1,281
Views
0
CrossRef citations to date
0
Altmetric
Articles

Analysis and mean-field derivation of a porous-medium equation with fractional diffusion

, , &
Pages 2217-2269 | Received 17 Sep 2021, Accepted 25 Aug 2022, Published online: 09 Sep 2022

Abstract

A mean-field-type limit from stochastic moderately interacting many-particle systems with singular Riesz potential is performed, leading to nonlocal porous-medium equations in the whole space. The nonlocality is given by the inverse of a fractional Laplacian, and the limit equation can be interpreted as a transport equation with a fractional pressure. The proof is based on Oelschläger’s approach and a priori estimates for the associated diffusion equations, coming from energy-type and entropy inequalities as well as parabolic regularity. An existence analysis of the fractional porous-medium equation is also provided, based on a careful regularization procedure, new variants of fractional Gagliardo–Nirenberg inequalities, and the div-curl lemma. A consequence of the mean-field limit estimates is the propagation of chaos property.

2000 Mathematics Subject Classification:

1. Introduction

The aim of this article is to derive and analyze the following nonlocal porous-medium equation: (1) tρ=div(ρP),P=(Δ)sf(ρ),ρ(0)=ρ0in Rd,(1) where 0<s<1,d2, and fC1([0,)) is a nondecreasing function satisfying f(0)=0. This model describes a particle system that evolves according to a continuity equation for the density ρ(x,t) with velocity v=P. The velocity is assumed to be the gradient of a potential, which expresses Darcy’s law. The pressure P is related to the density in a nonlinear and nonlocal way through P=(Δ)sf(ρ). The nonlocal operator (Δ)s can be written as a convolution operator with a singular kernel, (2) (Δ)su=K*u,K(x)=cd,s|x|2sd,xRd,(2) where cd,s=Γ(d/2s)/(4sπd/2Γ(s)) and Γ denotes the Gamma function [Citation1, theorem 5].

If s = 0, we recover the porous-medium equation (for nonnegative solutions), while the case s = 1 was investigated in [Citation2, Citation3] with f(u) = u for the evolution of the vortex density in a superconductor. Related models (with f(u) = u) appear in the dynamics of dislocations (line defects) in crystals [Citation4, (1.5)]. Other applications include particle systems with long-range interactions [Citation5, subsection 6.2]. The case 0<s<1 corresponds to long-range repulsive interactions. This model, still with f(u) = u, was investigated in [Citation4], but a mathematical justification is missing. In this article, we provide a rigorous derivation from an interacting particle system for general functions f(u). In this way, we aim to contribute to the understanding of mean-field limits involving nonquadratic nonlinearities.

EquationEquation (1) was first analyzed in [Citation6] with f(u) = u for nonnegative solutions and in [Citation7] with f(u)=|u|m2u (m > 1) for sign-changing solutions. The nonnegative solutions have the interesting property that they propagate with finite speed, which is not common in other fractional diffusion models [Citation6, Citation8]. EquationEquation (1) was probabilistically interpreted in [Citation9], and it was shown that the probability density of a so-called random flight process is given by a Barenblatt-type profile. Previous mean-field limits leading to (Equation1) were concerned with the linear case f(u) = u only; see [Citation10] (using the technique of [Citation11]) and [Citation12] (including additional diffusion as in (Equation7) below). In [Citation13], EquationEq. (1) (with f(u) = u) was derived in the high-force regime from the Euler–Riesz equations, which can be derived in the mean-field limit from interacting particle systems [Citation14]. A direct derivation from particle systems with Lévy noise was proved in [Citation15] for cross-diffusion systems, but still with f(u) = u. Up to our knowledge, a rigorous derivation of (Equation1) from stochastic interacting particle systems for general nonlinearities f(u) like power functions is missing in the literature. In this article, we fill this gap.

1.1. Problem setting

EquationEquation (1) is derived from an interacting particle system with N particles, moving in the whole space Rd. Because of the singularity of the integral kernel and the degeneracy of the nonlinearity, we approximate (Equation1) using three levels. First, we introduce a parabolic regularization adding a Brownian motion to the particle system with diffusivity σ(0,1) and replacing f by a smooth approximation fσ. Second, we replace the interaction kernel K by a smooth kernel Kζ with compact support, where ζ>0. Third, we consider interaction functions Wβ with β(0,1), which approximate the delta distribution. (We refer to Subsection 1.3 for the precise definitions.)

The particle positions are represented on the microscopic level by the stochastic processes XiN(t) evolving according to (3) dXiN(t)=Kζ*fσ(1Nj=1,jiNWβ(XjN(t)XiN(t)))dt+2σdBiN(t),XiN(0)=ξi,i=1,,N,(3) where the convolution has to be understood with respect to xi, (BiN(t))t0 are independent d-dimensional Brownian motions defined on a filtered probability space (Ω,F,Ft,P), and ξi are independent identically distributed random variables in Rd with the same probability density function ρσ0 (defined in (Equation12) below).

The mean-field-type limit is performed in three steps. First, for fixed (σ,β,ζ), system (Equation3) is approximated for N on the intermediate level by (4) dX¯iN(t)=Kζ*fσ(Wβ*ρσ,β,ζ(X¯iN(t),t))dt+2σdBiN(t),X¯iN(0)=ξi,i=1,,N,(4) where ρσ,β,ζ is the probability density function of X¯iN and a strong solution to (5) tρσ,β,ζσΔρσ,β,ζ=div(ρσ,β,ζKζ*fσ(Wβ*ρσ,β,ζ)),ρσ,β,ζ(0)=ρσ0in Rd.(5) System (Equation4) is uncoupled, since X¯iN depends on N only through the initial datum.

Second, passing to the limit (β,ζ)0 in the intermediate system leads on the macroscopic level to (6) dX̂iN(t)=K*fσ(ρσ(X̂iN(t),t))dt+2σdBiN(t),X̂iN(0)=ξi,i=1,,N,(6) where ρσ is the density function of X̂iN and a weak solution to (7) tρσ=σΔρσ+div(ρσ(Δ)sfσ(ρσ)),ρσ(0)=ρσ0in Rd.(7)

We perform the limits N and (β,ζ)0 simultaneously. In this limit, we use the logarithmic scaling β(logN)μ for some μ>0 between the strength of interaction β and the number of particles N. This can be viewed as a moderately interacting particle system. For the smoothing parameter ζ of the singularity from K, we can even allow an algebraic dependence on N, i.e., ζNν for some ν>0; see Theorem 2 for details. Our approach also implies the two-step limit but leading to weak convergence only, compared to the convergence in expectation obtained in Theorem 3.

Third, the limit σ0 is performed on the level of the diffusion equation, based on a priori estimates uniform in σ and the div-curl lemma.

The main result of this article is that the particles of system (Equation3) become independent in the limit with a common density function that is a weak solution to (Equation1)–(Equation2).

1.2. State of the art

We already mentioned that the existence of weak solutions to (Equation1) with f(u) = u was proved first in [Citation6]. The convergence of the weak solution to a self-similar profile was shown by the same authors in [Citation16]. The convergence becomes exponential, at least in one space dimension, when adding a confinement potential [Citation17]. EquationEquation (1) with f(u) = u was identified as the Wasserstein gradient flow of a square fractional Sobolev norm [Citation18], implying time decay as well as energy and entropy estimates. The Hölder regularity of solutions to (Equation1) was proved in [Citation19] for f(u) = u and in [Citation20] for f(u)=um1 and m2.

In the literature, related equations have been analyzed too. EquationEquation (1) for f(u) = u and the limit case s = 1 was shown in [Citation21] to be the Wasserstein gradient flow on the space of probability measures, leading to the well-posedness of the equation and energy-dissipation inequalities. The existence of local smooth solutions to the regularized EquationEq. (Equation7) are proved in [Citation22]. The solutions tρ=div(ρm1P) with P=(Δ)sρ in Rd propagate with finite speed if and only if m2 [Citation8]. The existence of weak solutions to this equation with P=(Δ)s(ρn) and n > 0 is proved in [Citation23] (in bounded domains). While (Equation1) has a parabolic-elliptic structure, parabolic-parabolic systems have been also investigated. For instance, the global existence of weak solutions to tρ=div(ρP) and tP+(Δ)sP=ρβ, where β>1, was shown in [Citation24]. In [Citation25], the algebraic decay toward the steady state was proved in the case β = 2. We also mention that fractional porous-medium equations of the type tρ+(Δ)s/2f(ρ)=0 in Rd have been studied in the literature; see, e.g., [Citation26]. Compared to (Equation1), this problem has infinite speed of propagation. For a review and comparison of this model and (Equation1), we refer to [Citation27].

There is a huge literature concerning mean-field limits leading to diffusion equations, and the research started already in the 1980s; we refer to the reviews [Citation28, Citation29] and the classical works of Sznitman [Citation30, Citation31]. Oelschläger proved the mean-field limit in weakly interacting particle systems [Citation32], leading to deterministic nonlinear processes, and moderately interacting particle systems [Citation33], giving porous-medium-type equations with quadratic diffusion. First, investigations of moderate interactions in stochastic particle systems with nonlinear diffusion coefficients were performed in [Citation34]. The approach of moderate interactions was extended in [Citation35, Citation36] to multispecies systems, deriving population cross-diffusion systems. Reaction-diffusion equations with nonlocal terms were derived in the mean-field limit in [Citation37]. The large population limit of point measure-valued Markov processes leads to nonlocal Lotka–Volterra systems with cross diffusion [Citation38]. Further references can be found in [Citation12, subsection 1.3].

Compared to previous works, we consider a singular kernel K and derive a partial differential equation without Laplace diffusion by taking the limit σ0. The authors of [Citation39] derived the viscous porous-medium equation by starting from a stochastic particle system with a double convolution structure in the drift term, similar to (Equation4). The main difference to our work is that (besides different techniques for the existence and regularity of solutions to the parabolic problems) we consider a singular kernel in one part of the convolution and a different scaling for the approximating regularized kernel Kζ=Kωζ*Wζ, where ωζ is a W1,(Rd) cutoff function (see Subsection 1.3 for the exact definition), in comparison to the interaction scaling Wβ*ρσ,β,ζ. The two different scalings β and ζ allow us to establish a result, for which the kernel regularization on the particle level does not need to be of logarithmic type but of power-law type only.

1.3. Main results and key ideas

We impose the following hypotheses:

(H1) Data: Let 0<s<1,d2.

(H2) ρ0L(Rd)L1(Rd) satisfies ρ00 in Rd and Rdρ0(x)|x|2d/(d2s)dx<.

(H3) Nonlinearity: fC1([0,)) is nondecreasing, f(0)=0, and uuf(u) for u > 0 is strictly convex.

Let us discuss these assumptions. We assume that d2; the case d = 1 can be treated if s < 1∕2; see [Citation6]. Extending the range of s to s < 0 leads to the fractional (higher-order) thin-film equation, which is studied in [Citation40]. The case 1<s<d/2 may be considered too, since it yields better regularity results; we leave the details to the reader. Conversely, the case sd/2 is more delicate since the multiplier in the definition of (Δ)s using Fourier transforms does not define a tempered distribution. The case s=d/2 for d2 (with a logarithmic Riesz kernel) was analyzed in [Citation10]. We need the moment bound for the initial datum ρ0 to prove the same moment bound for ρσ, which in turn is used several times, for instance to show the entropy balance and the convergence ρσρ as σ0 in the sense of Cweak0([0,T];L1(Rd)). The monotonicity of f and the strict convexity of uuf(u) are needed to prove the strong convergence of (ρσ), which then allows us to identify the limit of (fσ(ρσ)). An example of a function satisfying Hypothesis (H3) is f(u)=uβ with β1.

Our first result is concerned with the existence analysis of (Equation1). This result is needed to prove the main theorem below. We write ·p for the Lp(Rd) norm and define the so-called entropy density h:[0,)R by h(u)=0u1vf(w)wdwdvfor u0.

Theorem 1

(Existence of weak solutions to (Equation1)). Let Hypotheses (H1)–(H3) hold. Then, there exists a weak solution ρ0 to (Equation1) satisfying (i) the regularityρL(0,;L1(Rd)L(Rd)),(Δ)s/2f(ρ)L2(0,;L2(Rd)),tρL2(0,;H1(Rd)), (ii) the weak formulation(8) 0Ttρ,ϕdt+0TRdρ(Δ)sf(ρ)·ϕdxdt=0(8) for all ϕL2(0,T;H1(Rd)) and T > 0, (iii) the initial datum ρ(0)=ρ0 in the sense of H1(Rd), and (iv) the following properties for t > 0:

  • Mass conservation: ρ(t)1=ρ01,

  • Dissipation of the L norm: ρ(t)ρ0,

  • Moment estimate: sup0<t<TRdρ(x,t)|x|2d/(d2s)dxC(T),

  • Entropy inequality:Rdh(ρ(t))dx+0tRd|(Δ)s/2f(ρ)|2dxdsRdh(ρ0)dx.

Note that the Hardy–Littlewood–Sobolev-type inequality (B3) (see Appendix B) implies that ρ(Δ)sf(ρ)2=ρ(Δ)s/2[(Δ)s/2f(ρ)]2Cρd/(2s)(Δ)s/2f(ρ)2, such that ρ(Δ)sf(ρ)L2(Rd), and the weak formulation (Equation8) is well defined.

The main ideas of the proof of Theorem 1 are as follows. A priori estimates for strong solutions ρσ to the regularized EquationEq. (Equation7) are derived from mass conservation, the entropy inequality, and energy-type bounds. The energy-type bound allows us to show, for sufficiently small σ>0, that the L norm of ρσ is bounded by the L norm of ρ0, up to some factor depending on the moment bound for ρ0. The existence of a strong solution ρσ is proved by regularizing (Equation7) in a careful way to deal with the singular kernel. The regularized equation is solved locally in time by Banach’s fixed-point theorem. Entropy estimates allow us to extend this solution globally in time and to pass to the de-regularization limit. The second step is the limit σ0 in (Equation7). Since the bounds only provide weak convergence of (a subsequence of) ρσ, the main difficulty is the identification of the nonlinear limit fσ(ρσ). This is done by applying the div-curl lemma and exploiting the monotonicity of f and the strict convexity of uuf(u) [Citation41].

We already mentioned that the existence of local smooth solutions ρσ to (Equation7) has been proven in [Citation13]. However, we provide an independent proof that allows for global strong solutions and that yields a priori estimates needed in the mean-field limit.

Our second and main result is the propagation of chaos, which shows a mean-field-type convergence of the particle system (Equation3) to a solution of (Equation1). To define our particle system properly, we need some definitions. Introduce the smooth approximation (9) fσ(u)=0u(Γσ*(f1[0,)))(w)Ξ˜(σw)dw uR,(9) where the mollifier Γσ for σ>0 is given by Γσ(x)=σ1Γ1(x/σ), and Γ1C0(R) satisfies Γ10,Γ11=1, while the cutoff function Ξ˜C0(R) satisfies 0Ξ˜1 in R and Ξ˜(x)=1 for |x|1. Then, thanks to Γσ, we have fσC(R). The cutoff function guarantees that the derivatives Dkfσ are bounded and compactly supported for all k1. Furthermore, it holds that fσ0,fσ(0)=0.

In a similar way, we introduce the mollifier function Wβ for β>0 and xRd by (10) Wβ(x)=βdW1(x/β),W1C0(Rd) is symmetric,W10,W11=1.(10) Let us define the cutoff version of the singular kernel K by (11) K˜ζ:=Kωζ, where the cutoff function ωζW1,(Rd) is such that0ωζ(x)1 for xRd,ωζ2ζ,ωζ(x)=1 for all |x|ζ1,ωζ(x)=0 for all |x|2ζ1.(11) Then, the regularized kernel Kζ is given by Kζ(x):=K˜ζ*Wζ(x)for all xRd, where ζ>0. Let the cutoff function ΞC0(Rd) satisfy 0Ξ1 in Rd and Ξ(x)=1 for |x|1. Then, we define the regularized initial datum for xRd by (12) ρσ0(x)=κσ(Wσ*ρ0)(x)Ξ(σx),where κσ=Rdρ0(y)dyRd(Wσ*ρ0)(y)Ξ(σy)dy.(12) This definition guarantees the mass conservation since ρσ01=ρ01; see Subsection 2.1.

Note that our particle system (Equation3) depends on 4 parameters: NN denotes the number of particles, β>0 models the strength of interaction between the particles, ζ>0 describes the regularization of the singular kernel K, and σ>0 is a measure of the additional diffusion. The quantities (β,ζ,σ) are regularization parameters needed to overcome the singularity of the kernel K and the (possible) degeneracy of the nonlinearity f.

In the limit N,(β,ζ,σ)0, we prove the following propagation-of-chaos result.

Theorem 2

(Propagation of chaos). Let ζ2s1C1N1/4 and β3d7εlogN for some constants C1, ε>0, and let PN,σ,β,ζk(t) be the joint distribution of (X1N(t),,XkN(t)) for k1 and t(0,T). Then, there exists a subsequence in σ such thatlimσ0limN,(β,ζ)0PN,σ,β,ζk(t)=Pk(t)in the sense of distributions,where the limit is locally uniform in t, and the measure P(t) is absolutely continuous with respect to the Lebesgue measure with the probability density function ρ(t) that is a weak solution to (Equation1).

It is well known (see, e.g., [Citation31]) that the result of Theorem 2 implies the weak convergence of the empirical measure associated to the particle system (Equation3) toward the deterministic measure ρ(t), i.e., μN,σ,β,ζ(t)=i=1NδXiN(t)ρ(t), for a subsequence in σ. Furthermore, Theorem 2 shows that at any time t > 0, in the limit N,(β,ζ,σ)0, any finite selection of k particles in (Equation3) becomes independent with limiting distribution ρk(t). If EquationEq. (1) was uniquely solvable, we would obtain the convergence of the whole sequence in σ. Unfortunately, the regularity of the solution ρ to (Equation1) is too weak to conclude the uniqueness of weak solutions. Up to our knowledge, none of the known methods, such as [Citation42, Citation43], seem to be applicable to EquationEq. (Equation1).

Theorem 2 is proved in two steps: First, we show (strong) error estimates between particles of systems (Equation3) and (Equation6), respectively; see Proposition 3 below. Second, we show the weak convergence of (a subsequence of) ρσ to a solution ρ to (Equation1); see Corollary 13.

Proposition 3

(Error estimate for the stochastic system). Let XiN and X̂iN be the solutions to (Equation3) and (Equation6), respectively. We assume that ζ2s1C1N1/4 for some constant C1>0. Let δ(0,1/4) and a:=min{1,d2s}>0. Then, there exist constants ε>0, depending on σ and δ, and C2>0, depending on σ and T, such that if β3d7εlogN then,E(sup0<s<Tmaxi=1,,N|(XiNX̂iN)(s)|)C2(β+ζa)0 as (N,ζ,β)(,0,0).

The proposition is proved by estimating the differences E1(t):=E(sup0<s<tmaxi=1,,N|(XiNX¯iN)(s)|),E2(t):=E(sup0<s<tmaxi=1,,N|(X¯iNX̂iN)(s)|), and applying the triangle inequality. For the first difference, we estimate expressions like DkKζ*u for appropriate functions u and DkWβ for kN in terms of negative powers of β (here, Dk denotes the kth-order partial derivatives). Using properties of Riesz potentials, in particular Hardy–Littlewood–Sobolev-type inequalites (see Lemmas 22 and 23), we show that for some μi>0 (i = 1, 2, 3), E1(t)C(σ)βμ10tE1(s)ds+C(σ)βμ2ζμ3N1/2. By applying the Gronwall lemma and choosing a logarithmic scaling for β and an algebraic scaling for ζ with respect to N, we infer that E1(t)C(σ)Nμ4 for some μ4(0,1/4). For the second difference E2, we need the estimates Wβ*uuC(σ)β (Lemma 21), and (KζK)*ρσC(σ)ζa, ρσ,β,ζρσC(σ)(β+ζa) (Proposition 14), recalling that a=min{1,d2s}. The proof of these estimates is very technical. The idea is to apply several times fractional Gagliardo–Nirenberg inequalities that are proved in Appendix B and Hardy–Littlewood–Sobolev inequalities that are recalled in Lemmas 22 and 23. Then, after suitable computations, E2(t)C(σ)(β+ζa)+C(σ)0tE2(s)ds, and we conclude with Gronwall’s lemma that E2(t)C(σ)(β+ζa).

The article is organized as follows. The existence of global nonnegative weak solutions to (Equation1) is proved in Section 2 by establishing an existence analysis for (Equation7) and performing the limit σ0. Some uniform estimates for the solution ρρ,β,ζ to (Equation5) and for the difference ρσ,β,ζρσ are shown in Section 3. Section 4 is devoted to the proof of the error estimate in Theorem 3 and the propagation of chaos in Theorem 2. In Appendices A–C we recall some auxiliary results and Hardy–Littlewood–Sobolev-type inequalities, prove new variants of fractional Gagliardo–Nirenberg inequalities, and formulate a result on parabolic regularity.

1.3.1. Notation

We write ·p for the Lp(Rd) or Lp(R) norm with 1p. The ball around the origin with radius R > 0 is denoted by BR. The partial derivative /xi is abbreviated as i for i=1,,d, and Dα denotes a partial derivative of order |α|, where αN0d is a multi-index. The notation Dk refers to the kth-order tensor of partial derivatives of order kN. In this situation, the norm Dkup is the sum of all Lp norms of partial derivatives of u of order k. Finally, C > 0, C1>0, etc. denote generic constants with values changing from line to line.

2. Analysis of EquationEq. (Equation1)

In this section, we prove the existence of global nonnegative weak solutions to (Equation1). We first prove the existence of a solution ρσ to (Equation7) by a fixed-point argument, and then, perform the limit σ0. In Subsection 2.1, we prove some basic estimates for a strong solution ρσ to (Equation7). Entropy and moment estimates as well as higher-order estimates are derived in Subsections 2.2 and 2.3, respectively. The existence of a unique strong solution to (Equation7) is proved in Subsection 2.4 using a regularized version of (Equation7) and Banach’s fixed-point theorem. The strong L1(Rd) limit σ0 is performed in Subsection 2.5 using the div-curl lemma. Finally, Subsection 2.6 is concerned with the proof of a time-uniform weak L1(Rd) limit of (ρσ), which is needed in the proof of Proposition 3. Recall definition (Equation12) of the number κσ, which is stated in (iv) below.

Proposition 4.

Let Hypotheses (H1)–(H3) hold. Then, for all σ>0, there exists a unique weak solution ρσ0 to (Equation7) satisfying (i) the regularityρσL(0,;L1(Rd)L(Rd))C0([0,);L2(Rd)),ρσL2(0,;L2(Rd)),tρσL2(0,;H1(Rd)), (ii) the weak formulation of (Equation7) with test functions ϕL2(0,T;H1(Rd)), (iii) the inital datum ρσ(0)=ρσ0 in L2(Rd), and (vi) the following properties for t > 0, which are uniform in σ for sufficiently small σ>0:

  • Mass conservation: ρσ(t)1=ρ01.

  • Dissipation of the L norm: ρσL(0,;L(Rd))κσρ0L(Rd)Cρ0L(Rd).

  • Moment estimate: supt[0,)Rdρσ(x,t)|x|2dd2sdxCT.

  • Entropy inequality:Rdh(ρσ(T))dx+4σ0TRdfσ(ρσ)|ρσ|2dxdt+0TRd|(Δ)s/2fσ(ρσ)|2dxdtRdh(ρσ0)dxfor all T>0.

Additionally, for any T > 0, 1<p<, and 2q<, there exists C > 0, depending on T, σ, p, and q, such thatρσLp(0,T;W3,p(Rd))+tρσLp(0,T;W1,p(Rd))+ρσC0([0,T];W2,1(Rd)W3,q(Rd))C,i.e., ρσ is even a strong solution to (Equation7) and ρσC0([0,T];W2,1(Rd)W3,q(Rd)) for q2.

2.1. Basic estimates for ρσ

We prove a priori estimates in Lp spaces and an energy-type estimate. Let σ(0,1) and let ρσ be a nonnegative strong solution to (Equation7). Integration of (Equation7) in Rd and the definition of ρσ0 immediately yield the mass conservation (13) ρσ(t)1=ρσ01=ρ01for t>0.(13)

Lemma 5

(Energy-type estimate). Let FC2([0,)) be convex and let F(ρσ0)L1(Rd). Then,(14) ddtRdF(ρσ)dx=σRdF(ρσ)|ρσ|2dxcd,1s2RdRd(G(ρσ(x))G(ρσ(y)))(fσ(ρσ(x))fσ(ρσ(y)))|xy|d+2(1s)dxdy0,(14) where G(u):=0uvF(v)dv for u0 and cd,1s is defined after (Equation2).

Proof.

First, we assume that F is additionally bounded. Then, F(ρσ)F(0) is an admissible test function in the weak formulation of (Equation7), since |F(ρσ)F(0)|F|ρσ|. It follows from definition (B1) of the fractional Laplacian and integration by parts that ddtRdF(ρσ)dx+σRdF(ρσ)|ρσ|2dx=RdF(ρσ)ρσρσ·(Δ)sfσ(ρσ)dx=RdG(ρσ)·(Δ)sfσ(ρσ)dx=RdG(ρσ)(Δ)1sfσ(ρσ)dx=cd,1sRdRdG(ρσ(x))fσ(ρσ(x))fσ(ρσ(y))|xy|d+2(1s)dxdy. A symmetrization of the last integral yields (Equation14).

In the general case, we introduce Fk(u)=F(0)+F(0)u+0u0vmin{F(w),k}dwdv for k > 0. Then, Fk(u) is bounded and (Equation14) follows for F replaced by Fk. The result follows after taking the limit k using monotone convergence. □

We need a bound on κσ, defined in (Equation12), to derive uniform L(Rd) bounds for ρσ.

Lemma 6

(Bound for κσ). There exists C > 0 such that, for sufficiently small σ>0, 1κσ11CσE,where E:=1ρ01Rd(1+|x|2d/(d2s))ρ0(x)dx.

Proof.

By Young’s convolution inequality (Lemma 19), we have Rd(Wσ*ρ0)(x)Ξ(σx)dxWσ*ρ01Wσ1ρ01=ρ01, which shows that κσ1. To prove the upper bound, we use the triangle inequality |x||xy|+|y|: Rd(Wσ*ρ0)(x)Ξ(σx)dx{|x|1/σ}RdWσ(xy)ρ0(y)dydx=Rd(RdWσ(xy)dx)ρ0(y)dy{|x|>1/σ}RdWσ(xy)ρ0(y)dydxRdρ0(y)dyσ2d/(d2s){|x|>1/σ}Rd|x|2d/(d2s)Wσ(xy)ρ0(y)dydxRdρ0(y)dyσ2d/(d2s)RdRd|xy|2d/(d2s)Wσ(xy)ρ0(y)dydxσ2d/(d2s)RdRd|y|2d/(d2s)Wσ(xy)ρ0(y)dydx. Using the property Rd|z|2d/(d2s)Wσ(z)dzCσ2d/(d2s) for the second term on the right-hand side and WβL1(Rd)=1 for the third term, we find that Rd(Wσ*ρ0)(x)Ξ(σx)dxRdρ0(y)dyCσ4d/(d2s)Rdρ0(y)dyσ2d/(d2s)Rd|y|2d/(d2s)ρ0(y)dy. Because of σ2d/(d2s)σ for σ1, we obtain ρ01κσ=Rd(Wσ*ρ0)(x)Ξ(σx)dxRdρ0(y)dyCσRd(1+|y|2d/(d2s))ρ0(y)dyRdρ0(y)dyCσRdρ0(y)dy·E=ρ01(1CσE), which proves the lemma. □

Lemma 7

(Bounds for ρσ). The following bounds hold:(15) ρσ(t)κσρ0Cρ0,t>0,(15) (16) σρσL2(0,T;H1(Rd))ρ02,(16) where (Equation15) holds for sufficiently small σ>0.

Lemma 7 and mass conservation imply that ρσ(t)p is bounded for all t > 0 and 1p. Observe that κσ1 as σ0. So, if ρσ(t)ρ(t) a.e., the dissipation of the L norm follows, as stated in Theorem 1 (iv).

Proof.

The convexity of F shows that G, defined in Lemma 5, is nondecreasing. Therefore, (d/dt)RdF(ρσ)dx0 and supt>0RdF(ρσ(t))dxRdF(ρσ0)dx. We choose a convex function FC2([0,)) such that F(u) = 0 for uρσ0, F(u) > 0 for u>ρσ0 and satisfying F(u)Cu for u. Then, 0RdF(ρσ(t))dxRdF(ρσ0)dx=0for t>0. Consequently, ρσ(x,t)ρσ0κσρ0 for t > 0, showing the L(Rd) bound. Finally, choosing F(u)=u2 in Lemma 5, the L2(0,T;H1(Rd)) estimate follows. □

2.2. Entropy and moment estimates

We need a fractional derivative estimate for fσ(ρσ), which is not an immediate consequence of Lemma 5. To this end, we define the entropy density hσ(u)=0u1vfσ(w)wdwdv,u0.

Lemma 8

(Entropy balance). It holds for all t > 0 thatddtRdhσ(ρσ)dx+4σRdfσ(ρσ)|ρσ1/2|2dx+Rd|(Δ)s/2fσ(ρσ)|2dx=0.In particular, for all T > 0, there exists C > 0 such that(17) fσ(ρσ)L2(0,T;H1s(Rd))C.(17)

Proof.

The idea is to apply Lemma 5. Since hσC2([0,)), we cannot use the lemma directly. Instead, we apply it to the regularized function hσδ(u)=0u1vfσ(w)w+δdwdv,u0, where δ>0. Choosing F=hσδ in Lemma 5 gives (18) ddtRdhσδ(ρσ)dx+4σRdfσ(ρσ)ρσρσ+δ|ρσ1/2|2dx=cd,1s2RdRd(fσδ(ρσ(x))fσδ(ρσ(y)))(fσ(ρσ(x))fσ(ρσ(y))|xy|d+2(1s)dxdy,(18) where fσδ(u):=0u(v/(v+δ))fσ(v)dv for u0.

Step 1: Estimate of hσδ. The pointwise limit hσδ(ρσ)hσ(ρσ) holds a.e. in Rd×(0,T) as δ0. We observe that for all 0<u1, |hσδ(u)|sup0<v<1f(v)0uv1dwwdvCu(|logu|+1), while for all u > 1, since fσ0 in [0,), |hσδ(u)|01v1fσ(w)w+δdwdv+1u1vfσ(w)w+δdwdvC+1u1vfσ(w)dwdvC+0ufσ(v)dvC+ufσ(u). The last inequality follows after integration of fσ(v)fσ(v)+vfσ(v)=(vfσ(v)) in (0,u). Therefore, since ρσρσ0 a.e. in Rd×(0,), we find that |hσδ(ρσ)|Cρσ(|logρσ|+1)1{ρσ1}+C1{ρσ>1}C(ρσθ+ρσ), where θ(0,1) is arbitrary, and consequently, because of mass conservation, (19) Rd|hσδ(ρσ)|dxC+CRdρσθdx.(19)

Step 2: Estimate of Rdρσθdx. Let 0<α<1 and d/(d+α)<θ<1. Then, by Young’s inequality, Rdρσθdx=Rd(1+|x|2)αθ/2ρσθ(1+|x|2)αθ/2dxRd(1+|x|2)α/2ρσdx+CRd(1+|x|2)αθ/(2(1θ))dxRd(1+|x|2)α/2ρσdx+C, since the choice of θ guarantees that αθ/(2(1θ))<d/2, so Rd(1+|x|2)αθ/(2(1θ))dx<. To control the right-hand side, we need to bound a suitable moment of ρσ.

For this, we use the test function (1+|x|2)α/2ξk in the weak formulation of (Equation7), where ξkC02(Rd) is a cutoff function with the properties ξk(x)=1for |x|k,ξk(x)=0for |x|2k,k|ξk(x)|+k2|Δξk(x)|C,0ξk(x)1for xRd, and k > 1 is arbitrary. We find that Rd(1+|x|2)α/2ξkρσ(t)dx=Rd(1+|x|2)α/2ξkρσ0dx+σ0tRdρσξkΔ(1+|x|2)α/2dxds+σ0tRdρσ(2[(1+|x|2)α/2]·ξk+(1+|x|2)α/2Δξk)dxdsα0tRdρσξk(1+|x|2)α/21x·(Δ)sfσ(ρσ)dxds0tRdρσ(1+|x|2)α/2ξk·(Δ)sfσ(ρσ)dxds. Since α<1 and 0ξk1 in Rd, the terms involving Δ(1+|x|2)α/2 and x(1+|x|2)α/21 are bounded in Rd. It follows from the choice of ξk that |[(1+|x|2)α/2]·ξk|+|(1+|x|2)α/2Δξk|Ckα2,(1+|x|2)α/2|ξk|Ckα1. Thus, taking into account the assumption on ρ0 and mass conservation, sup0<t<TRd(1+|x|2)α/2ξkρσ(t)dxC+C0TRdρσ(Δ)s/2|(Δ)s/2fσ(ρσ)|dxdt. Next, we apply the Hardy–Littlewood–Sobolev inequality (see Lemma 22) and the Hölder inequality and use the fact that ρσ(t) is bounded in any Lp(Rd): sup0<t<TRd(1+|x|2)α/2ρσ(x,t)ξk(x)dxC+C0Tρσ2d/(d+2s)(Δ)s/2[(Δ)s/2fσ(ρσ)]2d/(d2s)dtC+0Tρσ2d/(d+2s)(Δ)s/2fσ(ρσ)2dtC(η)+η0T(Δ)s/2fσ(ρσ)22dt for all η>0. Since ξk(x)ξk+1(x) for xRd, k > 1, and ξk1 a.e. in Rd as k, we deduce from monotone convergence that in the limit k, sup0<t<TRd(1+|x|2)α/2ρσ(x,t)dxC(η)+η0T(Δ)s/2fσ(ρσ)22dt for all η>0. This proves that RdρσθdxC(η)+η0T(Δ)s/2fσ(ρσ)22dt.

Step 3: A priori estimate. Inserting the previous estimate into (Equation19) leads to sup0<t<TRd|hσδ(ρσ(x,t))|dxC(η)+η0T(Δ)s/2fσ(ρσ)22dt. We integrate (Equation18) in time and use the previous estimate: 4σ0TRdfσ(ρσ)ρσρσ+δ|ρσ1/2|2dxdt+cd,1s20TRdRd(fσδ(ρσ(x))fσδ(ρσ(y)))(fσ(ρσ(x))fσ(ρσ(y))|xy|d+2(1s)dxdydtRd|hσδ(ρσ(T))|dx+Rd|hσδ(ρσ0)|dxC(η)+η0T(Δ)s/2fσ(ρσ)22dt. We wish to pass to the limit δ0 in the previous inequality. We deduce from dominated convergence that fσδ(ρσ)fσ(ρσ) a.e. in Rd×[0,). The integrand of the second term on the left-hand side is nonnegative, and we obtain from Fatou’s lemma that (20) 4σ0TRdfσ(ρσ)|ρσ1/2|2dxdt+cd,1s20TRdRd(fσ(ρσ(x))fσ(ρσ(y)))2|xy|d+2(1s)dxdydtC(η)+η0T(Δ)s/2fσ(ρσ)22dt.(20) By the integral representation of the fractional Laplacian, cd,1s2RdRd(fσ(ρσ(x))fσ(ρσ(y)))2|xy|d+2(1s)dxdy=(Δ)s/2fσ(ρσ)22, the last term in (Equation20) can be absorbed for sufficiently small η>0 by the second term on the left-hand side. This leads to the estimate 4σ0TRdfσ(ρσ)|ρσ1/2|2dxdt+0TRd|(Δ)s/2fσ(ρσ)|2dxdtC. Thus, we can pass to the limit δ0 in (Equation18) giving the desired entropy balance. Finally, bound (Equation17) follows from the definition of the H1s(Rd) norm and the facts that fσ(ρσ)L2(Rd) since fσ is locally Lipschitz continuous, fσ(0)=0, and ρσ is bounded both in L(Rd) and L2(Rd) independently of σ. □

Lemma 9

(Moment estimate). It holds thatsup0<t<TRdρσ(x,t)|x|2d/(d2s)dxC,where C > 0 depends on T and the L1(Rd) norms of ρ0 and |·|2d/(d2s)ρ0.

Proof.

For the following computations, we would need to use cutoff functions to make the calculations rigorous. We leave the details to the reader, as we wish to simplify the presentation. Let m=2d/(d2s). Since |·|mρ0L1(Rd) by assumption, we can compute (21) ddtRdρσ(t)|x|mmdx=σ(m2+d)Rd|x|m2ρσdxRdρσ|x|m2x·(Δ)sfσ(ρσ)dxC|·|m2ρσ1+|·|m1ρσ2d/(d+2s)(Δ)sfσ(ρσ)2d/(d2s).(21) By Young’s inequality and mass conservation, we have |·|m2ρσ1CRd(1+|x|m)ρσdxC+CRd|x|mρσdx. It follows from (Equation17) that (Δ)sfσ(ρσ) is bounded in L2(0,T;Hs(Rd)). In particular, because of the Sobolev embedding Hs(Rd)Lm(Rd), (Δ)sfσ(ρσ)L2(0,T;Lm(Rd))C. Furthermore, using ρσL(0,;L(Rd)), Young’s inequality, and the property 2d/(d+2s)1 (recall that d2) |·|m1ρσ2d/(d+2s)2d/(d+2s)=Rdρσ2d/(d+2s)|x|2d(m1)/(d+2s)dxC+CRdρσ|x|2d(m1)/(d+2s)dx. Thus, we infer from (Equation21) and the identity 2d(m1)/(d+2s)=m that ddtRdρσ(t)|x|mmdxC+CRdρσ(t)|x|mdx, and Gronwall’s lemma concludes the proof. □

2.3. Higher-order estimate

We need some estimates in higher-order Sobolev spaces.

Proposition 10

(Higher-order regularity). Let T > 0, 1<p<, and 2q<. Then, there exists C > 0, depending on T, σ, p, and q, such thatρσLp(0,T;W3,p(Rd))+tρσLp(0,T;W1,p(Rd))+ρσC0([0,T];W2,q(Rd))C.

Proof.

Step 1: Case s > 1∕2. If s > 1∕2, then, w:=ρσ(Δ)sfσ(ρσ) does not involve any derivative of ρσ. Thus, wLp(0,T;Lp(Rd)) for p< and Lemma 26 in Appendix C implies that ρσLp(0,T;W1,p(Rd)). Iterating the argument leads to the conclusion. Thus, in the following, we can assume that 0<s1/2.

Step 2: Estimate of div w in Lp(0,T;W1,p(Rd)). We claim that w can be estimated in Lp(0,T;Lp(Rd)) for any p<. Then, by Lemma 26, ρσLp(0,T;Lp(Rd)). We use the L bound for ρσ, the fractional Gagliardo–Nirenberg inequality (Lemma 24), and Young’s inequality to find that wpC(Δ)sfσ(ρσ)pCfσ(ρσ)p2sfσ(ρσ)p12sC(η)+ηρσp, where η>0 is arbitrary. By estimate (C2) in Lemma 26, ρσ(Δ)sfσ(ρσ)p=wpC(η)+η(ρσ(Δ)sfσ(ρσ)p+T1/pρ0p). Choosing η>0 sufficiently small shows the claim.

Step 3: Estimate of div w in Lp(0,T;Lp(Rd)). We use Hölder’s inequality with 1/p=2s/(d+p)+1/q to obtain div wpρσ·(Δ)sfσ(ρσ)p+ρσ(Δ)1sfσ(ρσ)pρσ(d+p)/(2s)(Δ)sfσ(ρσ)q+C(Δ)1sfσ(ρσ)p. By the fractional Gagliardo–Nirenberg inequality (Lemma 25 with θ=1+d/pd/q2s and Lemma 24 with s replaced by 1s) and Young’s inequality, it follows that div wpCρσ(d+p)/(2s)fσ(ρσ)p1θfσ(ρσ)pθ+Cfσ(ρσ)psD2fσ(ρσ)p1sCρσ(d+p)/(2s)ρσpθ+Cfσ(ρσ)D2ρσ+fσ(ρσ)ρσρσp1sC(η)+Cρσ(d+p)/(2s)1/(1θ)+Cρσp+Cρσ2p2+ηD2ρσp, where η>0 is arbitrary. Taking the Lp(0,T) norm of the previous inequality and observing that p/(1θ)=(d+p)/(2s) (because of θ=d(1/p1/q)+12s), it follows that div wLp(0,T;Lp(Rd))C+CρσL(d+p)/(2s)(0,T;L(d+p)/(2s)(Rd))1/(1θ)+CρσLp(0,T;Lp(Rd))+CρσL2p(0,T;L2p(Rd))2+ηD2ρσLp(0,T;Lp(Rd)). Lemma 26 and Step 2 (ρσLp(0,T;Lp(Rd))) show that tρσLp(0,T;Lp(Rd))+(1Cη)D2ρσLp(0,T;Lp(Rd))C. Choosing η>0 sufficiently small, this yields tρσLp(0,T;Lp(Rd)) and ρσLp(0,T; W2,p(Rd)). We deduce from Lemma 20, applied to ρσ, that ρσL(0,T;Lq(Rd)) for any 2q<. (At this point, we need the restriction q2.)

Step 4: Higher-order regularity. To improve the regularity of ρσ, we differentiate (Equation7) in space. Recall that i=/xi,i=1,,d. Then, (22) tiρσσΔiρσ=j=1dij(ρσj(Δ)sfσ(ρσ))=j=1d(ij2ρσj(Δ)sfσ(ρσ)+iρσjj2(Δ)sfσ(ρσ)+jρσij2(Δ)sfσ(ρσ)+ρσijj3(Δ)sfσ(ρσ)).(22) We estimate the right-hand side term by term. Let 0<s1/2. First, by Hölder’s inequality with 1/p=1/q+1/r,1<p<q<,max{2,p}<r< and the fractional Gagliardo–Nirenberg inequality (Lemma 24), ij2ρσj(Δ)sfσ(ρσ)Lp(0,T;Lp(Rd))p0T||ij2ρσqpj(Δ)sfσ(ρσ)rpdtC0Tij2ρσqpfσ(ρσ)r(12s)pfσ(ρσ)r2spdtCfσ(ρσ)L(0,T;Lr(Rd))(12s)pfσ(ρσ)L(0,Lr(Rd))2sp0T||ij2ρσqpdtC. The second and third term on the right-hand side of (Equation22) can be treated in a similar way, observing that ij2(Δ)s=j(Δ)si. The last term is estimated according to ρσijj3(Δ)sfσ(ρσ)pCijj3(Δ)sfσ(ρσ)pCjj2fσ(ρσ)p2sjj2fσ(ρσ)p12sC(η)jj2fσ(ρσ)p+ηjj2fσ(ρσ)p, and the last expression can be absorbed by the corresponding estimate of Δiρσ from the left-hand side of (Equation22). Then, we deduce from Lemma 26 that tiρσ,ijj3ρσLp(0,T; Lp(Rd)) for all p > 1 and Lemma 20, applied to ij2ρσ, yields ij2ρσC0([0,T];Lq(Rd)) for all q2.

Next, if 1/2<s<1, we use the second inequality in Lemma 24 and argue similarly as before. This finishes the proof. □

Lemma 11.

Under the assumptions of Proposition 10, for every q2, there exists a constant C=C(q)>0, depending on σ, such thatρσC0([0,T];W2,1(Rd)W3,q(Rd))C.

The embedding W3,q(Rd)W2,(Rd) for q > d yields a bound for ρσ in C0([0,T]; W2,(Rd)).

Proof.

We first prove the bound in C0([0,T];W3,q(Rd)). By differentiating (Equation7) twice in space, estimating similarly as in Step 4 of the previous proof, and using the regularity results of Proposition 10, we can show that ρσ is bounded in L(0,T;W3,q(Rd)) for any q2.

It remains to show the C0([0,T];W2,1(Rd)) bound for ρσ. In view of mass conservation and Gagliardo–Nirenberg–Sobolev’s inequality, it suffices to show a bound for D2ρσ in L(0,T;L1(Rd)). To this end, we define the weights γn=(1+|x|2)n/2 for n0 and test EquationEq. (Equation7) for ρσ with vn:=γnρσ. Then, tvnσΔvn=div (vnK*fσ(ρσ))+In,vn(0)=γnρσ0in Rd,where In=2σγn·ρσσρσΔγnρσγn·K*fσ(ρσ). Arguing as in Step 4 of the previous proof, we can find a bound in L(0,T;W2,p(Rd)) for vn. Indeed, we can proceed by induction over n, since the additional terms in In can be controlled by Sobolev norms of v0,…,vn−1. The definition of ρσ0 implies that γnρσ0, γnρσ0L(Rd)L1(Rd) for every n0. Then, choosing n > d yields, for 0tT, that γnD2ρσpD2(γnρσ)p+2γn·ρσp+ρσD2γnpC(T). We conclude from γn1L(Rd)L1(Rd) that D2ρσ1γn1p/(p1)γnD2ρσpC(T). This proves the desired bound. □

2.4. Existence of solutions to (Equation7)

We show that the regularized EquationEq. (Equation7) possesses a unique strong solution ρσ.

Step 1: Existence for an approximated system. Let T > 0 arbitrary, define the spaces XT:=L2(0,T;H1(Rd))H1(0,T;H1(Rd))YT:=C0([0,T];L2(Rd)),YT,R:={uYT:uρσ0L(0,T;L2(Rd))R}, and consider the mapping S:vYTuYT, (23) tuσΔu=div (uKs(δ)*fσ(η)(v))in Rd×(0,T),u(0)=ρσ0in Rd,(23) where Ks(δ):RdR+ is a regularized version of Ks, defined by Ks(δ)=K˜s/2(δ)*K˜s/2(δ),K˜s/2(δ)(x)=cd,s/2{δsd+(sd)δsd1(|x|δ)for |x|<δ,|x|sdfor δ|x|δ1,[δds+(sd)δd+1s(|x|δ1)]+for |x|>δ1, and fσ(η) is given by fσ(η)(ρ)=0|ρ|fσ(u)min(1,uη1)du+η2ρ2,ρR. The regularization with parameter η is needed for the entropy estimates.

Lemma 12.

For any 0<s<1 and a.e. xRd, the function δK˜s/2(δ)(x) is nonincreasing for δ(0,1).

Proof.

Let rδ*=(ds+1)/((ds)δ). We can write K˜s/2(δ)(x)=cd,s/2Φδ(|x|) with Φδ(r)={δsd+(sd)δsd1(rδ)for r<δ,rsdfor δrδ1,δds+(sd)δd+1s(rδ1)for δ1<r<rδ*,0for rrδ*. Then, ΦδC0([0,))C1(0,rδ*), and its derivative equals Φδ(r)={(ds)max{r,δ}sd10r1,(ds)min{r,δ1}sd11r<rδ*. We show that Φδ(r) is nonincreasing in δ(0,1) for r1. We have for 1r<rδ*, Φδ(r)=Φδ(1)+1rΦδ(u)du=1(ds)1rmin{u,δ1}sd1du. Furthermore, we have Φδ(rδ*)=0, while min(u,δ1)=δ1>0 for u>rδ*, so it holds that Φδ(r)=(1(ds)1rmin{u,δ1}sd1du)+for r1. At this point, the above representation formula together with elementary monotonicity considerations show that Φδ(r) is nonincreasing in δ(0,1) for r1. It remains to show that Φδ(r) is nonincreasing in δ(0,1) for 0r<1. It holds that Φδ(r)=Φδ(1)r1Φδ(u)du=1+(ds)r1max{u,δ}sd1dufor 0r<1. Once again, we conclude from the above representation formula together with elementary monotonicity considerations that Φδ(r) is nonincreasing in δ(0,1) for 0r<1. This finishes the proof. □

We derive some estimates for fσ(η). First, we have 0fσ(η)(ρ)Cηρ2 for ρR, since fσ(η)(ρ)(η+η1max[0,η]fσ)ρ22for |ρ|η,fσ(η)(ρ)fσ(|ρ|)+η2ρ2(fση2+η2)ρ2for |ρ|>η. Furthermore, |Dfσ(η)(ρ)|=|ρ|ρ|fσ(|ρ|)min(1,|ρ|η1)+ηρ|(η+η1fσ)|ρ|, which implies that |Dfσ(η)(ρ)|Cη|ρ| for ρR. This shows that there exists C(η)>0 such that for any ρ1, ρ2R, |fσ(η)(ρ1)fσ(η)(ρ2)|C(η)(|ρ1|+|ρ2|)|ρ1ρ2|. It follows that fσ(η)(v)L(0,T;L1(Rd)) for vYT.

Since Ks(δ)L(Rd), a standard argument shows that (Equation23) has a unique solution uXTYT. Therefore, the mapping S is well-defined. Additionally, the nonnegativity of u follows immediately after by testing (Equation23) with min(0,u).

We show now that S is a contraction on YT,R for sufficiently small T > 0. We start with a preparation. By testing (Equation23) with u and taking into account the L bound for Ks(δ), we deduce from Young’s inequality for products and convolutions that Rdu(t)2dx+σ20tRd|u|2dxdτRd|ρσ0|2dx+C(δ,η,σ)0tu22v24dτ, since fσ(η)(v)1Cηv22 for vYT. Then, if vYT,R, we infer from Gronwall’s lemma that (24) Rdu(t)2dx+σ0tRd|u|2dxdτeC(σ,δ,η)R4tRd|ρσ0|2dxfor 0tT.(24) Let viYT,R and set ui=S(vi), i = 1, 2. We compute u1Ks(δ)*fσ(η)(v1)u2Ks(δ)*fσ(η)(v2)2(u1u2)Ks(δ)*fσ(η)(v1)2+u2Ks(δ)*(fσ(η)(v1)fσ(η)(v2))2u1u22Ks(δ)*fσ(η)(v1)+u22Ks(δ)*(fσ(η)(v1)fσ(η)(v2))u1u22Ks(δ)fσ(η)(v1)1+u22Ks(δ)fσ(η)(v1)fσ(η)(v2)1C(δ,η)(u1u22v122+u22(v12+v22)v1v22). Therefore, using (Equation24), for v1,v2YT,R, (25) u1Ks(δ)*fσ(η)(v1)u2Ks(δ)*fσ(η)(v2)2C(δ,η,R,T)(u1u22+v1v22).(25) Next, we write (Equation23) for (ui, vi) in place of (u, v), i = 1, 2, take the difference between the two equations, and test the resulting equation with u1u2: 12(u1u2)(t)22+σ0tRd|(u1u2)|2dxdτ=0tRd(u1u2)·(u1Ks(δ)*fσ(η)(v1)u2Ks(δ)*fσ(η)(v2))dxdτσ20tRd|(u1u2)|2dxdτ+12σ0tu1Ks(δ)*fσ(η)(v1)u2Ks(δ)*fσ(η)(v2)22dτ. It follows from (Equation25) that (u1u2)(t)22+σ0tRd|(u1u2)|2dxdτC(δ,η,R,T,σ)0t(u1u222+v1v222)dτ, and we conclude from Gronwall’s lemma that (u1u2)(t)22eC(δ,η,R,T,σ)t0Tv1v222dτfor 0tT. This inequality implies that S is a contraction in YT,R, provided that T is sufficiently small. Therefore, by Banach’s theorem, S admits a unique fixed point uYT,RYT for T > 0 sufficiently small.

It remains to show that the local solution can be extended to a global one. To this end, we note that the function uXT satisfies (Equation23) with v = u: (26) tuσΔu=div (uKs(δ)*fσ(η)(u))in Rd×(0,T),u(·,0)=ρσ0in Rd.(26) Then, defining the truncated entropy density h(η)(ρ)=0ρ0uDfσ(η)(v)v1dvdu,ρ0, and testing (Equation26) with Dh(η)(u) yields, in view of the definition of Ks(δ), that (27) Rdh(η)(u(t))dx+σ0tRdDfσ(η)(u)u1|u|2dxdτ+0tRd|K˜s/2(δ)*fσ(η)(u)|2dxdτ=Rdh(η)(ρσ0)dx(27) for 0tT. This inequality and the definitions of fσ(η) and h(η) yield a (δ,T)-uniform bound for u in L2(0,T;H1(Rd)), which in turn (together with (Equation26)) implies a (δ,T)-uniform bound for u in XT, and a fortiori in YT. This means that the solution u can be prolonged to the whole time interval [0,) and exists for all times.

Finally, we point out that, since Ks(δ)L2(Rd), then, Ks(δ)*fσ(η)(u)L(0,T;L2(Rd)) and so uKs(δ)*fσ(η)(u)L(0,T;L1(Rd)). This fact yields the conservation of mass for u, i.e., Rdu(t)dx=Rdρσ0dx for t > 0. Indeed, it is sufficient to test (Equation26) with a cutoff ψRC01(Rd) satisfying ψR(x)=1 for |x|<R,ψR(x)=0 for |x|>2R,|ψR(x)|CR1 for xRd, and then, to take the limit R.

Step 2: Limit δ0. Let u(δ) be the solution to (Equation26). An adaption of the proof of [Citation24, lemma 1] shows that the embedding H1(Rd)L1(Rd;(1+|x|2)κ/2)L2(Rd) is compact. Thus, because of the δ-uniform bounds for u(δ), the Aubin–Lions Lemma implies that (up to a subsequence) u(δ)u strongly in L2(0,T;L2(Rd)) for every T > 0. We wish now to study the convergence of the nonlinear and nonlocal terms in (Equation26)–(Equation27) as δ0.

It follows from (Equation27) that (up to a subsequence) (28) K˜s/2(δ)*fσ(η)(u(δ))Uweakly in L2(Rd×(0,T))as δ0.(28) To identify the limit U, we first notice that, by construction, 0K˜s/2(δ)Ks/2 a.e. in Rd. Furthermore, the Hardy–Littlewood–Sobolev inequality, the bound for fσ(η), and then, the Gagliardo–Nirenberg–Sobolev inequality yield that Ks/2*fσ(η)(u)(d+2)/(ds)Cfσ(η)(u)(d+2)/(d+2s/d)C(η)u(2d+4)/(d+2s/d)2C(η)u22(s+2)/(d+2)u22(ds)/(d+2). Therefore, since uL(0,T;L2(Rd))L2(0,T;H1(Rd)), 0T||Ks/2*fσ(η)(u)(d+2)/(ds)(d+2)/(ds)dtC(η)uL(0,T;L2(Rd))2(s+2)/(ds)0T||u22dtC(η,T), meaning that Ks/2*fσ(η)(u)L(d+2)/(ds)(Rd×(0,T)). Taking into account that fσ(η)(u)0 and that δK˜s/2(δ)(x)R is nonincreasing (see Lemma 12), we deduce from monotone convergence that (29) K˜s/2(δ)*fσ(η)(u)Ks/2*fσ(η)(u)strongly in L(d+2)/(ds)(Rd×(0,T)).(29) Furthermore, arguing as before and using the estimates for Dfσ(η) leads to K˜s/2(δ)*(fσ(η)(u(δ))fσ(η)(u))(d+2)/(ds)K˜s/2*|fσ(η)(u(δ))fσ(η)(u)|(d+2)/(ds)Cfσ(η)(u(δ))fσ(η)(u)(d+2)/(d+2s/d)C(η)|u|+|u(δ)|(2d+4)/(d+2s/d)uu(δ)(2d+4)/(d+2s/d)C(η)(u2(s+2)/(d+2)u2(ds)/(d+2)+u(δ)2(s+2)/(d+2)u(δ)2(ds)/(d+2))×uu(δ)2(s+2)/(d+2)(uu(δ))2(ds)/(d+2). Since u(δ) is bounded in L(0,T;L2(Rd))L2(0,T;H1(Rd)), it follows that (up to a subsequence) K˜s/2(δ)*(fσ(η)(u(δ))fσ(η)(u)) converges weakly to some limit in L(d+2)/(ds)(Rd×(0,T)). However, Hölder’s inequality and the fact that u(δ)u strongly in Lp(0,T;L2(Rd)) for every 2p<, which follows from 0T||u(δ)u2pdtsup0<t<T(u(δ)u)(t)2p20T||u(δ)u22dt0as δ0, imply that K˜s/2(δ)*(fσ(η)(u(δ))fσ(η)(u))0strongly in Lp(0,T;L(d+2)/(ds)(Rd)),p<d+2ds. We conclude that (30) K˜s/2(δ)*(fσ(η)(u(δ))fσ(η)(u))0weakly in L(d+2)/(ds)(Rd×(0,T)).(30)

We deduce from (Equation29) to (Equation30) that K˜s/2(δ)*fσ(η)(u(δ))Ks/2*fσ(η)(u)=(K˜s/2(δ)*fσ(η)(u)Ks/2*fσ(η)(u))+K˜s/2(δ)*(fσ(η)(u(δ))fσ(η)(u))0weakly in L(d+2)/(ds)(Rd×(0,T)), which, together with (Equation28), implies that U=Ks/2fσ(η)(u), that is, (31) K˜s/2(δ)*fσ(η)(u(δ))Ks/2*fσ(η)(u)weakly in L2(Rd×(0,T)).(31) Let ψC0(Rd×(0,T)). Because of Ks(δ)*fσ(η)(u(δ))=K˜s/2(δ)(K˜s/2(δ)*fσ(η)(u(δ))), we find that 0TRdψ·Ks(δ)*fσ(η)(u(δ))dxdt=0TRd(K˜s/2(δ)*fσ(η)(u(δ)))·(K˜s/2(δ)*ψ)dxdt. Our goal is to show that K˜s/2(δ)*ψKs/2*ψ strongly in L2(Rd×(0,T)) as δ0. We can assume without loss of generality that ψ0 a.e. in Rd×(0,T). Indeed, for general functions ψ, we may write ψ=ψ++ψ, where ψ+=max{0,ψ} and ψ=min{0,ψ}, and we have K˜s/2(δ)*ψ=K˜s/2(δ)*ψ+K˜s/2(δ)(ψ). Once again, since K˜s/2(δ)K˜s/2 a.e. in Rd, it is sufficient to show that Ks/2ψL2(Rd×(0,T)). The Hardy–Littlewood–Sobolev inequality (see Appendix B) yields 0TKs/2ψ22dtC0Tψ2d/(d+2s)2dt. It follows from (Equation31), the previous argument, and the fact that Ks*u=(Δ)su=Ks/2*Ks/2*u that 0TRdψ·Ks(δ)*fσ(η)(u(δ))dxdt0TRd(Ks/2*fσ(η)(u))·(Ks/2ψ)dxdt=0TRdψ·Ks*fσ(η)(u)dxdt for every ψL2(0,T;L2d/(d+2s)(Rd)), which means that (32) Ks(δ)*fσ(η)(u(δ))Ks*fσ(η)(u)weakly in L2(0,T;L2d/(d2s)(Rd)).(32) Since u(δ)u strongly in L2(0,T;L2(Rd)) and (u(δ)) is bounded in L(0,T;L1(Rd)) (via mass conservation), it also holds that u(δ)u strongly in L2(0,T;L2d/(d+2s)(Rd)). Therefore, the convergence (Equation32) is sufficient to pass to the limit δ0 in (Equation26).

Step 3: Limit η0 and conclusion. The limit δ0 in (Equation26) shows that the limit u solves (33) tuσΔu=div (uKs*fσ(η)(u))in Rd×(0,T),u(·,0)=ρσ0in Rd.(33) Fatou’s Lemma and the weakly lower semicontinuity of the L2 norm allow us to infer from (Equation27) that for t > 0, (34) Rdh(η)(u(t))dx+σ0tRdDfσ(η)(u)u1|u|2dxdτ+0tRd|Ks/2*fσ(η)(u)|2dxdτRdh(η)(ρσ0)dx.(34) At this point, all the bounds for u, derived in the previous subsections, and the moment estimate, contained in Lemma 9, can be proved like in Subsections 2.12.2. All these estimates are uniform in η. It is rather straightforward to perform the limit η0 in (Equation33)–(Equation34) to obtain a weak solution to (Equation7). However, the higher regularity bounds obtained in Subsection 2.3 imply that u is actually a strong solution to (Equation7), which in turn yields the uniqueness of u as a weak solution to (Equation7). This finishes the proof of Theorem 4.

2.5. Limit σ0

We prove that there exists a subsequence of (ρσ) that converges strongly in L1(Rd×(0,T)) to a weak solution ρ to (Equation1).

The uniform L(Rd×(0,T)) bound for ρσ in Lemma 7 implies that, up to a subsequence, ρσ*ρ weakly* in L(Rd×(0,T)) as σ0. We deduce from the uniform L(0,T;L1(Rd)) bound (Equation13) and the moment bound for ρσ in Lemma 9 that (ρσ) is equi-integrable. Thus, by the Dunford–Pettis theorem, again up to a subsequence, ρσρ weakly in L1(Rd×(0,T)). It follows from the L2(0,T;H1(Rd)) estimate (Equation16) that σΔρσ0 strongly in L2(0,T;H1(Rd)). The estimates in (Equation17) and Lemma 7 show that (tρσ) is bounded in L2(0,T;H1(Rd)) and consequently, up to a subsequence, tρσtρ weakly in L2(0,T;H1(Rd)). Therefore, the limit σ0 in (Equation7) leads to (35) tρ=div (ρσ(Δ)sfσ(ρσ)¯)in L2(0,T;H1(Rd)),(35) where the overline denotes the weak limit of the corresponding sequence.

We need to identify the weak limit on the right-hand side. The idea is to use the div-curl lemma [Citation41, theorem 10.21]. For this, we define the vector fields with d + 1 components Uσ:=(ρσ,ρσ(Δ)sfσ(ρσ)),Vσ:=(fσ(ρσ),0,,0). Let R > 0 be arbitrary and write BR for the ball around the origin with radius R. The L(Rd) bound (Equation15) for ρσ and the L2(0,T;H1s(Rd)) bound (Equation17) for fσ(ρσ) show that (Uσ) is bounded in Lp(BR×(0,T)) for some p > 1, while (Vσ) is bounded in L(BR×(0,T)). Furthermore, by (Equation17), div(t,x)Uσ=σΔρσ0strongly in L2(0,T;H1(BR))H1(BR×(0,T)),curl(t,x)VσL2(0,T;Hs(BR))Cfσ(ρσ)L2(0,T;Hs(BR))C, where curl(t,x)Vσ is the antisymmetric part of the Jacobian matrix of Vσ. Hence, by the compact embedding Hs(BR×(0,T))W1,r(BR×(0,T)) (since L2(0,T;Hs(BR))Hs(BR×(0,T))), the sequence (curl(t,x)Vσ) is relatively compact in W1,r(BR×(0,T)) for some r > 1. Therefore, we can apply the div-curl lemma giving Uσ·Vσ¯=Uσ¯·Vσ¯ or ρσfσ(ρσ)¯=ρfσ(ρσ)¯a.e. in BR×(0,T).

By definition (Equation9) of fσ(ρσ), it follows for arbitrary ρσ[0,L] and sufficiently large L > 0, that fσ(ρσ)=0ρσ(Γσ*(f1[0,)))(u)Ξ˜(σu)du=0ρσ0Γσ(uw)f(w)dwΞ˜(σu)du=0ρσ0Γσ(uw)f(w)dwΞ˜(σu)du=0(0ρσΓσ(uw)Ξ˜(σu)du)f(w)dw. We use the properties that (ρσ) is uniformly bounded and Ξ˜=1 in [1,1]. Then, choosing σ>0 sufficiently small, fσ(ρσ)=0(0ρσΓσ(uw)du)f(w)dw=0Γσ(ρσw)f(w)dw0Γσ(w)f(w)dw=RΓσ(ρσw)f˜(w)dwRΓσ(w)f˜(w)dw, setting f˜:=f1[0,). Hence, using f(0)=0, we find that fσ(ρσ)f(ρσ)=RΓσ(u)(f˜(u+ρσ)f˜(ρσ))duRΓσ(w)(f˜(w)f˜(0))dw. Taking into account the fundamental theorem of calculus for the function f˜C0W1,1(R), we can estimate as follows: |fσ(ρσ)f(ρσ)|  ess supusupp(Γσ)\{0}(|f˜(u+ρσ)f˜(ρσ)||u|+|f˜(u)f˜(0)||u|)RΓσ(w)|w|dw(maxξsupp(Γσ)[0,)(f(ξ+ρσ)+f(ξ)))RΓσ(w)|w|dw. Then, since Γσ(u)=σ1Γ1(σ1u), supp(Γσ)Bσ(0) is compact, fC1([0,)), and (ρσ) is uniformly bounded, we conclude that |fσ(ρσ)f(ρσ)|Cσ. This means that fσ(ρσ)f(ρσ)0 strongly in L(BR×(0,T)), and it shows that ρσf(ρσ)¯=ρf(ρσ)¯ a.e. in BR×(0,T). As f is nondecreasing, we can apply [Citation41, theorem 10.19] to infer that f(ρσ)¯=f(ρ) a.e. in BR×(0,T). Consequently, ρσf(ρσ)¯=ρf(ρ). As uuf(u) is assumed to be strictly convex, we conclude from [Citation41, theorem 10.20] that (ρσ) converges a.e. in BR×(0,T). Since (ρσ) is bounded in L(Rd×(0,T)), it follows that ρσρ strongly in Lp(BR×(0,T)) for all p<. Using the moment estimate from Lemma 9, we infer from lim supσ00TRd|ρσρ|dxdt=lim supσ00TRdBR|ρσρ|dxdtR2d/(d2s)lim supσ00TRdBRρσ(t,x)|x|2d/(d2s)dxR2d/(d2s)C0as R that ρσρ strongly in Lp(Rd×(0,T)) for all p<. The strong convergences of ρσ and fσ(ρσ) in Lp(Rd×(0,T)) for all p< allow us to identify the weak limit in (Equation35), proving the weak formulation (Equation8).

Finally, we deduce from the uniform L2(0,T;H1(Rd)) bound for tρσ and the fact that ρσρ strongly in Lp(Rd) for any p< that ρ(0)=ρ0 in the sense of H1(Rd). Properties (iv) of Theorem 1 follow from the corresponding expressions satisfied by ρσ in the limit σ0.

2.6. Time-uniform convergence of (ρσ)

The following lemma is needed in the proof of Proposition 3. It is essentially a consequence of the L2(0,T;H1(Rd)) bound of tρσ and the Ascoli–Arzelà theorem.

Corollary 13.

Under the assumptions of Theorem 1, it holds for all ϕL(Rd) that, possibly for a subsequence,RdρσϕdxRdρϕdx uniformly in [0,T].

Proof.

Let ϕC01(Rd) and 0t1<t2T. The uniform L2(0,T;H1(Rd)) bound of tρσ implies that |Rdρσ(t2)ϕdxRdρσ(t1)ϕdx|=|t1t2tρσ,ϕdt||t2t1|1/2tρσL2(0,T;H1(Rd))ϕH1(Rd)C|t2t1|1/2ϕH1(Rd). Hence, the sequence of functions tRdρσ(t)ϕds is bounded and equicontinuous in [0,T]. By the Ascoli–Arzelá theorem, up to a ϕ-depending subsequence, Rdρσϕdxξϕ strongly in C0([0,T]) as σ0. Since ρσ*ρ weakly* in L(0,T;L(Rd)), we can identify the limit, ξϕ=Rdρϕdx. Since H1(Rd) is separable, a Cantor diagonal argument together with a density argument allows us to find a subsequence (which is not relabeled) such that for all ϕH1(Rd), (36) RdρσϕdxRdρϕdx strongly in C0([0,T]).(36) Since (ρσ) is bounded in L(0,T;L2(Rd)), another density argument shows that this limit also holds for all ϕL2(Rd).

Now, let ϕL(Rd). Using ϕ1{|x|<R}L2(Rd), it follows from (Equation36) and the moment estimate for ρσ that lim supσ0sup0<t<T|Rdρσ(t)ϕdxRdρ(t)ϕdx|lim supσ0sup0<t<T|Rdρσ(t)ϕ1{|x|>R}dxRdρ(t)ϕ1{|x|>R}dx|R2d/(d2s)ϕlim supσ0sup0<t<TRd(ρσ(x,t)+ρ(x,t))|x|2d/(d2s)dxC(T)R2d/(d2s)ϕ0as R. This shows that limσ0sup0<t<T|Rdρσ(t)ϕdxRdρ(t)ϕdx|=0, concluding the proof. □

3. Analysis of EquationEq. (Equation5)

This section is devoted to the analysis of EquationEq. (Equation5), (37) tρσ,β,ζσΔρσ,β,ζ=div (ρσ,β,ζKζ*fσ(Wβ*ρσ,β,ζ)),t>0,ρσ,β,ζ(0)=ρσ0in Rd,(37) where Kζ=K˜ζ*Wζ and Wβ is defined in (Equation10), as well as to an estimate for the difference ρσ,β,ζρσ, which is needed in the mean-field analysis. The existence and uniqueness of a strong solution to (Equation37) follows from standard parabolic theory, since we regularized the singular kernel and smoothed the nonlinearity.

Proposition 14

(Uniform estimates). Let Hypotheses (H1)–(H3) hold and let T > 0, p > d. Set a:=min{1,d2s}, let ρσ be the strong solution to (Equation7), and let ρσ,β,ζ be the strong solution to (Equation5). Then, there exist constants C1>0, and ε0>0, both depending on σ, p, and T, such that if β+ζa<ε0 then,(38) ρσ,β,ζρσL(0,T;W2,p(Rd))C1(β+ζa),(38) (39) ρσ,β,ζL(0,T;W2,p(Rd))C1.(39)

Furthermore, for every q2, there exists C2=C2(q)>0, depending on σ and T, such that (40) (KζK)*ρσL(0,T;L(Rd))C2ζa,(40) (41) ρσ,β,ζL(0,T;W2,1(Rd)W3,q(Rd))C2.(41) The proof is presented in the following subsections. The most difficult part is the proof of (Equation38) in Subsection 3.1. We first prove an estimate for D2(ρσ,β,ζρσ) that depends on a lower-order estimate of this difference. Second, this lower-order estimate is shown by testing the equation satisfied by the difference ρσ,β,ζρσ with a suitable nonlinear test function. Based on the arguments of this section, estimates (Equation39)–(Equation41) are, then, shown in Subsections 3.23.4, respectively.

3.1. Proof of (Equation38)

We introduce the difference u:=ρσ,β,ζρσ, which satisfies (42) tuσΔu=div [(u+ρσ)Kζ*fσ(Wβ*(u+ρσ))ρσK*fσ(ρσ)]=D[u]+R[ρσ,u]+S[ρσ,u]in Rd,t>0,(42) and the initial datum u(0)=0 in Rd, where D[u]=div [uK*fσ(Wβ*u)],R[ρσ,u]=div [uK*(fσ(Wβ*(u+ρσ))fσ(Wβ*u))+ρσK*(fσ(Wβ*(u+ρσ))fσ(Wβ*ρσ))+ρσK*(fσ(Wβ*ρσ)fσ(ρσ))],S[ρσ,u]=div [(u+ρσ)(KζK)*fσ(Wβ*(u+ρσ))].

We show first an estimate for D2u that depends on a lower-order estimate for u.

Lemma 15

(Conditional estimate for u). For any p > d, there exists a number Γp(0,1) such that, if sup0<t<Tu(t)W1,p(Rd)Γp, then,D2uLp(0,T;Lp(Rd))C(uLp(0,T;W1,p(Rd))+β+ζa),recalling that a=min{1,d2s}, and where C > 0 is independent of u, β, and ζ, but may depend on σ.

Proof.

Let Γp(0,1) be such that sup0<t<Tu(t)W1,p(Rd)Γp. We will find a constraint for Γp at the end of the proof. The aim is to derive an estimate for the right-hand side of (Equation42) in Lp(0,T;Lp(Rd)). We observe that u(t)1ρσ,β,ζ1+ρσ12ρ01 for t[0,T]. In the following, we denote by C > 0 a generic constant that may depend on σ, without making this explicit. Furthermore, we denote by μ a generic exponent in (0, 1), whose value may vary from line to line.

Step 1: Estimate of D[u]. Let 1/2<s<1. Then, by the Hardy–Littlewood–Sobolev-type inequality (B3), D[u]pu·K*fσ(Wβ*u)p+uK*[fσ(Wβ*u)Wβ*u]pCupfσ(Wβ*u)d/(2s1)+Cud/(2s1)fσ(Wβ*u)up. We use the Young convolution inequality, the Gagliardo–Nirenberg inequality, the smoothness of fσ, the property fσ(0)=0, and the fact WβL1(Rd)=1 to estimate the terms on the right-hand side: Wβ*uuu11λupλCΓpλC,fσ(Wβ*u)maxU|fσ| Wβ*uC,fσ(Wβ*u)|fσ(0)|+maxU|fσ|Wβ*uC,ud/(2s1)u11μuμCuW1,p(Rd)μCΓpμC, where U:=[Wβ*u,Wβ*u] and λ>0,μ>0. Therefore, D[u]pCup and (43) D[u]Lp(0,T;Lp(Rd))CuLp(0,T;W1,p(Rd)).(43) Next, let 0<s1/2. Then, we write D[u]=u·K*[fσ(Wβ*u)Wβ*u]+uK*[fσ(Wβ*u)|Wβ*u|2]+uK*[fσ(Wβ*u)Wβ*Δu]=:D1+D2+D3. By the Hardy–Littlewood–Sobolev-type inequality (Lemma 22), D1pCud/(2s)fσ(Wβ*u)Wβ*upCud/(2s)up. Next, we apply the Gagliardo–Nirenberg inequality with λ=(1+1/d2s/d)/(1+2/d1/p): ud/(2s)Cu11λD2upλCD2upλ, which is possible as long as λ1/2 or equivalently d2s, which is true. Consequently, using Γp1, D1pCupD2upλCΓpλup1λD2upλC(δ)up+δD2up, where δ>0 is arbitrary. It follows from the Hardy–Littlewood–Sobolev-type inequality and the Gagliardo–Nirenberg inequality u2p2CD2upd/pup2d/pCΓpD2upd/pup1d/p that D2pCud/2sΓpD2upd/pup1d/pC(δ)up+δD2up. Finally, using similar ideas, we obtain D3pCud/(2s)ΔupCΓpμD2up. Summarizing the estimates for D1, D2, and D3 and integrating in time leads to (44) D[u]Lp(0,T;Lp(Rd))CuLp(0,T;W1,p(Rd))+CΓpμD2uLp(0,T;Lp(Rd)).(44)

Step 2: Estimate of R[ρσ,u]. We write R[ρσ,u]=R1+R2+R3 for the three terms in the definition of R[ρσ,u] below (Equation42).

Step 2a: Estimate of R1. If s > 1∕2, we can argue similarly as in the derivation of (Equation43), which gives R1Lp(0,T;Lp(Rd))CuLp(0,T;W1,p(Rd)). If 0<s1/2, we write R1=R11++R16, where R11=u·K*[fσ(Wβ*(u+ρσ))Wβ*ρσ],R12=uK*[fσ(Wβ*(u+ρσ))(Wβ*ρσ)·(Wβ*(u+ρσ))],R13=uK*[fσ(Wβ*(u+ρσ))Wβ*Δρσ],R14=u·K*[(fσ(Wβ*(u+ρσ))fσ(Wβ*u))Wβ*u],R15=uK*[(fσ(Wβ*(u+ρσ))Wβ*(u+ρσ)fσ(Wβ*u)(Wβ*u))·(Wβ*u)],R16=uK*[(fσ(Wβ*(u+ρσ))fσ(Wβ*u))Wβ*Δu]. All terms except the last one can be treated by the Hardy–Littlewood–Sobolev and Gagliardo–Nirenberg inequalities as before. For the last term, we use these inequalities and the L(Rd) bound for ρσ: R16pCud/(2s)(fσ(Wβ*(u+ρσ))fσ(Wβ*u))Wβ*ΔupCud/(2s)fσWβ*ρσWβ*ΔupCud/(2s)ΔupCΓpμD2up. We infer that (possibly with a different μ>0 than before) R1Lp(0,T;Lp(Rd))CuLp(0,T;W1,p(Rd))+CΓpμD2uLp(0,T;Lp(Rd)).

Step 2b: Estimate of R2. Since |fσ| is bounded on the interval [uρσ,u+ρσ], we obtain for s > 1∕2, R2Lp(0,T;Lp(Rd))CuLp(0,T;W1,p(Rd)). For 0<s1/2, we write R2=R21++R27, where R21=ρσ·K*[fσ(Wβ*(u+ρσ))Wβ*u],R22=ρσK*[fσ(Wβ*(u+ρσ))Wβ*(u+ρσ)·(Wβ*u)],R23=ρσK*[fσ(Wβ*(u+ρσ))Wβ*Δu],R24=ρσ·K*[(fσ(Wβ*(u+ρσ))fσ(Wβ*ρσ))Wβ*ρσ],R25=ρσK*[fσ(Wβ*(u+ρσ))(Wβ*u)·(Wβ*ρσ)],R26=ρσK*[(fσ(Wβ*(u+ρσ))fσ(Wβ*ρσ))|Wβ*ρσ|2],R27=ρσK*[(fσ(Wβ*(u+ρσ))fσ(Wβ*ρσ))Wβ*Δρσ]. Similar estimations as before allow us to treat all terms except the third one: R23pρσK*[(fσ(Wβ*(u+ρσ))fσ(Wβ*ρσ))Wβ*Δu]p+ρσK*[fσ(Wβ*ρσ)Wβ*Δu]p=:Q231+Q232. The first term can be estimated similarly as above by Q231CΓpμD2up, while Q232ρσΔK*[fσ(Wβ*ρσ)Wβ*u]p+ρσK*[Δfσ(Wβ*ρσ)Wβ*u]p+2ρσK*[fσ(Wβ*ρσ)·(Wβ*u)]p. It follows from ΔK*v=(Δ)1sv and the fractional Gagliardo–Nirenberg inequality (Lemma 24) that Q232CuW1,p(Rd)+ρσ(Δ)1s[fσ(Wβ*ρσ)Wβ*u]pCuW1,p(Rd)+Cρσfσ(Wβ*ρσ)Wβ*upsD2[fσ(Wβ*ρσ)Wβ*u]p1sCuW1,p(Rd)+Cups(up1s+up1s+D2up1s)CuW1,p(Rd)+CΓpD2up. This shows that R23pCuW1,p(Rd)+CΓpμD2up, and we conclude that R2Lp(0,T;Lp(Rd))CuLp(0,T;W1,p(Rd))+CΓpμD2uLp(0,T;Lp(Rd)).

Step 2c: Estimate of R3. We write R3=R31++R37, where R31=ρσ·K*[(fσ(Wβ*ρσ)fσ(ρσ))Wβ*ρσ],R32=ρσK*[(fσ(Wβ*ρσ)fσ(ρσ))|Wβ*ρσ|2],R33=ρσK*[fσ(ρσ)(Wβ*ρσρσ)·(Wβ*ρσ)],R34=ρσ·K*[fσ(ρσ)(Wβ*ρσρσ)],R35=ρσK*[fσ(ρσ)ρσ·(Wβ*ρσρσ)],R36=ρσK*[fσ(ρσ)(Wβ*ΔρσΔρσ)]R37=ρσK*[(fσ(Wβρσ)fσ(ρσ))Wβ*Δρσ] We start with the estimate of R31. We use the Hardy–Littlewood–Sobolev inequality (Lemma 22) and Lemma 21 to estimate Wβ*ρσρσ: R31Cρσd/sfσ(Wβ*ρσ)fσ(ρσ)pWβ*ρσd/sCρσd/s2max[0,2ρσ]|fσ| Wβ*ρσρσpC(σ)β, also taking into account the L(0,T;Lq(Rd)) bound for ρσ; see Proposition 10. With this regularity, we can estimate all other terms except R34 and R36. Since they have similar structures, we only treat R34. This term is delicate since the factor fσ(ρσ) cannot be bounded in Lq(Rd) for any q<. Therefore, one might obtain via Hardy–Littlewood–Sobolev’s inequality factors like ρσq1 and D2ρσq2 with either q1<2 or q2<2. However, for such factors, an L bound in time is currently lacking (Proposition 10 provides such a bound only for q2). Our idea is to add and subtract the term fσ(0) since |fσ(ρσ)fσ(0)|ρσmax[0,ρσ]|fσ|Cρσ can be controlled. This leads to R34pρσ·K*[(fσ(ρσ)fσ(0))(Wβ*ρβρβ)p+fσ(0)ρσ·K*(Wβ*ρσρσ)pCβ+|fρ(0)|ρσ·K*(Wβ*ρσρσ)p=:Cβ+Q341, as the first term can be estimated in a standard way. For the estimate of Q341, we need to distinguish two cases.

If 1/2<s1, we infer from the Hardy–Littlewoord–Sobolev-type inequality (B3) that Q341Cρσd/(2s1)Wβ*ρσρσpCρσd/(2s1)ρσpβCβ. Next, let 0<s1/2. Then, we apply the Hardy–Littlewoord–Sobolev-type inequality (B2), the standard Gagliardo–Nirenberg inequality for some λ>0, and Lemma 21: Q341Cρσd/(2s)Wβ*ρσρσpCρσ11λD2ρσpλ(βD2ρσp)Cβ. We conclude that R34pCβ and eventually R3Lp(0,T;Lp(Rd))Cβ. Summarizing the estimates for R1, R2, and R3 finishes this step: (45) R[ρσ,u]Lp(0,T;Lp(Rd))CuLp(0,T;W1,p(Rd))+Cβ+CΓpμD2uLp(0,T;W1,p(Rd)).(45)

Step 3: Estimate of S[ρσ,u]. We formulate this term as S[ρσ,u]=S1++S4, where S1=div [u(KζK)*(fσ(Wβ*(u+ρσ))fσ(Wβ*ρσ))],S2=div (u(KζK)*fσ(Wβ*ρσ)),S3=div [ρσ(KζK)*(fσ(Wβ*(u+ρσ))fσ(Wβ*ρσ))],S4=div (ρσ(KζK)*fσ(Wβ*ρσ)). The terms S1, S2, and S3 can be treated as the terms in R[ρσ,u], since they have the same structure and the techniques used to estimate integrals involving K can be applied to those involving Kζ. This leads to (for some μ>0) (46) S1+S2+S3Lp(0,T;Lp(Rd))CuLp(0,T;W1,p(Rd))+CΓpμD2uLp(0,T;Lp(Rd)).(46) It remains to estimate S4. We write S4=S41+S42+S43, where S41=ρσ·(KζK)*[fσ(Wβ*ρσ)Wβ*ρσ],S42=ρσ(KζK)*[fσ(Wβ*ρσ)|Wβ*ρσ|2],S43=ρσ(KζK)*[fσ(Wβ*ρσ)Wβ*Δρσ]. Observe that, because of the definition of Kζ=K˜ζWζ with K˜ζ=Kωζ (defined in (Equation11)), we have (KζK)*v=K*(Wζ*vv)(K(1ωζ))Wζv for every function v for which the convolution is defined, and therefore, by the Hardy–Littlewood–Sobolev-type inequality (B2), Young’s convolution inequality, and Lemma 21, ρσ(KζK)*vpCρσd/(2s)Wζ*vvp+Cρσp(K(1ωζ))vCρσd/(2s)vpζ+CρσpK1Rd\B(0,ζ1)v1Cρσd/(2s)vpζ+Cζd2sρσpv1, Given the regularity properties of ρσ (see Lemma 11) and the assumptions on fσ, it follows that (47) S4Lp(0,T;Lp(Rd))Cζmin{1,d2s}.(47)

We conclude from (Equation46) and (Equation47) that (48) S[ρσ,u]Lp(0,T;Lp(Rd))CuLp(0,T;W1,p(Rd))+Cζa+CΓpμD2uLp(0,T;Lp(Rd)),(48) where a:=min{1,d2s}.

Step 4: End of the proof. Summarizing (44, 45), and (Equation48), we infer that the right-hand side of (Equation42) can be bounded (for some μ>0) by D[u]+R[ρσ,u]+S[ρσ,u]Lp(0,T;Lp(Rd))CuLp(0,T;W1,p(Rd))+C(β+ζa)+CΓpμD2uLp(0,T;Lp(Rd)). By parabolic regularity (C1), D2uLp(0,T;Lp(Rd))CuLp(0,T;W1,p(Rd))+C(β+ζa)+CΓpμD2uLp(0,T;Lp(Rd)). Choosing Γp>0 sufficiently small finishes the proof. □

It remains to estimate the Lp(0,T;W1,p(Rd)) norm of u. This is done in the following lemma.

Lemma 16

(Unconditional estimate for u). For any p > d, there exist constants C > 0, and ε0>0, both depending on σ, p, and T, such that for β+ζa<ε0, uL(0,T;W1,p(Rd))C(β+ζa).recalling that a:=min{1,d2s}.

Proof.

The idea is to test (Equation42) with p|u|p2updiv (|u|p2u). Integration by parts and some elementary computations lead to Rdpdiv (|u|p2u)Δudx=pi,jRd|u|p2iuijj2udx=pi,jRdj(|u|p2iu)ij2udx=pRd|u|p2|D2u|2dx+p2jRdj(|u|p2)j(|u|2)dx=pRd|u|p2|D2u|2dx+jRd4p(p2)(j(|u|p/2))2dx. Consequently, we have (49) pu(t)W1,p(Rd)p+σp(p1)0tRd|u|p2|u|2dxds+σ0tRd(p|u|p2|D2u|2+4(p2)p1|(|u|p/2)|2)dxds=p0tRd(|u|p2udiv (|u|p2u))(D[u]+R[ρσ,u]+S[ρσ,u])dxds=:Q[u].(49)

We infer from Lemmas 20 and 26 that uC0([0,T];W1,p(Rd)). Therefore, since u(0)=0, it holds that u(t)W1,p(Rd)Γp for all t[0,T*] and T*:=sup{t0(0,T):u(t)W1,p(Rd)Γp for 0tt0}. Let t[0,T*]. We have shown in the proof of the previous lemma that D[u]+R[ρσ,u]+S[ρσ,u]Lp(0,t;Lp(Rd))CuLp(0,t;W1,p(Rd))+C(β+ζa). Hence, we can estimate the right-hand side Q[u] of (Equation49) as follows: Q[u]C0tRd(|u|p1+|u|p2|D2u|)|D[u]+R[ρσ,u]+S[ρσ,u]|dxC(uLp(0,t;Lp(Rd))p1+uLp(0,t;Lp(Rd))p/21|u|p/21|D2u|L2(0,t;L2(Rd)))×(uLp(0,t;W1,p(Rd))+β+ζa)C(δ,p,t)(uLp(0,t;W1,p(Rd))p+(β+ζa)p)+δ|u|p/21|D2u|L2(0,t;L2(Rd))2, where δ>0. Choosing δ sufficiently small, the last term is absorbed by the corresponding expression on the left-hand side of (Equation49), and we infer from (Equation49) that for 0tT*, u(t)W1,p(Rd)pC(p,t)0tuW1,p(Rd)pds+C(p,t)(β+ζa)p. We assume without loss of generality that C(p, t) is nondecreasing in t. Then, Gronwall’s lemma implies that for 0tT*, u(t)W1,p(Rd)pC(p,T)(β+ζa)p0teC(p,T)(ts)ds(β+ζa)eC(p,T)t. Choosing ε0=12Γpexp(C(p,T)T/p)<1, we find that u(t)W1,p(Rd)Γp/2 for β+ζa<ε0 and 0tT*. By definition of T*, it follows that T*=T. In particular, u(t)W1,p(Rd)C(β+ζa) for 0<t<T, which finishes the proof. □

3.2. Proof of (Equation38) and (Equation39)

Combining Lemmas 15 and 16 leads to (50) uLp(0,T;W2,p(Rd))C(σ,p,T)(β+ζa),where a=min{1,d2s},(50) as long as β+ζa<ε0 and p > d. Next, we differentiate (Equation42) with respect to xi (writing i for /xi): t(iu)σΔ(iu)=i(D[u]+R[ρσ,u]+S[ρσ,u]),iu(0)=0in Rd. Taking into account estimate (Equation50) and arguing as in the proof of Lemma 15, we can show that for δ>0, i(D[u]+R[ρσ,u]+S[ρσ,u])Lp(0,T;Lp(Rd))C(p,σ,δ)(β+ζa)+δD3uLp(0,T;Lp(Rd)). We infer from parabolic regularity (Lemma 26) for sufficiently small δ>0 that tDuLp(0,T;Lp(Rd))+D3uLp(0,T;Lp(Rd))C(p,σ)(β+ζa). Then, Lemma 20, applied to Du, leads to (Equation38), which with Proposition 10 implies (Equation39).

3.3. Proof of (Equation40)

Let xRd. We use the definitions of Kζ and Wζ to find that |(KζK)*ρσ(x)|=|RdWζ(xy)((K*ρσ)(x)((Kωζ)*ρσ)(y))dy|RdWζ(xy)|xy||(K*ρσ)(x)(K*ρσ)(y)||xy|dy+(K(1ωζ))*ρσK*ρσRdWζ(z)|z|dz+K1Rd\B(0,ζ1)ρσ1ζK*ρσRdW1(y)|y|dy+ζd2sρσ1. Let ϕC0(Rd) be such that supp(ϕ)B2 and ϕ=1 in B1. Then, (since we can assume without loss of generality that ζ<1), by arguing like in the derivation of (Equation47), we obtain |(KζK)*ρσ(x)|Cζmin{1,d2s}((Kϕ)*ρσ+(K(1ϕ))*ρσ+ρσ1), A computation shows that for p>max{d/(2s),2}, (Kϕ)*ρσ=(Kϕ)*ρσKϕp/(p1)ρσpCρσp,(K(1ϕ))*ρσ(K(1ϕ))ρσ1Cρσ1, where we note that K1B2Lp/(p1) if p>d/(2s). Then, in view of the regularity of ρσ in Lemma 11, we find that (KζK)*ρσL(0,T;L(Rd))Cζa.

3.4. Proof of (Equation41)

The L(0,T;W2,1(Rd)W3,q(Rd)) bound for ρσ,β,ζ is shown in a similar way as the corresponding bound for ρσ in Lemma 11.

4. Mean-field analysis

This section is devoted to the proof of Proposition 3 and Theorem 2. The existence of solutions to (Equation4) and (Equation6) as well as the existence of density functions is shown in Subsection 4.1. In Subsection 4.2, we estimate the difference XiNX¯iN of the processes of the original system (Equation3) and the intermediate system (Equation4), while the difference X¯iNX̂iN of the processes of the intermediate system (Equation4) and the macroscopic system (Equation6) is estimated in Subsection 4.3. These estimates are combined in Subsection 4.4 to conclude with the proof of Proposition 3 and Theorem 2.

4.1. Existence of density functions for (Equation4) and (Equation6)

First, we show that the coefficients of the stochastic differential EquationEq. (Equation6), satisfied by X̂N, are globally Lipschitz continuous and of at most linear growth. The latter condition follows from |K*fσ(ρσ(x,t))|K*fσ(ρσ)L(0,T;L(Rd))CK*fσ(ρσ)L(0,T;W1,p(Rd))Cfσ(ρσ)L(0,T;W1,r(Rd))C(σ), where p > d and r=dp/(d+2s) according to the Hardy–Littlewood–Sobolev inequality, and we used the regularity bounds for ρσ from Lemma 26. The global Lipschitz continuity is a consequence of the mean-value theorem, the Hardy–Littlewood–Sobolev inequality, and the W2,(Rd) regularity of ρσ from Lemma 11: sup0<t<T|K*fσ(ρσ(x,t))K*fσ(ρσ(y,t))|sup0<t<TD2K*fσ(ρσ(·,t))|xy|=sup0<t<TK*(fσ(ρσ)ρσρσ+fσ(ρσ)D2ρσ)(·,t)|xy|C(σ)|xy|. These two conditions yield the existence and uniqueness of solutions to the associated particle systems [Citation44, theorems 2.5 and 2.9]. Moreover, by [Citation45, theorem 2.3.1], the law of the process X̂iN is absolutely continuous with respect to the Lebesgue measure. By Radon-Nikodym’s theorem, there exists a density function û(t) for all t > 0 on Rd, which is measurable and integrable with respect to the Lebesgue measure. (Since all X̂iN are copies of the same process, their density functions are the same almost everywhere.) The processes X̂iN(t) have continuous paths, which implies the continuity of the distribution function of X̂iN(t) with respect to time, and this implies in turn the Bochner measurability of û(t). Clearly, we have sup0<t<Tû(t)L1(Rd)=1, which shows that ûL(0,T;L1(Rd)).

Similar arguments show that X¯iN(t) has a density function u¯L(0,T;L1(Rd)).

Next, we show that û and u¯ can be identified with the weak solutions ρσ and ρσ,β,ζ, respectively, using Itô’s lemma. Indeed, let ϕC0(Rd×[0,T]). We infer from Itô’s formula that ϕ(X̂iN(t),t)=ϕ(X̂iN(0),0)+0tsϕ(X̂iN(s),s)ds+σ0tΔϕ(X̂iN(s),s)ds0tK*fσ(ρσ(X̂iN(s),s))·ϕ(X̂iN(s),s)ds+2σ0tϕ(X̂iN(s),s)·dBiN(s). Taking the expectation, the Itô integral vanishes, and we end up with (51) Rdϕ(x,t)û(x,t)dx=Rdϕ(x,0)ρσ0(x)dx+0tRdsϕ(x,s)û(x,s)dxds+σ0tRdΔϕ(x,s)û(x,s)dxds0tRdK*fσ(ρσ(x,s))·ϕ(x,s)û(x,s)dxds.(51) Hence, û is a very weak solution in the space L(0,T;L1(Rd)) to the linear equation (52) tû=σΔû+div (ûK*fσ(ρσ)),û(0)=ρσ0in Rd,(52) where ρσ is the unique solution to (Equation7).

It can be shown that (Equation52) is uniquely solvable in the class of functions in L(0,T;L1(Rd)). This implies that û=ρσ in Rd×(0,T) (and similarly u¯=ρσ,β,ζ). The proof is technical but standard; see, e.g., [Citation35, theorem 7] for a sketch of a proof.

Another approach is as follows. Because of the linearity of (Equation51), it is sufficient to prove that û0 in Rd×(0,T) if ρσ0=0. First, we verify that v:=K*fσ(ρσ)L(0,T;W1,(Rd)) and ûLp(0,T;Lp(Rd)) for p<d/(d1). Then, by density, (Equation51) holds for all ϕW1,q(0,T;Lq(Rd))Lq(0,T;W2,q(Rd)) with q > d and ϕ(T)=0. Choosing ψ to be the unique strong solution to the dual problem tψ+σΔψ=v·ψ+g,ψ(T)=0in Rd in the very weak formulation of (Equation51), we find that 0TRdûgdxdt=0 for all gC0(Rd×(0,T)), which implies that û=0.

4.2. Estimate of XiNX¯iN

We derive an estimate for the expectation of the difference XiNX¯iN. To this end, we need to estimate the difference of the microscopic average N1j=1,jiNWβ(XjNXiN) and the macroscopic average Wβ*ρβ,ζ,σ(X¯iN). By a careful choice of β and ζ, we show that this estimate is of the order N1/4+δ for δ>0.

Lemma 17.

Let XiN and X¯iN be the solutions to (Equation3) and (Equation4), respectively, and let δ(0,1/4). Under the assumptions of Theorem 3 on β and ζ, it holds thatE(sup0<s<Tmaxi=1,,N|(XiNX¯iN)(s)|)CN1/4+δ.

Proof.

To simplify the presentation, we set Ψ(x,t):=fσ(1Nj=1,jiNWβ(XjN(t)x)),Ψ¯(x,t):=fσ(1Nj=1,jiNWβ(X¯jN(t)x)), and we write ρ:=ρσ,β,ζ. Taking the difference of EquationEqs. (Equation3) and Equation(Equation4) in the integral formulation leads to (53) sup0<s<t|(XiNX¯iN)(s)|0t|Kζ*(Ψ(XiN(s),s)fσ(Wβ*ρ(X¯iN(s),s)))|ds0t|Kζ*(Ψ(XiN(s),s)Ψ¯(X¯iN(s),s))|ds+0t|Kζ*(Ψ¯(X¯iN(s),s)fσ(Wβ*ρ(X¯iN(s),s)))|ds=:I1+I2.(53)

Step 1: Estimate of I1. To estimate I1, we formulate I1=I11+I12+I13, where I11=0t|Kζ*(Ψ(XiN(s),s)Ψ(X¯iN(s),s))|ds,I12=0t|Kζ*(Ψ(X¯iN(s),s)Ψ¯(XiN(s),s))|ds,I13=0t|Kζ*(Ψ¯(XiN(s),s)Ψ¯(X¯iN(s),s))|ds.

We start with the first integral: I110tD2Kζ*Ψ(·,s)sup0<r<smaxi=1,,N|(XiNX¯iN)(r)|ds. We claim that (54) DkKζ*Ψ(·,s)C(σ)β(k+1)(d+k)1,kN.(54) For the proof, we introduce Φ(x,y):=fσ(1Nj=1N1Wβ(yjx))for xRd,y=(y1,,yN1)R(N1)d. Then, by definition of Kζ, DkKζ*Ψ(·,t)supyRN1Wζ*Kωζ*DkΦ(·,y). We estimate the right-hand side: Wζ*(Kωζ*DkΦ(·,y))Wζ1Kωζ*DkΦ(·,y)CKωζ*DkΦ(·,y)W1,p(Rd)CK*|DkΦ(·,y)|p+CK*|Dk+1Φ(·,y)|pCDkΦ(·,y)r+CDk+1Φ(·,y)r, where we used the Hardy–Littlewood–Sobolev inequality for r=dp/(d+2ps) in the last step. It follows from the Faàdi Bruno formula, after an elementary computation that the last term is estimated according to Dk+1Φ(·,y)rr=Rd|Dk+1(fσ(1Nj=1N1Wβ(yjx)))|rdxC(k,N)max=1,,k+1fσ()rDkWβkrmax0jkRd|Dj+1Wβ(x)|rdxC(k,N)max=1,,k+1fσ()rβ(d+k)krβ(d+k+1)r+dC(k,N,σ)β(d+k)(k+1)rr, since DkWβCβ(d+k) and Dj+1WβrCβ(d+j+1)+d/r. This verifies (Equation54). We infer from (Equation54) with k = 2 that I11Cβ3d70tsup0<r<smaxi=1,,N|(XiNX¯iN)(r)|ds. The term I13 is estimated in a similar way, with Ψ replaced by Ψ¯: I13Cβ3d70tsup0<r<smaxi=1,,N|(XiNX¯iN)(r)|ds. The estimate of the remaining term I12 is more involved. Since Wβ is assumed to be symmetric, we find that I12=|0tRdKζ(y){fσ(1Nj=1,jiNWβ(XjN(s)X¯iN(s)+y))fσ(1Nj=1,jiNWβ(X¯jN(s)XiN(s)+y))}dyds|C0tRdKζ(y)|fσ(1NjiWβ(XjN(s)X¯iN(s)+y))×1Nji(Wβ(XjN(s)X¯iN(s)+y)Wβ(X¯jN(s)XiN(s)+y))+{fσ(1NjiWβ(XjN(s)X¯iN(s)+y))fσ(1NjiWβ(X¯jN(s)XiN(s)+y))}×1NjiWβ(X¯jN(s)XiN(s)+y)|dydsCfσ0tsup0<s<tmaxi=1,,N|(XiNX¯iN)(s)|1NjiRdKζ(y)|D2Wβ(y+ξij(s))|dyds+Cfσ0tsup0<s<tmaxi=1,,N|(XiNX¯iN)(s)| Kζ*Wβds, where ξij(s) is a random value. We write K1=K|B1,K2=K|RdB1 and note that K˜ζK for all ζ>0. Then, RdKζ(y)|D2Wβ(y+ξij(s))|dyB1+ζ(K1*Wζ)(y)|D2Wβ(y+ξij(s))|dy+Rd\B1ζ(K2*Wζ)(y)|D2Wβ(y+ξij(s))|dyK1*WζLθ/(θ1)(B1+ζ)D2Wβ(·+ξij(s))Lθ(B1+ζ)+K2*WζD2Wβ(·+ξij(s))L1(Rd\B1ζ)K1Lθ/(θ1)(B1)D2Wβ(·+ξij(s))Lθ(B1+ζ)+K2D2Wβ(·+ξij(s))L1(Rd)C(D2Wβ+D2Wβ1)Cβd2. Observe that we did not use the compact support for K˜ζ (which depends on ζ), because a negative exponent of ζ at this point would lead to a logarithmic connection between ζ and N in the end, which we wish to avoid.

Furthermore, by the convolution, Sobolev, and Hardy–Littlewood–Sobolev inequalities as well as the fact that |K˜ζ*Wβ|=|(Kwζ)*Wζ*Wβ|K*|Wζ|*|Wβ|, Kζ*Wβ=Wζ*K˜ζ*WβK˜ζ*WβK˜ζ*WβCK˜ζ*WβW1,p(Rd)C(K*|Wβ|pp+K*|D2Wβ|pp)1/pCWβW1,r(Rd)Cβd2+d/r, where we recall that r>d/(2s) and we choose p > d satisfying 1/p=2s/d1/r. The previous two estimates lead to I12C(σ)βd20tsup0<r<smaxi=1,,N|(XiNX¯iN)(r)|ds. We summarize: (55) I1C(σ)β3d70tsup0<r<smaxi=1,,N|(XiNX¯iN)(r)|ds.(55)

Step 2: Estimate of I2. We take the expectation of I2 and use the mean-value theorem: (56) E(I2)=0tE|RdKζ(y){fσ(1NjiWβ(X¯jN(s)X¯iN(s)+y))fσ(Wβ*ρ(X¯iN(s)y,s))}dy|dsN1fσK˜ζ*Wζ10tsupyRdE(ji|bij(y,s)|)ds,(56) where bij(y,s)=Wβ(X¯jN(s)X¯iN(s)+y)NN1Wβ*ρ(X¯iN(s)y,s). We deduce from WζL1(Rd)Cζ1 that K˜ζ*Wζ1Cζ1K˜ζ1Cζ2s1, due to the compact support of K˜ζ(x)=|x|2sdωζ(x)C|x|2sd1|x|2ζ1 and {|x|<2/ζ}|x|2sddx={|y|<2}ζd|y/ζ|2sddy=Cζ2s. We claim that E(ji|bij(y,s)|)C(σ)βd/2N1/2 for all yRd. To show the claim, we compute the expectation E[(jibij(y,s))2]. We estimate first the terms with kj (omitting the argument (y, s) to simplify the notation). Then, an elementary but tedious computation leads to E(bjibki)=RdRdRd(Wβ(xjxi+y)NN1Wβ*ρ(xiy))×(Wβ(xkxi+y)NN1Wβ*ρ(xiy))ρ(xi)ρ(xj)ρ(xk)dxidxjdxk=Rd(Wβ*ρ(xiy)NN1Wβ*ρ(xiy))2ρ(xi)dxiN2ρL(0,T;L(Rd))Wβ*ρL(0,T;L2(Rd))2C(σ)N2Wβ12C(σ)N2. The diagonal terms contribute in the following way: E(bji2)=RdRd(Wβ(xjxi+y)NN1Wβ*ρ(xiy))2ρ(xi)ρ(xj)dxidxj=Rd((Wβ2*ρ)(xiy)2NN1(Wβ*ρ)(xiy)2+N2(N1)2(Wβ*ρ)(xiy)2)ρ(xi)dxiC(σ)(Wβ2*ρL(0,T;L1(Rd))+Wβ*ρL(0,T;L2(Rd))2)C(σ)βd, since Wβ2*ρ2Wβ21ρ2CWβ22βdC. This shows that E(ji|bji(y,s)|)(E[jibji(y,s)]2)1/2C(σ)βd/2N1/2. We infer that (Equation56) becomes (57) I2C(σ)ζ2s1βd/2N1/2.(57)

Step 3: End of the proof. We insert (Equation55) and (Equation57) into (Equation53) to infer that E1(t):=E(sup0<s<tmaxi=1,,N|(XiNX¯iN)(s)|)C(σ)β3d70tE1(s)ds+C(σ)ζ2s1βd/2N1/2. By Gronwall’s lemma, E1(t)C(σ)ζ2s1βd/2N1/2exp(C(σ)β3d7T),0tT. We choose ε=δ˜/(C(σ)T) for some arbitrary δ˜(0,1/4). Then, since by assumption, βd/2β3d7εlogN and ζ2s1C1N1/4, we find that E1(t)C(σ)C1εlog(N)N1/4exp(C(σ)TεlogN)=C1δ˜Tlog(N)N1/4+δ˜, proving the result. □

4.3. Estimate of X¯iNX̂iN

Next, we compute the expectation of X¯iNX̂iN by estimating the difference between Kζ and K as well as the difference between Wβ*ρ(X¯iN) and ρσ(X̂iN). The estimate depends on β and ζ.

Lemma 18.

Let X¯iN and X̂iN be the solutions to (Equation4) and (6), respectively. Then, there exists a constant C > 0, depending on σ, such thatE(sup0<t<Tmaxi=1,,N|(X¯iNX̂iN)(t)|)C(β+ζa),where a:=min{1,d2s}.

Proof.

We compute the difference |(X¯iNX̂iN)(t)|=|0t(Kζ*fσ(Wβ*ρ(X¯iN(s),s))K*fσ(ρσ(X̂iN(s),s)))ds|J1+J2+J3, where ρ:=ρσ,β,ζ, the convolution is taken with respect to xi, and J1=|0tKζ*(fσ(Wβ*ρ(X¯iN(s),s))fσ(Wβ*ρ(X̂iN(s),s)))ds|,J2=|0tKζ*(fσ(Wβ*ρ(X̂iN(s),s))fσ(ρσ(X̂iN(s),s)))ds|,J3=|0t(KζK)*fσ(ρσ(X̂iN(s),s))ds|.

Step 1: Estimate of J1. We write Kζ*fσ()=Kζ*fσ and add and subtract the expression fσ(Wβ*ρ(X¯iNy))Wβ*ρ(X̂iNy): J1=0tRdKζ(y)(fσ(Wβ*ρ(X¯iN(s)y))Wβ*[ρ(X¯iN(s)y)ρ(X̂iN(s)y)][fσ(Wβ*ρ(X̂iN(s)y))fσ(Wβ*ρ(X¯iN(s)y))]Wβ*ρ(X̂iN(s)y))dydsfσ0tRd|Kζ(y)Wβ*(ρ(X¯iN(s)y)ρ(X̂iN(s)y))|dyds+fσWβ*ρL(0,T;L(Rd))×0tRd|Kζ(y)Wβ*(ρ(X̂iN(s)y)ρ(X¯iN(s)y))|dyds. By the mean-value theorem and using Wβ1=1, we obtain for some random variable ξij(s), (58) J1fσW2,(R)ρL(0,T;L(Rd))0tsup0<r<ssupi=1,,N|(X¯iNX̂iN)(r)|×Rdk=12|Kζ(y)DkWβ*ρ(y+ξij(s),s)|dyds.(58) We need to estimate the last integral. For this, we write for k = 1, 2 Rd|Kζ(y)DkWβ*ρ(y+ξij(s),s)|dyK1k+K2k,whereK1k:=B1+ζ|K1*Wζ(y)DkWβ*ρ(y+ξij(s),s)|dy,K2k:=RdB1ζ|K2*Wζ(y)DkWβ*ρ(y+ξij(s),s)|dy, where K1=K|B1 and K2=K|RdB1. Note that K˜ζK. A similar argument as for the estimate of I12 in the proof of Lemma 17 shows that for θ>max{d/(2s),d}, K1k+K2kC(DkWβ*ρL(0,T;Lθ(Rd))+DkWβ*ρL(0,T;L1(Rd)))C(DkρL(0,T;Lθ(Rd))+DkρL(0,T;L1(Rd)))C(σ), where we used Proposition 14 ((Equation39) and (Equation41)) with p=θ in the last inequality. We conclude from (Equation58) that (59) J1C(σ)0tsup0<r<smaxi=1,,N|(X¯iNX̂iN)(r)|ds.(59)

Step 2: Estimate of J2. We treat the two cases s < 1∕2 and s1/2 separately. Let first s1/2. Then, J2=|0tK˜ζ*Wζ*(fσ(Wβ*ρ(X̂iN(s),s))fσ(ρσ(X̂iN(s),s)))ds|TK˜ζ*(fσ(Wβ*ρ)fσ(ρσ))L(0,T;L(Rd)). Recalling the definition of K˜ζ=Kωζ in (Equation11) and writing K˜ζu=Ku[(1ωζ)K]u+[Kωζ]u for u=fσ(Wβ*ρ)fσ(ρσ), we find that (60) J2C(T)(K*uL([0,T];L(Rd))+[(1ωζ)K]*uL([0,T];L(Rd))+[Kωζ]*uL(0,T;L(Rd))).(60) We estimate the right-hand side term by term. Because of K*v={(Δ)1/2vfor s=1/2(K)vfor s>1/2, we use Sobolev’s embedding W1,p(Rd)L(Rd) for any p > d, and then, the boundedness of the Riesz operator (Δ)1/2:Lp(Rd)Lp(Rd) [Citation46, chapter IV, §3.1] in case s = 1/2 or the Hardy–Littlewood–Sobolev inequality for α˜=α1/2>0 (see Lemma 22) in case s > 1∕2 to control the first norm in (Equation60) by K*uL(0,T;L(Rd))C(K*uL(0,T;Lp(Rd))+j=1dK*DjuL(0,T;Lp(Rd)))CuL(0,T;W1,r(Rd))=Cfσ(Wβ*ρ)fσ(ρσ)L(0,T;W1,r(Rd)), where r = p in case s = 1/2 and r=pd/(d+2s1) in case s > 1∕2. Choosing p>d+(2s1) guarantees that r > d always holds.

For the second norm in (Equation60), Hölder’s inequality yields for q > d and 1/q+1/q=1, for every t > 0, [(1ωζ)K]*u(t)L(Rd)1ωζL(Rd)KLq({|x|>2ζ1})u(t)Lq(Rd)KLq({|x|>2ζ1})u(t)Lq(Rd), which can be bounded by Cζ12s+d/qu(t)Lq(Rd), since KLq({|x|>2ζ1})qC{|x|>2ζ1}|x|(2sd1)qdx=Cζd{|y|>2}|y/ζ|(2sd1)qdyCζd+(1+d2s)q. By similar arguments and the fact that ωζLCζ, we find that KωζLq({|x|<2ζ1})Cζ1+d2sd/q, and hence, using q=q/(q1), we conclude for the second and third term in (Equation60) that [(1ωζ)K]*u(t)L(Rd)+[Kωζ]*u(t)L(Rd)Cζ12s+d/qu(t)Lq(Rd). The choice d<qd/(2s1) guarantees on the one hand that q > d and on the other hand that the exponent 12s+d/q is strictly positive (which allows us to use the property ζ12s+d/q<1).

Using these estimates in (Equation60), we arrive (for s1/2) at J2C(T)(fσ(Wβ*ρ)fσ(ρσ)L(0,T;W1,r(Rd))+fσ(Wβ*ρ)fσ(ρσ)L(0,T;Lq(Rd))), where we recall that r,q>d. These norms can be estimated by fσ(Wβ*ρ(t))fσ(ρσ(t))Lq(Rd)fσWβ*ρ(t)ρσ(t)Lq(Rd) and (fσ(Wβ*ρ)fσ(ρσ))(t)Lr(Rd)fσ(Wβ*ρρσ)(t)Lr(Rd)+fσ(Wβρρσ)(t)Lr(Rd)ρσ(t)L(Rd). The L(Rd×(0,T)) bound for ρσ from Lemma 11 and the definition of fσ finally show for s1/2 and r,q>d that (61) J2C(σ,T)(Wβ*ρρσL(0,T;W1,r(Rd))+Wβ*ρρσL(0,T;Lq(Rd))).(61) Now, let s < 1∕2. In this case, we cannot estimate K and put the gradient to the second factor of the convolution. Adding and subtracting an appropriate expression as in Step 1, using the embedding W1,p(Rd)L(Rd) for p > d, the estimate KζK, and the Hardy–Littlewood–Sobolev inequality, we find that J2=|0tRdKζ(y)((fσ(Wβ*ρ(X̂iN(s)y))fσ(ρσ(X̂iN(s)y)))Wβ*ρ(X̂iN(s)y)fσ(ρσ(X̂iN(s)y))(ρσ(X̂iN(s)y)Wβ*ρ(X̂iN(s)y)))dyds|fσWβ*ρ0tRdKζ(y)|ρσ(X̂iN(s)y)Wβ*ρ(X̂iN(s)y)|dyds+fσ0tRdKζ(y)|ρσ(X̂iN(s)y)Wβ*ρ(X̂iN(s)y)|dydsmax{ρL(0,T;L(Rd)),1}fσW1,T(K*|(Wβ*ρρσ)|L(0,T;L(Rd))+K*|(Wβ*ρρσ)|L(0,T;L(Rd)))C(σ,T)(ρL(0,T;L(Rd))+1)|α|2Wβ*DαρDαρσL(0,T;Lr(Rd)), where r > d is such that 1/r=2s/d+1/p (this is needed for the Hardy–Littlewood– Sobolev inequality) and p > d (because of Sobolev’s emebdding). Note that r > d can be only guaranteed if s < 1∕2. Together with the fact that ρL(0,T;L(Rd))C(σ) (choose q > d in (Equation41) and use Sobolev’s embedding), this shows that for s < 1∕2, (62) J2C(σ,T)|α|2Wβ*DαρDαρσL(0,T;Lr(Rd)).(62) It follows from estimate (Equation38) and Lemma 21 in Appendix A for p > d that (Wβ*DαρDαρσ)(t)Lp(Rd)C(DαρLp(Rd)β+β+ζa)C(σ,T)(β+ζa), where we used the L(0,T;W3,p(Rd)) estimate for ρ=ρσ,β,ζ in (Equation41). Then, we deduce from estimates (Equation61) and (Equation62) that for all 0<s<1, J2C(σ,T)(β+ζa), where we recall that a=min{1,d2s}.

Step 3: Estimate of J3 and end of the proof. Arguing similarly as in Subsection 3.3, we have (KζK)*ρσL(0,T;L(Rd))Cζa(D2ρσL(0,T;Lp(Rd))+ρσL(0,T;L1(Rd))). This implies that (63) J3fσ(KζK)*ρσL(0,T;L(Rd))C(σ)ζa.(63) Taking the expectation, we infer from (Equation59) to (Equation63) that E2(t):=E(sup0<s<tmaxi=1,,N|(X¯iNX̂iN)(s)|)C(σ)(β+ζa)+C(σ)0tE2(s)ds, An application of Gronwall’s lemma gives the result. □

4.4. Proof of Theorem 2 and Proposition 3

Lemmas 17 and 18 show that E(sup0<s<Tmaxi=1,,N|(XiNX̂iN)(s)|)C(N1/4+δ+β+ζmin{1,d2s}), and this expression converges to zero as N and (β,ζ)0 under the conditions stated in Theorem 3. This result implies the convergence in probability of the k-tuple (X1N,,XkN) to (X̂1N,,X̂kN). Since convergence in probability implies convergence in distribution, we obtain limN,(β,ζ)0PN,β,σk(t)=Pσk(t)locally uniform in time, where PN,β,σk(t) and Pσk(t) denote the joint distributions of (X1N,,XkN)(t) and (X̂1N,, X̂kN)(t), respectively. By Subsection 4.1, Pσ(t) is absolutely continuous with the density function ρσ(t). Using the test function ϕ=1(,x]d in Corollary 13, we have, up to a subsequence, Pσ(t,(,x]d)=(,x]dρσ(y,t)dy(,x]dρ(y,t)dy=:P(t,(,x]d) locally uniformly for t > 0. Since the convergence also holds for the initial condition, the result is shown.

Additional information

Funding

The second and third authors have been partially supported by the Austrian Science Fund (FWF), grants P30000, P33010, F65, and W1245. The fourth author acknowledges support from the Alexander von Humboldt Foundation. This work has received funding from the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme, ERC Advanced Grant no. 101018153.

References

  • Stinga, P. R. (2019). User’s guide to the fractional Laplacian and the method of semigroups. In: Kochubei, A., Luchko, Y., eds. Handbook of Fractional Calculus with Applications, Vol. 2. Berlin: De Gruyter, pp. 235–266.
  • Chapman, J., Rubinstein, J., Schatzman, M. (1996). A mean-field model for superconducting vortices. Eur. J. Appl. Math. 7(2):97–111. DOI: 10.1017/S0956792500002242.
  • E, W. (1994). Dynamics of vortex liquids in Ginzburg–Landau theories with applications in superconductivity. Phys. Rev. B Condens. Matter. 50(2):1126–1135. DOI: 10.1103/physrevb.50.1126.
  • Biler, P., Imbert, C., Karch, G. (2015). The nonlocal porous medium equation: Barenblatt profiles and other weak solutions. Arch. Rational Mech. Anal. 215(2):497–529. DOI: 10.1007/s00205-014-0786-1.
  • Vázquez, J. L. (2017). The mathematical theories of diffusion. Nonlinear and fractional diffusion. In: Bonforte, M., Grillo, G., eds. Nonlocal and Nonlinear Diffusions and Interactions: New Methods and Directions. Lect. Notes Math. 2186. Cham: Springer, pp. 205–278.
  • Caffarelli, L., Vazquez, J. L. (2011). Nonlinear porous medium flow with fractional potential pressure. Arch. Rational Mech. Anal. 202(2):537–565. DOI: 10.1007/s00205-011-0420-4.
  • Biler, P., Imbert, C., Karch, G. (2011). Barenblatt profiles for a non local porous medium equation. C. R. Acad. Sci. Paris Ser. I 349(1112):641–645. DOI: 10.1016/j.crma.2011.06.003.
  • Stan, D., del Teso, F., Vázquez, J. L. (2014). Finite and infinite speed of propagation for porous medium equations with fractional pressure. C. R. Acad. Sci. Paris Ser. I 352(2):123–128. DOI: 10.1016/j.crma.2013.12.003.
  • Płociniczak, L. (2019). Derivation of the nonlocal pressure form of the fractional porous medium equation in the hydrological setting. Commun. Nonlin. Sci. Numer. Simul. 76:66–70. DOI: 10.1016/j.cnsns.2019.04.014.
  • Duerinckx, M. (2016). Mean-field limits for some Riesz interaction gradient flows. SIAM J. Math. Anal. 48(3):2269–2300. DOI: 10.1137/15M1042620.
  • Serfaty, S. (2016). Mean-field limits of the Gross–Pitaevskii and parabolic Ginzburg–Landau equations. J. Amer. Math. Soc. 30(3):713–768. DOI: 10.1090/jams/872.
  • Olivera, C., Richard, A., Tomašević, M. (2020). Quantitative particle approximation of nonlinear Fokker–Planck equations with singular kernel. Ann. Sc. Norm. Super. Pisa Cl. Sci.
  • Choi, Y.-P., Jeong, I.-J. (2021). Relaxation to fractional porous medium equation from Euler–Riesz system. J. Nonlin. Sci. 31(95):28.
  • Serfaty, S. (2020). Mean field limit for Coulomb-type flows. Duke Math. J. 169(15):2887–2935. DOI: 10.1215/00127094-2020-0019.
  • Daus, E. S., Ptashnyk, M., Raithel, C. (2022). Derivation of a fractional cross-diffusion system as the limit of a stochastic many-particle system driven by Lévy noise. J. Diff. Eqs. 309: 386–426.
  • Caffarelli, L., Vázquez, J. L. (2011). Asymptotic behaviour of a porous medium equation with fractional diffusion. Discrete Contin. Dyn. Sys. 29(4):1393–1404. DOI: 10.3934/dcds.2011.29.1393.
  • Carrillo, J. A., Huang, Y., Santos, M. C., Vazquez, J. L. (2015). Exponential convergence towards stationary states for the 1D porous medium equation with fractional pressure. J. Differ. Equ. 258(3):736–763. DOI: 10.1016/j.jde.2014.10.003.
  • Lisini, S., Mainini, E., Segatti, A. (2018). A gradient flow approach to the porous medium equation with fractional pressure. Arch. Rational Mech. Anal. 227(2):567–606. DOI: 10.1007/s00205-017-1168-2.
  • Caffarelli, L., Soria, F., Vázquez, J. L. (2013). Regularity of solutions of the fractional porous medium flow. J. Eur. Math. Soc. 15(5):1701–1746. DOI: 10.4171/JEMS/401.
  • Imbert, C., Tarhini, R., Vigneron, F. (2020). Regularity of solutions of a fractional porous medium equation. Interfaces Free Bound 22(4):401–442. DOI: 10.4171/IFB/445.
  • Ambrosio, L., Serfaty, S. (2008). A gradient flow approach to an evolution problem arising in superconductivity. Commun. Pure Appl. Math. 61(11):1495–1539. DOI: 10.1002/cpa.20223.
  • Choi, Y.-P., Jeong, I.-J. (2021). Classical solutions for fractional porous medium flow. Nonlin. Anal. 210 (112393):13.
  • Nguyen, Q.-H., Vazquez, J. L. (2018). Porous medium equation with nonlocal pressure in a bounded domain. Commun. Partial Differ. Eqs. 43(10):1502–1539. DOI: 10.1080/03605302.2018.1475492.
  • Caffarelli, L., Gualdani, M., Zamponi, N. (2020). Existence of weak solutions to a continuity equation with space time nonlocal Darcy law. Commun. Partial Differ. Equ. 45(12):1799–1819. DOI: 10.1080/03605302.2020.1814325.
  • Daus, E. S., Gualdani, M., Zamponi, N. (2020). Long time behavior and weak-strong uniqueness for a nonlocal porous media equation. J. Differ. Equ. 268(4):1820–1839. DOI: 10.1016/j.jde.2019.09.029.
  • de Pablo, A., Quiros, F., Rodriguez, A., Vázquez, J. L. (2012). A general fractional porous medium equation. Commun. Pure Appl. Math. 65(9):1242–1284. DOI: 10.1002/cpa.21408.
  • Vázquez, J. L. (2012). Nonlinear diffusion with fractional Laplacian operators. In: Holden, H., Karlsen, K., eds. Nonlinear Partial Differential Equations, Abel Symposia, Vol. 7. Berlin: Springer, pp. 1–27.
  • Golse, F. (2003). The mean-field limit for the dynamics of large particle systems. Journ. Equ. Dériv. Partielles 1–47. DOI: 10.5802/jedp.623.
  • Jabin, P.-E., Wang, Z., Bellomo, N., Degond, P., Tadmor, E. (2017). Mean field limit for stochastic particle systems. In: Active Particles, Vol. 1. Boston: Springer, pp. 379–402.
  • Sznitman, A. S. (1984). Nonlinear reflecting diffusion process, and the propagation of chaos and fluctuations associated. J. Funct. Anal. 56(3):311–336. DOI: 10.1016/0022-1236(84)90080-6.
  • Sznitman, A. S. (1991). Topics in propagation of chaos. In: Hennequin, P. L., ed. École D’Été de Probabilités de Saint-Flour XIX-1989, Lect. Notes Math. 1464. Berlin: Springer, pp. 165–251.
  • Oelschläger, K. (1984). A martingale approach to the law of large numbers for weakly interacting stochastic processes. Ann. Probab. 12:458–479.
  • Oelschläger, K. (1990). Large systems of interacting particles and the porous medium equation. J. Differ. Eqs. 88(2):294–346. DOI: 10.1016/0022-0396(90)90101-T.
  • Jourdain, B., Méléard, S. (1998). Propagation of chaos and fluctuations for a moderate model with smooth initial data. Ann. Inst. H. Poincaré Probab. Stat. 34(6):727–766. DOI: 10.1016/S0246-0203(99)80002-8.
  • Chen, L., Daus, E. S., Holzinger, A., Jüngel, A. (2021). Rigorous derivation of population cross-diffusion systems from moderately interacting particle systems. J. Nonlin. Sci. 31(94):38.
  • Chen, L., Daus, E. S., Jüngel, A. (2019). Rigorous mean-field limit and cross diffusion. Z. Angew. Math. Phys. 70(4):21. DOI: 10.1007/s00033-019-1170-7.
  • Ichikawa, K., Rouzimaimaiti, M., Suzuki, T. (2012). Reaction diffusion equation with non-local term arises as a mean field limit of the master equation. Discrete Contin. Dyn. Syst. S 5(1):115–126. DOI: 10.3934/dcdss.2012.5.115.
  • Fontbona, J., Méléard, S. (2015). Non local Lotka–Volterra system with cross-diffusion in an heterogeneous medium. J. Math. Biol. 70(4):829–854. DOI: 10.1007/s00285-014-0781-z.
  • Figalli, A., Philipowski, R. (2008). Convergence to the viscous porous medium equation and propagation of chaos. Alea 4:185–203.
  • Lisini, S. (2020). Fractional high order thin film equation: gradient flow approach. Preprint available at https://arxiv.org/pdf/2007.00459.
  • Novotný, E. F. A. (2009). Singular Limits in Thermodynamics of Viscous Fluids. Basel: Birkhäuser.
  • Bonforte, M., Sire, Y., Vazquez, J. L. (2015). Existence, uniqueness and asymptotic behaviour for fractional porous medium equations on bounded domains. DCDS-A. 35(12):5725–5767. DOI: 10.3934/dcds.2015.35.5725.
  • Cifani, S., Jakobsen, E. (2011). Entropy solution theory for fractional degenerate convection-diffusion equations. Ann. Inst. H. Poincaré Anal. Non Linéaire 28(3):413–441. DOI: 10.1016/j.anihpc.2011.02.006.
  • Karatzas, I., Shreve, S. (1991). Brownian Motion and Stochastic Calculus, 2nd ed. New York, NY: Springer.
  • Nualart, D. (2006). The Malliavin Calculus and Related Topics. Berlin: Springer.
  • Stein, E. (1970). Singular Integrals and Differentiability Properties of Functions. Princeton: Princeton University Press.
  • Brézis, H. (2011). Functional Analysis, Sobolev Spaces and Partial Differential Equations. New York, NY: Springer.
  • Roubíček, T. (2013). Nonlinear Partial Differential Equations with Applications. Basel: Birkhäuser.
  • Lieberman, G. M. (1996). Second Order Parabolic Differential Equations. Singapore: World Scientific.

Appendix A.

Auxiliary results

We recall some known results. The following result is proved in [Citation47, theorem 4.33].

Lemma 19

(Young’s convolution inequality). Let 1p,r,uLp(Rd),vLq(Rd), and 1/p+1/q=1+1/r. Then, u*vLr(Rd) andu*vrupvq.

The following lemma slightly extends [Citation48, lemma 7.3] from the L2 to the Lp setting.

Lemma 20.

Let p2 and T > 0. Then, the following embedding is continuous:Lp(0,T;W1,p(Rd))W1,p(0,T;W1,p(Rd))C0([0,T];Lp(Rd)).

Proof.

Let uLp(0,T;W1,p(Rd))W1,p(0,T;W1,p(Rd)) and 0t1t2T. Then, (A1) |Rd|u(t2)|pdxRd|u(t1)|pdx|=|t1t2tu,p|u|p2udt|ptuLp(t1,t2;W1,p(Rd))|u|p2uLp(t1,t2;W1,p(Rd)),(A1) where p=p/(p1). Direct computations using Young’s inequality lead to |u|p2uLp(t1,t2;W1,p(Rd))p=Ct1t2Rd(|u|p+|u|p(p2)|u|p)dxdtCt1t2u(t)W1,p(Rd)pdt. We infer from (A1) and the continuity of the integrals with respect to the time integration boundaries that tu(t)p is continuous and (A2) sup0<t<Tu(t)pu(0)p+CtuLp(t1,t2;W1,p(Rd))+CuLp(0,T;W1,p(Rd)).(A2) Next, let t[0,T] be arbitrary and let τn0 as n such that t+τn[0,T]. Estimate (A2) implies that (u(t+τn))nN is bounded in Lp(Rd). Thus, there exists a subsequence (τn) of (τn) such that u(t+τn)v(t) weakly in Lp(Rd) as n for some v(t)Lp(Rd). We can show, using estimate (A2) and dominated convergence for the integral 0TRd(u(t+τn,x)v(t,x))ϕ(t,x)dxfor ϕC0(Rd×(0,T)) that in the limit n 0TRd(u(t,x)v(t,x))ϕ(t,x)dx=0, which yields v(t)=u(t).

Moreover, since tu(t)p is continuous, we have u(t+τn)pu(t)p. Since Lp(Rd) is uniformly convex, we deduce from [Citation47, prop. 3.32] that u(t+τn)u(t) strongly in Lp(Rd). Since the limit is unique, the whole sequence converges. Together with (A2), this concludes the proof. □

Let W1C0(Rd) be nonnegative with RdW1(x)dx=1 and define Wβ(x)=βdW1(x/β) for xRd and β>0.

Lemma 21.

Let 1p< and uW1,p(Rd). Then,Wβ*uupCβup.

Proof.

We use Hölder’s inequality and the fact that WβL1(Rd)=1 to find that Wβ*uupp=Rd|RdWβ(xy)(u(x)u(y))dy|pdxRd(RdWβ(xy)dy)p1(RdWβ(xy)|u(x)u(y)|pdy)dx=RdRdWβ(z)|z|p|u(y+z)u(y)|p|z|pdydzuppRdWβ(z)|z|pdzCβpupp, which shows the lemma. □

Appendix B.

Fractional Laplacian

We recall that the fractional Laplacian (Δ)s for 0<s<1 can be written as the pointwise formula (B1) (Δ)su(x)=cd,sRdu(x)u(y)|xy|d+2sdy,where cd,s=4sΓ(d/2+s)πd/2|Γ(s)|,(B1) uHs(Rd), and the integral is understood as principal value if 1/2s<1 [Citation1, theorem 2]. The inverse fractional Laplacian (Δ)s is defined in (Equation2). The following lemma can be found in [Citation46, chapter V, subsection 1.2].

Lemma 22

(Hardy–Littlewood–Sobolev inequality). Let 0<s<1 and 1<p<. Then, there exists a constant C > 0 such that for all uLp(Rd), (Δ)suqCup,where 1p=1q+2sd.

Applying Hölder’s, and then, Hardy–Littlewood–Sobolev’s inequality gives the following result.

Lemma 23.

Let 0<s<1 and 1p<q<. Then, there exists C > 0 such that for all uLq(Rd),vLr(Rd), (B2) u(Δ)svpCuqvr,1q+1r=1p+2sd,(B2) (B3) u(Δ)svpCuqvr,1q+1r=1p+2s1d,s>12.(B3)

Lemma 24

(Fractional Gagliardo–Nirenberg inequality I). Let d2 and 1<p<. Then, there exists C > 0 such that for all uW1,p(Rd) or uW2,p(Rd), respectively, (Δ)supCup12sup2sif 0<s1/2,(Δ)supCup1sD2upsif 1/2<s1.

Proof.

It follows from the properties of the Riesz and Bessel potentials [Citation46, theorem 3, p. 96] that the operator (Δ)s:W1,p(Rd)Lp(Rd) is bounded for 0<s1/2, while the operator (Δ)s:W2,p(Rd)Lp(Rd) is bounded for 1/2<s1. Thus, if 0<s1/2, (Δ)supC(up+up)for uW1,p(Rd). Replacing u by uλ(x)=λd/p2su(λx) with λ>0 yields (Δ)sup=(Δ)suλpC(uλp+uλp)=Cλ2s(up+λup). We minimize the right-hand side with respect to λ giving the value λ0=2s(12s)1up up1 and therefore, (Δ)supCu12sup2s. The case 1/2<s1 is proved in a similar way. □

Lemma 25

(Fractional Gagliardo–Nirenberg inequality II). Let d2,0<s1/2,p(1,), and q[p,). If p<d/(2s), we assume additionally that qdp/(d2sp). Then, there exists C > 0 such that for all uW1,p(Rd), (Δ)suqCup1θupθ,where θ=1+d/pd/q2s[0,1].

Proof.

The statement is true for s = 1/2 since the operator (Δ)1/2:Lq(Rd)Lq(Rd) is bounded for any q(1,) [Citation46, theorem 3, p. 96]. Then, the inequality follows from the standard Gagliardo–Nirenberg inequality.

Thus, let 0<s<1/2. We claim that it is sufficient to prove that (Δ)s:W1,p(Rd)Lq(Rd) is bounded. Indeed, assume that (B4) (Δ)suqC(up+up)for uW1,p(Rd).(B4) Replacing, as in the proof of Lemma 24, u by uλ(x)=λd/q1+2su(λx) with λ>0 yields (Δ)suqCλθ(up+λup), where θ is defined in the statement of the theorem. Minimizing the right-hand side with respect to λ gives the value λ0=θ(1θ)1upup1 and therefore, (Δ)suqCup1θupθ. It remains to show Equation(B4). To this end, we distinguish two cases. First, let p<d/(2s). By assumption, pqr(1):=dp/(d2sp). We apply the Hardy–Littlewood–Sobolev inequality (Lemma 22) to find that (Δ)sur(1)CupC(up+up). Furthermore, by using (in this order) the boundedness of (Δ)1/2:Lp(Rd)Lp(Rd), lemma 2 in [Citation46, p. 133], Equationeq. (Equation40) in [Citation46, p. 135], and theorem 3 in [Citation46, p. 135f], (B5) (Δ)sup=(Δ)1/2(Δ)1/2supC(Δ)1/2supC(IΔ)1/2supC(IΔ)1/2upC(up+up).(B5) These inequalities hold for any p(1,). Now, it is sufficient to interpolate with 1/q=μ/p+(1μ)/r(1): (Δ)suq(Δ)supμ(Δ)sur(1)1μC(up+up). Second, let pd/(2s). We choose λ(0,d/(2sp))(0,1) and apply the Hardy–Littlewoord– Sobolev inequality: (Δ)sur(λ)=(Δ)λs(Δ)(1λ)sur(λ)C(Δ)(1λ)sup, where r(λ)=dp/(d2sλp). Since (1λ)s<1/2, we deduce from Equation(B5) that (Δ)sur(λ)C(up+up). Since r(λ) as λd/(2sp), the result follows. □

Appendix C.

Parabolic regularity

Lemma 26

(Parabolic regularity). Let 1<p<, T > 0 and let u be the (weak) solution to the heat equationtuΔu=v,u(0)=u0in Rd,where vLp(0,T;Lp(Rd)) and u0W2,p(Rd). Then, there exists C > 0, depending on T and p, such that(C1) tuLp(0,T;Lp(Rd))+D2uLp(0,T;Lp(Rd))C(vLp(0,T;Lp(Rd))+D2u0Lp(Rd)).(C1)

Furthermore, if v=div w for some wLp(0,T;Lp(Rd;Rd)), then,(C2) uLp(0,T;Lp(Rd))C(wLp(0,T;Lp(Rd))+T1/pu0Lp(Rd)).(C2)

Proof.

We use a known result on the parabolic regularity for the equation (C3) tûΔû=v,û(0)=0in Rd.(C3) It holds that [Citation49] (C4) tûLp(0,T;Lp(Rd))+D2ûLp(0,T;Lp(Rd))CvLp(0,T;Lp(Rd)).(C4) We apply this result to û=uetΔu0, where etΔu0 is the solution to the homogeneous heat equation in Rd with initial datum u0. Then, û solves Equation(C3) and satisfies estimate Equation(C4). Inserting the definition of û and observing that D2(etΔu0)pCD2u0p, we obtain Equation(C1).

If v=div w for some wLp(0,T;Lp(Rd;Rd)), the uniqueness of solutions to the heat equation yields u=etΔu0+div U, where U solves tUΔU=w,U(0)=0in Rd. Then, we deduce from the regularity result of [Citation49] with û=U and v = w that D2ULp(0,T;Lp(Rd))CwLp(0,T;Lp(Rd)). Since u=etΔu0+div U, inequality Equation(C2) follows. □