehensive analysis can be found in [HM14]. In the present article we will focus on the convergence in entropy.
Main contribution: We present a novel method to derive propagation of chaos in entropy on the whole space for both non-conservative field and potential field possessing a convolution structure. Inspired by Oelschläger [Oel87], the presented method is based on the crucial observation that, under the convolution structure, the expectation of mollified norm and the modulated energy (also as a weighted -norm) can be estimated using the dynamics of the underlying systems in conjunction with the propagation of chaos in probability, as demonstrated in [LP17, HLP20, FHS19, CNP23]. The key contribution of the present work lies in the technique of combining propagation of chaos in probability [LP17, HLL19a, LY19, FHS19, HKPZ19, CCS19, HLP20, CLPY20, CNP23] with the underlying entropy structure from [JW16, Ser20, CH23] and the fluctuation estimates in [Oel87]. Consequently, we prove that convergence in probability for an interaction kernel, which is obtained by some type of mollification technique, implies convergence in relative entropy for an algebraic cut-off . This demonstrates that convergence in probability is actually a quite strong convergence result.
We emphasize that the main quantitative estimate, Theorem 3.3, is presented in a general manner and can easily be extended to a multi-dimensional setting, allowing its application to a wide range of kernels. We refer to Remark 3.4 for more details and to Section 5 for some interesting examples from the fields of chemotaxis and opinion dynamics. In particular, the method can be further applied in handling the attractive and repulsive Coulumb interaction potential in dimension , which includes the Keller–Segel model. Finally, we derive an estimate on the supremum norm in time of the relative entropy between the law of the approximated particle system and the chaotic law of the approximated mean-field SDE system of rate greater than . Moreover, the approximation is of algebraic order, which is sharper than the logarithmic cut-off derived from the standard coupling method [Szn91, LY19].
Theorem 3.3 can be considered as an intermediate result on the approximated level. On the one hand, the remaining limit of the regularized mean-field equation to the mean-field equation reduces to a question regarding the convergence on the PDE level. On the other hand, the convergence of the regularized particle system to the particles system is a question about the stability of solutions to the stochastic differential equations. For bounded interaction kernels, we also provide both convergences in the -norm. Consequently, we prove the -convergence of the -th marginal of the Liouville equation to the -th chaotic law of the non-linear diffusion-aggregation equation. This final convergence result is only presented for bounded kernels since, in general, the existence for the linear Liouville equation (2.6) on is not given, see [BJW19, Proposition 4.2] for the torus setting.
Related literature: The study of propagation of chaos for a globally Lipschitz continuous interaction force has already a fairly long history, see e.g. [McK67, Szn91, HM14]. One of the first idea was to utilize the coupling method, i.e. comparing with their associated McKean–Vlasov SDEs.
Motivated by models, particularly from physics, with bounded measurable or even singular interaction force kernels, extensive efforts have been devoted to investigated propagation of chaos for particle systems with such kernels. Initially, approaches to treat such irregular kernels were often based on compactness methods in combination with the martingale problems associated to the McKean–Vlasov SDEs, see e.g. [Oel84, Osa87, Gär88, FJ17, GQ15, LLY19, LLY19, ORT20]. For general -interaction force kernels , the propagation of chaos was demonstrated for first and second order systems on the torus [BJS22] and on the whole space [HRZ22, Han22, Lac23]. Another approach, initiated by Lazarovici and Pickl for the Vlasov–Poisson system [LP17], allows to deduce propagation of chaos in probability. This method is well-suited for singular interaction kernels, even when the underlying systems may not be well-defined.
For the moderate interacting system in deriving porous medium equation, Oeschläger has actually a series of contributions many decades ago, for example in [Oel84, Oel87]. Especially, for the fluctuation analysis, a smoothed estimate with convergence rate has been obtained. The convolution structure of the moderate interaction played important roles. In the estimates proposed in [Oel87], the repulsive moderate interaction provides an essential quantity to absorb the rests from interacting effect. Recently, [Hol23] obtained Oeschläger’s estimate for the moderate interacting system with attractive potential, under the assumption that the convergence of probability for the moderating interacting particle system holds, which is still an open problem. Recently [CH23], derived also a connection between the relative entropy and the regularized -norm in the moderate interaction framework by directly citing the estimate from [Oel87]. The novelty of our work is that we do not follow the framework provided by [Oel87], but generate a direct estimation method in a general framework.
Another way to treat singular kernels such as the Coulomb potential for was investigated in the deterministic setting [Ser20, NRS22] ( as well as in the random setting [JW18, BJW19, BJW20, RS23] (). The aforementioned references introduced the modulated free energy, which is a practical quantity suited for the Coulomb case. In particular, it metrize the weak convergence of the empirical measures [RS23]. A drawback of the modulated free energy approach in combination with the relative entropy is the torus domain as well as the requirement of entropy solutions on the particle level (microscopic level), see [BJW19, Proposition 4.2], which is non-trivial outside a setting on the torus. Furthermore, in order to apply the large deviation result in [JW18], strict conditions are required on the initial data and the solution of the Fokker–Planck equation. Recently, Wang and Feng extended these results to the -viscous point vortex model on the whole space . The idea is to show exponential decay of the solution [FW23, Theorem 4.4] to be able to apply the large deviation result in [JW18]. Again strict restrictions on the initial conditions such as exponential decay of the initial data are necessary.
In the present article we manage to avoid the large deviation principle [JW18] and the strict conditions on the initial data by utilizing the convergence in probability, see (2.15) below. We also can treat general forces such as rotational fields or magnetic fields in physics. We also manage to derive quantitative bounds on singular forces such as attractive Coulomb interaction kernels on the whole space, which to our knowledge require approximation techniques by the nature of their singularities on the level of the Lioville equation. The price we pay lies in the obtained convergence rate. While [JW18] can establish the convergence in the sense of the Boltzmann entropy on the level of the associated Fokker–Planck equations with an order of , we achieve a rate of for some . Nevertheless, the convergence is faster than and, therefore, we are optimistic that this result can be used as a stepping stone for Gaussian fluctuation.
Organization of the paper: In Section 2 we introduce the notation, the interacting particle systems and their associated diffusion-aggregation equations, give the necessary assumptions, and list the main results of this paper. We present the main ideas and the main estimate (Theorem 3.3) in Section 3. In Section 4, we demonstrate the propagation of chaos in the case of bounded interaction forces for the non-regularized systems by establishing the convergence of the approximated PDEs to the non-approximated counterparts. In Section 5, we showcase the applicability of the developed method by discussing, e.g., the regularized, singular Keller–Segel models and bounded confidence models.
Acknowledgments: P. Nikolaev and D. J. Prömel would like to deeply thank L. Chen for fruitful discussions and suggestions leading to a significant improvement of the present work.
2. Problem setting, preliminaries and main results
In this section we introduce the basic setting, the interacting particle systems, their associated partial differential equations, some preliminary results, as well we the main results of this article.
2.1. Particle systems
In this subsection we introduce the probabilistic setting, in particular, for the -particle system and the associated McKean–Vlasov equation. To that end, let be a complete probability space with right-continuous filtration and , , be independent one-dimensional Brownian motions with respect to . In the following, we use the notation to represent that is the law of random variable .
The -particle system is given by
(2.1) |
where is the diffusion parameter and is independent of the Brownian motions , . The particle system (2.1) induces in the limiting case the following i.i.d. sequence of mean-field particles
(2.2) |
where denotes the probability density of the i.i.d. random variable .
To introduce the regularized versions of (2.1) and (2.2), we take the smooth approximation of and replace the drift term with its approximation. Hence, the regularized microscopic -particle system is given by
(2.3) |
and the regularized mean-field trajectories by
(2.4) |
where denotes the probability density of the i.i.d. random variable .
Finally, let the empirical measure of the regularized interaction system be given by
(2.5) |
where is the Dirac measure.
2.2. Associated PDEs
Itô’s formula implies that the associated probability densities of the particle systems, introduced in Subsection 2.1, satisfy partial differential equations (PDEs). Indeed, the interacting particle system (2.1) induces the following Liouville equation on ,
(2.6) |
for , the system (2.2) induces the non-linear aggregation-diffusion equation
(2.7) |
the regularized particle system (2.3) the Liouville equation
(2.8) |
and the regularized system (2.4) the aggregation-diffusion equation
(2.9) |
Note that we use and for the solutions of the PDEs (2.7) and (2.9) as well as for the probability densities of the particle systems (2.2) and (2.4), respectively, since these objects coincide by the superposition principle, see [BR20], in combination with existence results of densities for considered SDEs, see [Rom18].
Furthermore, we need to define the marginal of the system of rank ,
(2.10) |
We remark that the -th martingale solves the following Liouville equation
(2.11) |
Similar to (2.10) we denote by the m-th marginal of the approximated Lioville equation, i.e.
which solves (2.11) with instead of . Additionally, we define the chaotic law
which solves the following equation
with initial data .
2.3. Preliminary results
In this subsection, we gather essential definitions, the requisite function spaces and preliminary results for the well-posedness of the above mentioned SDEs and PDEs.
Throughout the entire paper, we use the generic constant for inequalities, which may change from line to line. The constants are always fix and will be given by our Assumptions 2.5, 2.6.
For we denote by the space of measurable functions whose -th power is Lebesgue integrable (with the usual modification for ) equipped with the norm , by the space of all measurable functions such that , by the space of all infinitely differentiable functions with compact support on , and by the space of all Schwartz functions, see [Yos80, Chapter 6] for more details.
Let be a Banach space. We denote by the space of all strongly measurable functions such that
The Banach space consists of all continuous functions , equipped with the norm
For a smooth function and a multiindex with length , we denote the derivative with respect to by , where we write or for . The derivative with respect to time we denote by . For we define the Fourier transform and inverse Fourier transform by
We denote the Bessel potential for each and by
and define the Bessel potential space for and by
Applying [Tri83, Theorem 2.5.6] we can characterize the above Bessel potential spaces for and as Sobolev spaces
where is to be understood as weak derivatives [AF03] and is the set of all multi-indices. Moreover, we will use the following abbreviation .
For the partial differential equations (2.6), (2.7), (2.8) and (2.9) we rely on the concept of weak solutions, which we recall in the next definition.
Definition 2.1 (Weak solutions).
Definition 2.2 (Weak solutions).
By the regularity of the solution in Definition 2.2 we can actually weaken the assumption on in equations (2.12) and (2.13) to .
Remark 2.3.
The divergence structure of the PDEs (2.7) and (2.9), respectively, implies mass conservation/the normalisation condition
for all under Assumption 2.4. This is an immediate consequence by plugging in a cut-off sequence, see [Bre11, Lemma 8.4], which converges to the constant function as a test function in (2.13).
Throughout the entire paper we make the following assumptions on the initial condition of the interacting particle system and the interaction force kernel .
Assumption 2.4.
The initial condition fulfills
(2.14) |
We recall some general facts, which will be used throughout the article. First, we notice that we have a solution of the regularized PDE (2.8) in the sense of Definition 2.1, which follows from the regularity of . We also have a solution in the sense of Definition 2.1 in the case and the equation is linear. By standard SDE theory we also obtain strong solutions , to the regularized SDEs (2.3), (2.4). For the well-posedness of the particle system (2.1) and McKean–Vlasov SDE (2.2) we refer to [HRZ22, Theorem 3.7] and [HRZ22, Theorem 4.10], respectively. Additionally, [CNP23, Section 3] guarantees the well-posedness of PDEs (2.9), (2.7), which are bounded in time and space uniformly in . Consequently, our framework is well-defined and, in particular, the empirical measure given by (2.5) is well-defined.
The analysis of the entropy relies on the convergence of the particle system (2.3) to the particle system (2.4) in probability. Hence, we introduce the following convergence in probability assumption.
Assumption 2.5.
This assumptions is satisfied by a variety of models [LP17, HLL19a, LY19, FHS19, HKPZ19, CCS19, HLP20, CLPY20]. In particular for bounded or even singular kernels this assumption is fulfilled, see [CNP23].
Furthermore, we need the following law of large numbers result.
Assumption 2.6.
Let and be given by (2.4). Assume further that , , with , and define for the following sets
Then, for each there exists a such that
(2.16) |
for every , where the constant is independent of .
2.4. Main results:
Let with a given mollification kernel and let be a cut-off function, which satisfies , on and on , . We need the following assumptions on the mollified version of interaction force separately to state the main result of this paper.
Definition 2.7.
We say are admissible approximations, if and with
(2.17) |
for some and . We say admissible approximations are strongly admissible approximations, if the above inequality holds for instead of .
In general we will consider two type of forces. First, and second . The potential field structure of the latter one will be required for the definition of the modulated energy (see Section 3). This assumption on include many different forces, where no potential field is needed.
Remark 2.8.
Some typical examples for the above structure are as follows:
-
(1)
The interaction force kernel . Then and is just the standard mollified version of .
-
(2)
If for , we can choose and , which is also just a mollification of .
-
(3)
If we may choose and , where is defined as a cut-off function to guarantee integrability of the mollification .
The first main result of this paper is the propagation of chaos on the mollified level with :
Theorem 2.9.
Let and be the non-negative solutions of (2.8) and of (2.9) respectively. Assume that the convergence in probability, Assumption 2.5, and the law of large numbers, Assumption 2.6 hold for . Let and be admissible in the sense of Definition 2.7 with rate . Then there exists a such that , the following propagation of chaos result holds for between (2.8) and of (2.9).
(2.18) |
where is the -marginal density of .
Remark 2.10.
In obtaining the estimate for smoothed modulated energy, the proof has been done with the identity
where is the reflection. Again choosing for instance we may borrow an additional factor from the mollification kernel , which will weaken the convergence rate estimate, or in other words, one has to choose even smaller to achieve the order . The restriction is in place to guarantee the order . The convergence of the relative entropy holds also without this restriction.
Additionally, for bounded force, we know from [CNP23] that convergence in probability holds for approximations , which satisfy a local Lipschitz bound. Therefore, we can obtain the propagation of chaos result without mollification.
Theorem 2.11.
Remark 2.12.
The Theorem holds for more general approximation as long as the approximation and the convergence in probability holds. We refer to [CNP23] for an overview of the topic of convergence in probability in the bounded case .
Let us finish the section with an overview over the constants:
-
•
provides the rate on the distance of the particles in the convergence in probability
and in the law of large numbers
-
•
provides the maximum interval for the cut-off parameter , for which the convergence in probability and law of large numbers hold.
-
•
is the convergence rate of the approximated particles such that .
-
•
provide the maximum intervals such that the relative entropy and modulated energy converges with rate greater than , (see (2.18)).
3. Relative entropy method
This section is devoted to present the relative entropy method for the moderate interacting problem and its connection to the estimate proposed by Oelschläger [Oel87]. We derive the smoothed estimate for given force (no requirement as a potential field), and the smoothed modulated energy for potential field with convolution structure. Both lead to the estimate of the relative entropy between and .
The main idea is to use the assumption of convergence in probability (Assumption 2.5), the structure of the PDEs (2.8), (2.9) and the law of large number (Assumption 2.6). Applying the Csiszár–Kullback–Pinsker inequality [Vil09, Chapter 22] we provide an estimate on the -norm of the marginals and for fix .
We emphasize that the method developed in Theorem 3.3 can be applied in different settings. Indeed, since we are working on the approximation level, our assumptions are only needed in the regularized setting. Hence, in general the assumptions on , and itself can be chosen more irregular, extending even to singular models. We refer to Remark 3.4 and the applications Section 5 for more details.
3.1. Relative entropy and modulated energy
In this section we introduce our main quantities the relative entropy and the modulated free energy. We then show the connection between the -norm
(3.1) |
the relative entropy as well as the expectation of the modulated free energy . This can be viewed as a combination of Oelschläger’s results on moderated interaction and fluctuations [Oel87] and the relative entropy method developed among others by Serfaty, Jabin, Wang, Bresch and Lacker [JW16, JW18, BJW19, BJW20, Ser20, NRS22, RS23, BJS22, Lac23] for the mean-field setting. The aim is to demonstrate how both concepts connect under the convolution assumption. Finally, we derive an estimate on the relative entropy in terms of the above -norm.
Following [BJW19], we introduce the modulated free energy
where
is the relative entropy introduced in [JW16] and if is a potential
is the expectation of the modulated energy. We refer to [BJW19] and the references therein for more details on the modulated free energy.
Let us now explore some connections between the relative entropy and the structure presented by Oelschläger [Oel87]. We start by rewriting the expectation of the free energy by using our convolution structure. A straightforward calculation shows
(3.2) |
where is the reflection. Applying Young’s inequality we see that it is enough to control a term of the form
for some function , where we just write for simplicity and understand that we can chose in all calculations below. Hence, in order to estimate we can estimate the -difference between the convoluted empirical measure and the solution the law of the mean-field limit (2.2). This will be accomplished in Theorem 3.3.
But let us recall that our initial goal is to estimate the relative entropy and not . Therefore, let us connect the relative free energy to the -norm of the derivative .
Lemma 3.1.
Proof.
Let us compute the time derivative of the relative entropy
For we have further estimates
Substituting the above estimate into the first inequality, while recalling that , proves the lemma. ∎
Remark 3.2.
Depending on the regularity of and one may choose to interchange the roles in the estimate. Generally, one should choose the more regular function to be . Indeed, in the above estimate we need only the -norm of , while later on in Theorem 3.3 we need the -norm as well as the -norm of not only the function but also of its derivatives. Moreover, if the force is a potential field, the last term has the following structure
which will also be estimated by Theorem 3.3. Hence, we do not lose convergence rates in the case , but as already mentioned, we obtained an additional estimate on the modulated energy .
3.2. -estimate
In this section we concentrate on estimating the rest term in the entropy estimate (3.3).
We present the main theorem of the article, which is formulated for a function , which depends on . This presentation is motivation by our case . We emphasize that the function in the following Theorem can be chosen independent of , but than the estimate has no connection to the modulated energy or the relative entropy (see Lemma 3.1).
Theorem 3.3.
Remark 3.4.
The only ingredients we need for completing the proof of theorem 3.3 are the convergence in probability of the particle system to the mean-field limit (2.15) as well as the law of large numbers (2.16). But the convergence in probability and the law of large numbers are known for a variety of interaction force kernels, see for instance [LP17, FHS19, HLL19b, HLP20]. Hence, this result can be extended for a variety of interaction force kernels. Moreover, the kernels can also be -dimensional since the estimates we used are dimension-free. We refer to Section 5 for applicable models such as the case with Coulomb force. Actually, the estimates become dimension dependent by the choice of . Consequently, the rate of convergence becomes dependent on the dimension. Nevertheless, the steps of the proof work analogously in multi-dimensional setting by replacing the multiplication with the scalar product, the absolute value with the Euclidean norm and the Itô’s formula with its multidimensional counter part.
Remark 3.5.
The results in Theorem 3.3 state that is close to in the mollified -norm. By the propagation of chaos we expect that this quantity should be small since should ideally vanish in the limit. The majority of work, which lies ahead, is to estimate this -norm with a good rate. In the process we will also obtain an estimate on the derivative . This is no surprise, since the estimate follows the structure of the classic a priori -estimate for the parabolic equation [WYW06, Chapter 3]. As a result, we obtain in the -norm an -bound and as usual an -bound for the derivative. In combination with Lemma 3.1 this will allow us to obtain a bound on the relative entropy . Additionally, if the interaction force is a potential field we obtain an estimate for by equality (3.2).
Let us start by describing the dynamic of the empirical measure . Applying Itô’s formula to a sufficiently smooth function , we obtain
Taking the expectation and using the fact that we have a density of , provides a weak formulation of the Lioville equation (2.8). If we want to compare it to the mean-field law, we need to make the crucial observation that the stochastic integral in the above equation should vanish after taking the expectation. In other words, we have no term in the regularized PDE (2.9), which corresponds to the stochastic integral. If the integrand is smooth enough then obviously the stochastic integral vanishes. However, we need to compute the following difference
Therefore, we need somehow transfer the naive approach to the more complex expected value. Applying the above dynamic we prove the following lemma, which allows us to treat the convolution as if the stochastic integral vanishes.
Lemma 3.6.
Let defined by (2.5). Then, we have the following inequality
Proof.
We use Itô’s formula, the dynamics (2.3) and the Burkholder–Davis–Gundy inequality to find
It remains to estimate the last term by the Burkholder–Davis–Gundy (BDG) inequality,
Inserting this calculation into the previous inequality proves the lemma. ∎
Proof of Theorem 3.3.
By Lemma 3.6 we can ignore the stochastic integral in the processes , which determine the empirical measure . Hence, let us write
for the convolution after applying Itô’s formula but without the stochastic integral. Then, we have
where we notice that for the initial time , we have by definition. Let us remark that since all integrands are smooth enough we have . Next, plugging in and differentiate we obtain
Similar, is a weak solution to our PDE (2.9), which implies
Combing the last two calculations, we find
The goal is now to insert back into the equation. Hence, for the absorption term we have
and for the last term
Applying Lemma 3.6 and put together the above estimates we have shown
(3.4) | ||||
Now we want to estimate each term on its own. We will split the fourth terms into fourth separate lemmata to keep a readable structure. The theorem follows immediately by combining Lemma 3.8 and the inequalities (3.5), (3.9), (3.20) in the lemmata below. We will summarize the estimate after we prove the following lemmata.
Lemma 3.7 (Initial Value Inequality).
Let the assumptions of Theorem 3.3 hold true. Then
(3.5) |
Proof.
We compute
where we used the fact that the initial particles are i.i.d. and Young’s inequality for convolutions in the last step. ∎
Lemma 3.8 (Absorbation Inequality).
Let the assumptions of Theorem 3.3 hold true. Then
Proof.
Before we begin the proof of this lemma, we will provide an overview of our approach. Our main strategy is to utilize the convergence in probability of the particle to their mean-field limit (Assumption 2.5) in combination with the law of large numbers (Assumption 2.6). This implies that the ”bad set”, where the particles are apart is small in probability with arbitrary algebraic convergence rate. Therefore, we may assume that is close to , and we formally replace the empirical measure of with the empirical measure associated with . However, has more desirable properties. For instance, the particles are independent and have density and often even . This allows us to apply the law of large numbers (2.16), which ultimately proves the claim.
Let us start by splitting our probability space into two sets. On one set the particles are close to the mean-field particles in probability and “satisfy” the law of large numbers. The other set we take as the complement , which has small probability by inequalities (2.15) and (2.16).
More precisely, we have
(3.6) |
for some such that and we have the estimate for all by (2.15) and (2.16). Let us rewrite the last Lebesgue integral on the left-hand side of our claim as follows
We are going to estimate each term by itself.
On the set : In order to estimate the first term above we let and will not write the indicator function. Then we have
For the first term we obtain
(3.7) |
Here we used integration by parts in the first step, the property of the set in the fourth step. As always, we neglect the last term by absorbing it into the diffusion in our statement.
We treat the term using the law of large numbers property of the second term in . For we rewrite
(3.8) |
For the first term we obtain
(3.9) |
where we used the property of the set in the second step and Young’s inequality.
Using the fact that we are still on the set we obtain for the second term the following estimate
(3.10) |
In the above calculations we used Young’s inequality in the first step, Jensen inequality in the second estimate, the property of the set in the third estimate.
In order to estimate the last term in (3.2) we use the independence of our mean-field particles . Hence, we can no longer do the estimates pathwise and need to take advantage of the expectation. First, applying Young’s inequality we find
As always, the last term is going to be absorbed. For the first term, we recall that our statement has an supremum over all and an expectation. Hence, it is enough to estimate
Let us denote for fix
Then we notice that
Furthermore, we have the random variables are pairwise independent. Hence, if we find
We notice that we have
On the other hand by using the trivial inequality and Young’s inequality for convolution we obtain
Hence, the estimate for follows by the previous law of large numbers argument and is obtained in the following
(3.11) | ||||
By combining the estimates (3.2)(3.2)(3.11) with (3.2) and (3.7) we obtain the estimate on the set
(3.12) | ||||
It remains to obtain an estimate on the complement of .
On the set : Applying Young’s inequality, multiple Hölder’s inequalities, the fact that , we obtain
Combined with the estimate on the set , we obtained the result. ∎
Lemma 3.9 (Stochastic Remaining Term Inequality).
Let the assumptions of Theorem 3.3 hold true. Then
(3.13) | ||||
Proof.
We carry out a similar strategy as in the previous Lemma 3.8. Again, we want to split into a good and bad set. Remember the definition of set in (3.2), we do the estimates on and its complement separately.
On the set : Let , then we insert the i.i.d. process and split the estimate further into two terms
Further, utilizing the property of the set and the Burkholder–Davis–Gundy inequality we obtain
(3.14) | ||||
where we have used the estimate
(3.15) | ||||
This completes the estimate of on the set . Next, for we rewrite in the following way
For the first term , applying the the property of the set , we find with the help of the estimate (3.15) for the stochastic term that
(3.16) | ||||
where we used Fubini’s Theorem in the second step. For the term compute
where we utilized the property of in the third step, followed by the application of Hölder’s inequality and Minkowski’s inequality. Consequently, applying the Burkholder–Davis–Gundy inequality we obtain
(3.17) | ||||
For we use again the Burkholder–Davis–Gundy inequality to estimate the stochastic integral and by the law of large number argument similar to the term in Lemma 3.8, and obtain
(3.18) | ||||
This completes the estimate on the set , namely
(3.19) | ||||
On the set : Using for all by Assumption 2.6, the Burkholder–Davis–Gundy inequality, Hölder’s inequality, we obtain
This completes the estimate on the set and we have shown our Lemma. ∎
Lemma 3.10 (Stochastic Integral Inequality).
Under the assumptions of Theorem 3.3 we have the following -estimate for the stochastic integral,
(3.20) |
Proof.
An application of the Burkholder–Davis–Gundy inequality implies
∎
Continuation of the proof of theorem 3.3.
We are ready to input the estimates from above lemmata in the the inequality (3.4). We find
The above estimate is the most general one we obtain. In the following we simplify it to derive a usable estimates. In the process we may loose some convergence rate, depending on the concrete problem at hand. Noticing that by mass conservation
by keeping all the and dependent terms and put all the other constants into a universal constant , which depends on , , , , we obtain
In the above estimates, and are also used and the Theorem is proven. ∎
In the our main setting we provide the following rough estimate.
Corollary 3.11.
Let and be admissible with rates . If Theorem 3.3 holds, then
Proof.
Estimating all norms of by and using Young’s inequality to find
Hence the right hand side of the main inequality in Theorem 3.3 can be estimated by
∎
Now, that we have proven our main estimate, we are ready to demonstrate the relative entropy estimates by combining Theorem 3.3 and Lemma 3.1. We start with the first main result of this paper
Proof of Theorem 2.9.
We combine the assumptions of Theorem 3.3 and the results from Lemma 3.1, Theorem 3.3 and Corollary 3.11, to find a small so that for , and small
This allows us to demonstrate strong convergence in the -norm. Indeed, let us recall the Csiszár–Kullback–Pinsker inequality [Vil09, Chapter 22], which states that for any and function we have
(3.21) |
and the relative entropy inequaity [DMM01, Lemma 3.9]
(3.22) |
for . Consequently,
In the case the estimate (2.18) is derived analogously. The key is to recognize that we actually derived an estimate on the derivative of , which we have not used so far. In the case of we utilize it and as a result we obtain the same convergence rates. The estimate for the modulated energy follows also directly from equality (3.2), Young’s inequality and an application of Theorem 3.3 for and under the assumption that are strongly admissible. ∎
3.3. Special Choices of and
We present a series of corollaries for Theorem 3.3 for different choices of . In most applications we want to take a mollified sequence. In the special case we obtain the following corollary.
Corollary 3.12.
Suppose Theorem 3.3 holds true. Let be a mollification, then for with some and , for some , then we obtain the following -estimate ,
for a constant , which depends on , , , . In particular if and the above estimate holds with and .
Proof.
If , we obtains easily that
Therefore, we obtained with ,
The second claim follows by Young’s inequality and the scaling of the mollifier. More precisely,
∎
Corollary 3.13.
Proof.
Next, we provide a similar corollary in the case the force is a potential and has a convolution structure.
Corollary 3.14.
Proof.
Since we know that and therefore we only need to estimate the -norm for in inequality (3.3) to obtain the convergence rates. We emphasize that in Theorem 3.3 we also obtained an estimate on the gradient . Consequently, we can use Theorem 3.3 for the function . By the estimate
for any we know that
(3.23) | |||
Plugging in all estimates with into Theorem 3.3 and having equality (3.2) in mind we obtain the rate of and the estimate on the modulated energy. ∎
We have now shown in two cases how to derive explicit estimates on the relative entropy with the help of Theorem 3.3. In general, if the function have low regularity, we need to mollify them to make them admissible. Hence, we borrow the necessary regularity from and consequently, get higher rates of in our estimates. Compare for instance Corollary 3.13 and Corollary 3.14. Therefore the estimate by using the regularity of the term, will lead to weaker convergence rates. The benefit is of course that one does not require a potential field and the convolution structure of the potential.
4. De-regularization of the high dimensional PDE and the limiting PDE
The goal of this section is to prove Theorem 2.11, i.e. the strong form of propagation of chaos on the PDE level in the -norm. For the de-regularization of Liouville equation (2.8) we need . We take the following approximation . We need convergence results between and as well as and . The latter convergence was shown in [CNP23, Section 3]. More precisely,
(4.1) |
It remains to show that the approximated Liouville equation converges in entropy to the Liouville equation. An application of inequality (3.21) implies also the -convergence.
Lemma 4.1.
Proof.
We start by computing the time derivative of . We have
Now, it is enough to show, that the last term vanishes for and consequently for . We start by using the fact that the particle system (2.3) is exchangeable. We obtain
Hence, we obtained an expression in which the dimension does not change in the limit. By applying mass conservation, the Csiszár–Kullback–Pinsker inequality (3.21) and inequality (3.22) we further estimate the term
Plugging this estimate into our above entropy calculation and taking the supremum in time proves the lemma. ∎
Combing both implies the strong convergence on the PDE-level of any observable to the law in the -norm.
Proof of theorem 2.11.
For , let , therefore we take and with . By assumption of the Theorem (see also [CNP23, Theorem 6.1]), there exists a such that for all the convergence in probability, Assumption 2.5, and the law of large numbers, Assumption 2.6 both hold. Therefore we can apply the result from Corollary 3.13 for and obtain the convergence of the relative entropy to zero. We can even get better convergence rate , since we are not interested in the order of convergence of the relative entropy. Applying (3.21) and (3.22), we obtain
As mentioned the second term converges to zero. For the first term we use the inequality in Lemma 4.1 together with the fact that the converges to zero and the dominated convergence to obtain
where the last equality follows by (4.1). Consequently, it remains to show that the third term vanishes, i.e.
(4.2) |
Again this follows by (4.1) and an induction argument. Indeed let us assume , then by mass conservation we have
which proves the initial case for the induction. Now, by the same argument one can prove the induction step and therefore equation (4.2). ∎
5. Application
We provide some examples for which Theorem 2.9 can be shown with the same techniques developed in Section 3. In particular we demonstrate the convergence in relative entropy in the attractive Coulomb case on the whole space. Note that the rate of converges may vary across these examples. As stated in Remark 3.4 we only need the existence of approximated PDE (2.9), the particle system (2.4), the convergence in probability of the particle system to the mean-field limit (Assumption (2.5)) and the law of large numbers (Assumption 2.6). Since we are working on the regularized level, we can often assume the existence of the above results.
Although the result in Theorem 2.9 also works with rotational field, it worth to study directly a convolution type of potential field to achieve better cut-off rate, in other words, to allow bigger . For a given potential field, the challenging part is to find a convolution structure for the potential described in Section 3. The first idea to obtain interesting kernels, beside the Delta-Distribution, which was given in [Oel87], is to look at infinite divisible distributions. Assume that is infinitely divisible. Then, there exists a such that . Hence, if we can approximate the antiderivative of our kernel by a infinitely divisible distribution (multiplied by a constant if necessary) we are able to find candidates for interesting kernels.
Another powerful tool is the Fourier analysis. On the Fourier side the equation becomes
which can be explored. In particular for singular kernels we have representations of the Fourier transforms, see for instance [Ste70]. Consequently, we can use this approach to obtain a wide range of interesting examples used in biology or physics.
In the rest of the section we provide some fascinating examples for which the case of convolution structure in Theorem 3.3 can be obtained.
5.1. Uniform bounded confidence model
Let be a complex-valued function. Then
is a Lipschitz-continuous function with bounded support. Furthermore, we have
Consequently, the uniform bounded confidence model, satisfies the assumption of Section 3 with the usual mollification approximation. Also, it is well known that the indicator function for all . We also have the convergence in probability by [CNP23, Lemma 4.7, Theorem 6.1.]. Hence, we obtain the following proposition
Proposition 5.1.
Let be given above, then the first marginal of the law of the system converges to the law of in the -norm.
5.2. Parabolic-Elliptic Keller–Segel System
In this subsection we provide an approximation for the elliptic-parabolic Keller–Segel model [KS70] in . The underlying PDE is given by
for . Decoupling the above system by setting with being the fundamental solution of the Laplace equation we can formally derive the above equation from the particle system (2.3) with the interaction force kernel . In particular, if we have
is the fundamental solution of the Laplace equation.
In the following we present two approaches to mollify our kernel. For the first approach, let us define a mollification kernel , which satisfies , and and is infinitely differentiable. As always we set . Then satisfies all properties of [HLL19a, Theorem 2.1]. Hence, the convergence in probability Assumption 2.5, the law of large numbers Assumption 2.6 is satisfied.
Hence, under consideration of Remark 2.10 we can obtain a relative entropy convergence results on the approximated -dimensional attractive Keller–Segel system on the whole space . We formulate the following proposition as combination of Lemma 3.1 and Theorem 3.3.
Proposition 5.2.
Remark 5.3.
By going through the proof of Theorem 3.3 one can obtain a convergence rate and precise condition for . Furthermore, by inequality (3.21) we have proven convergence of the -norm of the marginals
It is also well-known that under additional assumptions on the initial data and in the sub-critical regime in the case the density converges in . Hence, we have shown that in the sub-critical case the density of the two marginal converges in to the solution of the Keller–Segel equation.
In the case we can obtain an even better approximation, which has a symmetric convolution structure given by . Indeed, define the approximation of as . Then, for and
(5.1) |
we have
(5.2) |
More precisely, for fix we have for all . Hence the Fourier transform is well-defined and by the Hardy–Littlewood–Sobolev inequality [Ste70, Chapter 5, Theorem 1], [LL01, Corollary 5.10] and the Fourier transform exists. Similar . A simple calculation shows
since and is a Schwartz function. As a result, to verify (5.2) we need to show
where the right-hand is square integrable. Now, by [LL01, Corollary 5.10] we have
where the left-hand side is in by similar arguments as before. Therefore, (5.2) is proven and we can find an appropriate approximation for the Keller–Segel interaction kernel. In particular, we can derive similar estimates to (3.23) with the help of Fourier analysis and the Hardy–Littlewood–Sobolev inequality [Ste70, Chapter 5. Theorem 1]. Clearly, this estimates will now depend on the dimension and therefore the convergence rate parameters also depend on the dimension .
Proposition 5.4.
Another approach to approximate the Coulomb kernel is by utilizing the following approximation in dimension ,
(5.3) |
where
(5.4) |
is the Weierstrass kernel. Indeed, we note first that the square root is well-defined since the Fourier transform of a Gaussian is still a Gaussian or in other words the normal distribution is infinitely divisible. More precisly, by [LL01, Theorem 5.2] we have
Hence, similar to the first approximation, we obtain Proposition 5.4. By using the Weierstrass kernel over an abstract mollification kernel we obtain explicit sharp convergence rates. For instance, using Plancherel theorem we obtain
Remark 5.5.
The above potential is attractive and therefore as far as we know regularization/approximation is necessary to obtain a solution of the underlying Lioville equation on . Nevertheless, one can obtain tightness of the empirical measure in the super subcritical regime [FJ17]. Our approach provides propagation of chaos of the intermediate system on the level of the relative entropy. Hence, it can be used as a tool to develop further results on the propagation of chaos for the Keller–Segel model without regularization.
5.3. Parabolic-Elliptic Keller–Segel System with Bessel potential
Let us recall the parabolic-elliptic Keller–Segel model [KS70] in given by
Again solving the second equation by setting
with the function defined by
(5.5) |
we can decouple the system and obtain an analogous result by using the following approximations of ,
(5.6) |
or
(5.7) |
where is the Weierstrass kernel given by (5.4). Setting
it can be shown similar to the elliptic-parabolic Keller–Segel model that
Consequently, we obtain the analogous result.
Proposition 5.6.
Let with defined by (5.5) (5.6) or (5.7) and suppose the Assumptions 2.5 and 2.6 hold. Moreover, for this let be the solution of the Lioville equation (2.8) and be the solution to regularized Keller–Segel equation, i.e. to the PDE (2.9). Then, there exists a depending on the dimension such that for there exists a such that
Remark 5.7.
By going through the proof of Theorem 3.3 one can obtain a convergence rate and precise condition for . We also assumed the convergence probability, since we can not reference a concrete result. Nevertheless, we think that this assumption should be true for good enough initial data.
References
- [AF03] Robert A. Adams and John J. F. Fournier, Sobolev spaces, second ed., Pure and Applied Mathematics (Amsterdam), vol. 140, Elsevier/Academic Press, Amsterdam, 2003.
- [BJS22] Didier Bresch, Pierre-Emmanuel Jabin, and Juan Soler, A new approach to the mean-field limit of Vlasov–Fokker–Planck equations, ArXiv preprint arXiv:2203.15747 (2022).
- [BJW19] Didier Bresch, Pierre-Emmanuel Jabin, and Zhenfu Wang, On mean-field limits and quantitative estimates with a large class of singular kernels: application to the Patlak-Keller-Segel model, C. R. Math. Acad. Sci. Paris 357 (2019), no. 9, 708–720.
- [BJW20] by same author, Mean-field limit and quantitative estimates with singular attractive kernels, ArXiv preprint arXiv:2011.08022 (2020).
- [BR20] Viorel Barbu and Michael Röckner, From nonlinear Fokker-Planck equations to solutions of distribution dependent SDE, Ann. Probab. 48 (2020), no. 4, 1902–1920.
- [Bre11] Haim Brezis, Functional analysis, Sobolev spaces and partial differential equations, Universitext, Springer, New York, 2011.
- [CCH14] José Antonio Carrillo, Young-Pil Choi, and Maxime Hauray, The derivation of swarming models: mean-field limit and Wasserstein distances, Collective dynamics from bacteria to crowds, CISM Courses and Lect., vol. 553, Springer, Vienna, 2014, pp. 1–46.
- [CCS19] José A. Carrillo, Young-Pil Choi, and Samir Salem, Propagation of chaos for the Vlasov-Poisson-Fokker-Planck equation with a polynomial cut-off, Commun. Contemp. Math. 21 (2019), no. 4, 1850039, 28.
- [CH23] Alexandra Holzinger Chen, Li and Xiaokai Huo, Quantitative convergence in relative entropy for a moderately interacting particle system on , ArXiv preprint arXiv:2311.01980 (2023).
- [CLPY20] Li Chen, Xin Li, Peter Pickl, and Qitao Yin, Combined mean field limit and non-relativistic limit of Vlasov-Maxwell particle system to Vlasov-Poisson system, J. Math. Phys. 61 (2020), no. 6, 061903, 21.
- [CNP23] Li Chen, Paul Nikolaev, and David J. Prömel, Well-posedness of diffusion-aggregation equations with bounded kernels and their mean-field approximations, ArXiv preprint arXiv:2310.13463 (2023).
- [DMM01] P. Del Moral and L. Miclo, Genealogies and increasing propagation of chaos for Feynman-Kac and genetic models, Ann. Appl. Probab. 11 (2001), no. 4, 1166–1198.
- [DTGC14] Pia Domschke, Dumitru Trucu, Alf Gerisch, and Mark A. J. Chaplain, Mathematical modelling of cancer invasion: implications of cell adhesion variability for tumour infiltrative growth patterns, J. Theoret. Biol. 361 (2014), 41–60.
- [Eva15] Lawrence C. Evans, Partial differential equations, second edition, reprinted with corrections ed., Graduate studies in mathematics; volume 19, Providence, Rhode Island, 2015.
- [FHS19] Razvan C. Fetecau, Hui Huang, and Weiran Sun, Propagation of chaos for the Keller-Segel equation over bounded domains, J. Differential Equations 266 (2019), no. 4, 2142–2174.
- [FJ17] Nicolas Fournier and Benjamin Jourdain, Stochastic particle approximation of the Keller-Segel equation and two-dimensional generalization of Bessel processes, Ann. Appl. Probab. 27 (2017), no. 5, 2807–2861.
- [FW23] Xuanrui Feng and Zhenfu Wang, Quantitative Propagation of Chaos for 2D Viscous Vortex Model on the Whole Space, ArXiv preprint arXiv:2310.05156 (2023).
- [Gär88] Jürgen Gärtner, On the McKean-Vlasov limit for interacting diffusions, Math. Nachr. 137 (1988), 197–248.
- [GQ15] David Godinho and Cristobal Quiñinao, Propagation of chaos for a subcritical Keller–Segel model, Annales de l’Institut Henri Poincaré, Probabilités et Statistiques 51 (2015), no. 3, 965–992.
- [Han22] Yi Han, Entropic propagation of chaos for mean field diffusion with interactions via hierarchy, linear growth and fractional noise, ArXiv preprint arXiv:2205.02772 (2022).
- [HKPZ19] Seung-Yeal Ha, Jeongho Kim, Peter Pickl, and Xiongtao Zhang, A probabilistic approach for the mean-field limit to the Cucker-Smale model with a singular communication, Kinet. Relat. Models 12 (2019), no. 5, 1045–1067.
- [HL09] Seung-Yeal Ha and Jian-Guo Liu, A simple proof of the Cucker-Smale flocking dynamics and mean-field limit, Commun. Math. Sci. 7 (2009), no. 2, 297–325.
- [HLL19a] Hui Huang, Jian-Guo Liu, and Jianfeng Lu, Learning interacting particle systems: diffusion parameter estimation for aggregation equations, Math. Models Methods Appl. Sci. 29 (2019), no. 1, 1–29.
- [HLL19b] by same author, Learning interacting particle systems: diffusion parameter estimation for aggregation equations, Math. Models Methods Appl. Sci. 29 (2019), no. 1, 1–29.
- [HLP20] Hui Huang, Jian-Guo Liu, and Peter Pickl, On the mean-field limit for the Vlasov-Poisson-Fokker-Planck system, J. Stat. Phys. 181 (2020), no. 5, 1915–1965.
- [HM14] Maxime Hauray and Stéphane Mischler, On Kac’s chaos and related problems, J. Funct. Anal. 266 (2014), no. 10, 6055–6157.
- [Hol23] Alexandra Holzinger, Rigorous derivations of diffusion systems from moderately interacting particle models, Dissertation, Technische Universität Wien]. reposiTUm. https://doi.org/10.34726/hss.2023.112186 (2023).
- [Hor04] Dirk Horstmann, From 1970 until present: the Keller-Segel model in chemotaxis and its consequences. II, Jahresber. Deutsch. Math.-Verein. 106 (2004), no. 2, 51–69.
- [Hos20] Noorazar Hossein, Recent advances in opinion propagation dynamics: a 2020 survey, The European Physical Journal Plus 135 (2020), no. 6, 1–20.
- [HP09] T. Hillen and K. J. Painter, A user’s guide to PDE models for chemotaxis, J. Math. Biol. 58 (2009), no. 1-2, 183–217.
- [HRZ22] Zimo Hao, Michael Röckner, and Xicheng Zhang, Strong convergence of propagation of chaos for McKean–Vlasov SDEs with singular interactions, ArXiv preprint arXiv:2204.07952 (2022).
- [Jab14] Pierre-Emmanuel Jabin, A review of the mean field limits for Vlasov equations, Kinet. Relat. Models 7 (2014), no. 4, 661–711.
- [JW16] Pierre-Emmanuel Jabin and Zhenfu Wang, Mean field limit and propagation of chaos for Vlasov systems with bounded forces, J. Funct. Anal. 271 (2016), no. 12, 3588–3627.
- [JW18] by same author, Quantitative estimates of propagation of chaos for stochastic systems with kernels, Invent. Math. 214 (2018), no. 1, 523–591.
- [KS70] Evelyn F. Keller and Lee A. Segel, Initiation of slime mold aggregation viewed as an instability, J. Theoret. Biol. 26 (1970), no. 3, 399–415.
- [Lac23] Daniel Lacker, Hierarchies, entropy, and quantitative propagation of chaos for mean field diffusions, Probab. Math. Phys. 4 (2023), no. 2, 377–432.
- [LL01] Elliott H. Lieb and Michael Loss, Analysis, second ed., Graduate Studies in Mathematics, vol. 14, American Mathematical Society, Providence, RI, 2001.
- [LLY19] Lei Li, Jian-Guo Liu, and Pu Yu, On the mean field limit for Brownian particles with Coulomb interaction in 3D, J. Math. Phys. 60 (2019), no. 11, 111501, 34.
- [Lor07] Jan Lorenz, Continuous Opinion Dynamics Under Bounded Confidence: A Survey, International Journal of Modern Physics C 18 (2007), no. 12, 1819–1838.
- [LP17] Dustin Lazarovici and Peter Pickl, A mean field limit for the Vlasov–Poisson system, Arch. Ration. Mech. Anal. 225 (2017), no. 3, 1201–1231.
- [LY19] Jian-Guo Liu and Rong Yang, Propagation of chaos for the Keller-Segel equation with a logarithmic cut-off, Methods Appl. Anal. 26 (2019), no. 4, 319–347.
- [McK67] H. P. McKean, Jr., Propagation of chaos for a class of non-linear parabolic equations, Stochastic Differential Equations (Lecture Series in Differential Equations, Session 7, Catholic Univ., 1967), Air Force Office Sci. Res., Arlington, Va., 1967, pp. 41–57.
- [NRS22] Quoc-Hung Nguyen, Matthew Rosenzweig, and Sylvia Serfaty, Mean-field limits of Riesz-type singular flows, Ars Inven. Anal. (2022), Paper No. 4, 45.
- [Oel84] Karl Oelschläger, A martingale approach to the law of large numbers for weakly interacting stochastic processes, Ann. Probab. 12 (1984), no. 2, 458–479.
- [Oel87] Karl Oelschläger, A fluctuation theorem for moderately interacting diffusion processes, Probab. Theory Related Fields 74 (1987), no. 4, 591–616.
- [ORT20] Christian Olivera, Alexandre Richard, and Milica Tomasevic, Quantitative approximation of nonlinear fokker-planck equations with singular potential by interacting particle systems.
- [Osa87] Hirofumi Osada, Propagation of chaos for the two-dimensional Navier-Stokes equation, Probabilistic methods in mathematical physics (Katata/Kyoto, 1985), Academic Press, Boston, MA, 1987, pp. 303–334.
- [Rom18] Marco Romito, A simple method for the existence of a density for stochastic evolutions with rough coefficients, Electron. J. Probab. 23 (2018), Paper no. 113, 43.
- [RS23] Matthew Rosenzweig and Sylvia Serfaty, Global-in-time mean-field convergence for singular Riesz-type diffusive flows, Ann. Appl. Probab. 33 (2023), no. 2, 754–798.
- [Ser20] Sylvia Serfaty, Mean field limit for Coulomb-type flows, Duke Math. J. 169 (2020), no. 15, 2887–2935.
- [Ste70] Elias M. Stein, Singular integrals and differentiability properties of functions, Princeton Mathematical Series, No. 30, Princeton University Press, Princeton, N.J., 1970.
- [Szn91] Alain-Sol Sznitman, Topics in propagation of chaos, École d’Été de Probabilités de Saint-Flour XIX—1989, Lecture Notes in Math., vol. 1464, Springer, Berlin, 1991, pp. 165–251.
- [TBL06] Chad M. Topaz, Andrea L. Bertozzi, and Mark A. Lewis, A nonlocal continuum model for biological aggregation, Bull. Math. Biol. 68 (2006), no. 7, 1601–1623.
- [Tri83] Hans Triebel, Theory of function spaces, Monographs in Mathematics, vol. 78, Birkhäuser Verlag, Basel, 1983.
- [Vil09] Cédric Villani, Optimal transport, Grundlehren der mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 338, Springer-Verlag, Berlin, 2009, Old and new.
- [WYW06] Zhuoqun Wu, Jingxue Yin, and Chunpeng Wang, Elliptic & parabolic equations, World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ, 2006.
- [Yos80] Kôsaku Yosida, Functional analysis, sixth ed., Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 123, Springer-Verlag, Berlin-New York, 1980.