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

Large friction-High force fields limit for the nonlinear Vlasov–Poisson–Fokker–Planck system

José A. Carrillo
Mathematical Institute
University of Oxford, Oxford OX2 6GG, UK.
[email protected]
Young-Pil Choi
Department of Mathematics
Yonsei University, 50 Yonsei-Ro, Seodaemun-Gu, Seoul 03722, Republic of Korea.
[email protected]
 and  Yingping Peng
School of Mathematical Sciences
University of Electronic Science and Technology of China, Chengdu, 611731, China.
yingping_\_peng163.com
Abstract.

We provide a quantitative asymptotic analysis for the nonlinear Vlasov–Poisson–Fokker–Planck system with a large linear friction force and high force-fields. The limiting system is a diffusive model with nonlocal velocity fields often referred to as aggregation-diffusion equations. We show that a weak solution to the Vlasov–Poisson–Fokker–Planck system strongly converges to a strong solution to the diffusive model. Our proof relies on the modulated macroscopic kinetic energy estimate based on the weak-strong uniqueness principle together with a careful analysis of the Poisson equation.

Key words and phrases:
Vlasov–Poisson–Fokker–Planck equation, aggregation-diffusion equation, relative entropy, modulated energy, high force-field limit.

1. Introduction

We are concerned in this work with the asymptotic analysis of the nonlinear Vlasov–Poisson–Fokker–Planck (in short, VPFP) system with a linear damping in a high force-field regime. More precisely, let f=f(x,v,t)f=f(x,v,t) be the particle, electron for instance, distribution function in the phase space (x,v)(x,v) at time t>0t>0. Our main system reads as

tf+vxf1mev((v+xV+xΦ)f)=1τe𝒩FP(f),(x,v,t)d×d×+,ΔxΦ=ρ,ρ(x,t)=df(x,v,t)𝑑v,\displaystyle\begin{aligned} &\partial_{t}f+v\cdot\nabla_{x}f-\frac{1}{m_{e}}\nabla_{v}\cdot((v+\nabla_{x}V+\nabla_{x}\Phi)f)=\frac{1}{\tau_{e}}\mathcal{N}_{FP}(f),\quad(x,v,t)\in\mathbb{R}^{d}\times\mathbb{R}^{d}\times\mathbb{R}_{+},\cr&-\Delta_{x}\Phi=\rho,\quad\rho(x,t)=\int_{\mathbb{R}^{d}}f(x,v,t)\,dv,\end{aligned} (1.1)

where V:d+V:\mathbb{R}^{d}\to\mathbb{R}_{+} represents the confinement potential, and 𝒩FP\mathcal{N}_{FP} denotes the nonlinear Fokker–Planck operator [33] given by

𝒩FP(f):=v((vu)f+σevf),where u:=dvf𝑑vdf𝑑v.\mathcal{N}_{FP}(f):=\nabla_{v}\cdot\left((v-u)f+\sigma_{e}\nabla_{v}f\right),\quad\mbox{where }u:=\frac{\int_{\mathbb{R}^{d}}vf\,dv}{\int_{\mathbb{R}^{d}}f\,dv}.

Here the positive constants mem_{e} and τe\tau_{e} represent the mass of the electrons and the relaxation time, respectively, and σe=(kBTth)/me\sqrt{\sigma_{e}}=\sqrt{(k_{B}T_{th})/m_{e}} denotes the thermal velocity, with the Planck constant kBk_{B} and the temperature of the thermal bath TthT_{th}. Other physical constants are taken to be unity for simplicity. The first term in 𝒩FP(f)\mathcal{N}_{FP}(f) describes the nonlinear relaxation towards the local velocity uu, and it can be rigorously derived from velocity alignment forces [25].

For the system (1.1), we are interested in the common large friction and high force fields regime. To be more specific, we consider a small mass of electrons mem_{e} and a fast relaxation time τe\tau_{e}. In particular, for sufficiently small ε(0,1)\varepsilon\in(0,1), we choose me=εm_{e}=\varepsilon, Tth1T_{th}\simeq 1, and τe=ε2+δ\tau_{e}=\varepsilon^{2+\delta} for some δ>0\delta>0. Due to the relaxation between mem_{e} and TthT_{th}, we obtain σe=kBε1\sigma_{e}=k_{B}\varepsilon^{-1}, and by setting kB=1k_{B}=1, we deduce from (1.1) that

tf+vxf1εv((v+xV+xΦ)f)=1ε2+δv((vu)f+1εvf),ΔxΦ=ρ.\displaystyle\begin{aligned} &\partial_{t}f+v\cdot\nabla_{x}f-\frac{1}{\varepsilon}\nabla_{v}\cdot((v+\nabla_{x}V+\nabla_{x}\Phi)f)=\frac{1}{\varepsilon^{2+\delta}}\nabla_{v}\cdot\left((v-u)f+\frac{1}{\varepsilon}\nabla_{v}f\right),\cr&-\Delta_{x}\Phi=\rho.\end{aligned} (1.2)

Here we omitted the ε\varepsilon-dependence of solutions ff for the sake of notational simplicity. For the rest of this paper, without loss of generality, we assume that the particle distribution function ff has a unit mass, i.e. d×df𝑑x𝑑v=1\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f\,dxdv=1 for all times due to the conservation of mass.

We study the asymptotic behavior of the nonlinear VPFP system (1.2) as ε0\varepsilon\to 0. To be more specific, we will show that the VPFP system (1.2) converges towards the following drift-diffusion model with nonlocal velocity fields often referred to as aggregation-diffusion equations:

tρ¯+x(ρ¯u¯)=0,ρ¯u¯=ρ¯(xV+xΦ¯+xlogρ¯),ΔxΦ¯=ρ¯.\partial_{t}\bar{\rho}+\nabla_{x}\cdot(\bar{\rho}\bar{u})=0,\quad\bar{\rho}\bar{u}=-\bar{\rho}(\nabla_{x}V+\nabla_{x}\bar{\Phi}+\nabla_{x}\log\bar{\rho}),\quad-\Delta_{x}\bar{\Phi}=\bar{\rho}. (1.3)

Note that if we define the free energy E[ρ¯]E[\bar{\rho}] associated to the above by

E[ρ¯]=dρ¯logρ¯dx+dVρ¯𝑑x+12dΦ¯ρ¯𝑑x,E[\bar{\rho}]=\int_{\mathbb{R}^{d}}\bar{\rho}\log\bar{\rho}\,dx+\int_{\mathbb{R}^{d}}V\bar{\rho}\,dx+\frac{1}{2}\int_{\mathbb{R}^{d}}\bar{\Phi}\bar{\rho}\,dx,

the velocity field u¯\bar{u} of equation (1.3) can be rewritten as

u¯=xδEδρ¯[ρ¯],\bar{u}=-\nabla_{x}\frac{\delta E}{\delta\bar{\rho}}[\bar{\rho}],

where δEδρ¯\frac{\delta E}{\delta\bar{\rho}} denotes the variation of the free energy EE with respect to the mass density ρ¯\bar{\rho}. This shows that the equation (1.3) has a gradient flow structure [23, 8, 1, 6].

1.1. Formal derivation

Let us explain how we can derive the equations (1.3) as the limiting equation of (1.2) as ε0\varepsilon\to 0 at the formal level. If we define local particle density and moment by

ρ:=df𝑑vandρu:=dvf𝑑v,\rho:=\int_{\mathbb{R}^{d}}f\,dv\quad\mbox{and}\quad\rho u:=\int_{\mathbb{R}^{d}}vf\,dv, (1.4)

then by taking into account zero and first moments on ff of (1.2), we find that ρ\rho and uu satisfy

tρ+x(ρu)=0,t(ρu)+x(ρuu)+1εxρ=1ερu1ερ(xV+xΦ)+e,ΔxΦ=ρ,\displaystyle\begin{aligned} &\partial_{t}\rho+\nabla_{x}\cdot(\rho u)=0,\cr&\partial_{t}(\rho u)+\nabla_{x}\cdot(\rho u\otimes u)+\frac{1}{\varepsilon}\nabla_{x}\rho=-\frac{1}{\varepsilon}\rho u-\frac{1}{\varepsilon}\rho(\nabla_{x}V+\nabla_{x}\Phi)+e,\cr&-\Delta_{x}\Phi=\rho,\end{aligned} (1.5)

where ee is given by

e:=x(d(uuvv+1ε𝕀)f𝑑v).e:=\nabla_{x}\cdot\left(\int_{\mathbb{R}^{d}}\left(u\otimes u-v\otimes v+\frac{1}{\varepsilon}\mathbb{I}\right)f\,dv\right). (1.6)

If we set a local Maxwellian

Mu(x)ε(v):=εd/2(2π)d/2exp(ε|u(x)v|22),M^{\varepsilon}_{u(x)}(v):=\frac{\varepsilon^{d/2}}{(2\pi)^{d/2}}\exp\left(-\frac{\varepsilon|u(x)-v|^{2}}{2}\right),

then

dMu(x)ε(v)𝑑v=1for all xd\int_{\mathbb{R}^{d}}M^{\varepsilon}_{u(x)}(v)\,dv=1\quad\mbox{for all }x\in\mathbb{R}^{d}

and

𝒩FP(f)=1εv(fvlogfMuε).\mathcal{N}_{FP}(f)=\frac{1}{\varepsilon}\nabla_{v}\cdot\left(f\nabla_{v}\log\frac{f}{M^{\varepsilon}_{u}}\right).

On the other hand, for 0<ε10<\varepsilon\ll 1, we expect from (1.2) that 𝒩FP(f)0\mathcal{N}_{FP}(f)\simeq 0. This infers

f(x,v)ρ(x)Mu(x)ε(v)=εd/2(2π)d/2ρ(x)exp(ε|u(x)v|22)for0<ε1.f(x,v)\simeq\rho(x)M^{\varepsilon}_{u(x)}(v)=\frac{\varepsilon^{d/2}}{(2\pi)^{d/2}}\rho(x)\exp\left(-\frac{\varepsilon|u(x)-v|^{2}}{2}\right)\quad\mbox{for}\quad 0<\varepsilon\ll 1.

Moreover we notice that

d(vv)exp(ε|u(x)v|22)𝑑v\displaystyle\int_{\mathbb{R}^{d}}(v\otimes v)\exp\left(-\frac{\varepsilon|u(x)-v|^{2}}{2}\right)\,dv =d(uu)exp(ε|u(x)v|22)𝑑v\displaystyle=\int_{\mathbb{R}^{d}}(u\otimes u)\exp\left(-\frac{\varepsilon|u(x)-v|^{2}}{2}\right)\,dv
+d(vu)(vu)exp(ε|u(x)v|22)𝑑v\displaystyle\quad+\int_{\mathbb{R}^{d}}(v-u)\otimes(v-u)\exp\left(-\frac{\varepsilon|u(x)-v|^{2}}{2}\right)\,dv
=(εd/2(2π)d/2)1((uu)+1ε𝕀).\displaystyle=\left(\frac{\varepsilon^{d/2}}{(2\pi)^{d/2}}\right)^{-1}\left((u\otimes u)+\frac{1}{\varepsilon}\mathbb{I}\right).

Thus we obtain

ex(ρd(uuvv+1ε𝕀)Muε𝑑v)=0e\simeq\nabla_{x}\cdot\left(\rho\int_{\mathbb{R}^{d}}\left(u\otimes u-v\otimes v+\frac{1}{\varepsilon}\mathbb{I}\right)M^{\varepsilon}_{u}\,dv\right)=0

for 0<ε10<\varepsilon\ll 1. Putting this into (1.5) yields

tρ+x(ρu)=0,\displaystyle\partial_{t}\rho+\nabla_{x}\cdot(\rho u)=0,
t(ρu)+x(ρuu)+1εxρ=1ερu1ερ(xV+xΦ),\displaystyle\partial_{t}(\rho u)+\nabla_{x}\cdot(\rho u\otimes u)+\frac{1}{\varepsilon}\nabla_{x}\rho=-\frac{1}{\varepsilon}\rho u-\frac{1}{\varepsilon}\rho(\nabla_{x}V+\nabla_{x}\Phi),
ΔxΦ=ρ\displaystyle-\Delta_{x}\Phi=\rho

for 0<ε10<\varepsilon\ll 1, and again we deduce from the momentum equation in the above system that

ρuρ(xV+xΦ)xρ=ρ(xV+xΦ+xlogρ)\rho u\simeq-\rho(\nabla_{x}V+\nabla_{x}\Phi)-\nabla_{x}\rho=-\rho(\nabla_{x}V+\nabla_{x}\Phi+\nabla_{x}\log\rho)

for 0<ε10<\varepsilon\ll 1. This reduces to our main limiting equations (1.3).

Remark 1.1.

The solution of the Poisson equation Φ\Phi can be uniquely expressed as a convolution KρK\star\rho, i.e. Φ=Kρ\Phi=K\star\rho, where the Coulomb interaction potential KK is explicitly given by

K(x)={|x|2for d=1,12πlog|x|for d=2,1(d2)|B(0,1)|1|x|d2for d3,K(x)=\left\{\begin{array}[]{ll}-\frac{|x|}{2}&\textrm{for $d=1$,}\\[5.69054pt] -\frac{1}{2\pi}\log|x|&\textrm{for $d=2$,}\\[5.69054pt] \frac{1}{(d-2)|B(0,1)|}\frac{1}{|x|^{d-2}}&\textrm{for $d\geq 3$},\end{array}\right.

where |B(0,1)||B(0,1)| denotes the volume of unit ball B(0,1)B(0,1) in d\mathbb{R}^{d}, i.e. |B(0,1)|=πd/2/Γ(d/2+1)|B(0,1)|=\pi^{d/2}/\Gamma(d/2+1). Here Γ\Gamma is the Gamma function.

1.2. Literature review

In the absence of the nonlinear Fokker–Planck operator, the asymptotic analysis for the kinetic equation (1.1) with regular interaction forces in the high force-field regime, i.e. (1.1) with me=εm_{e}=\varepsilon and 𝒩FP0\mathcal{N}_{FP}\equiv 0, is first studied in [22] and refined in [20]. To be more precise, the following aggregation equation can be rigorously derived from (1.1) with replacing xΦ\nabla_{x}\Phi by xK~ρ\nabla_{x}\tilde{K}\star\rho and taking me=εm_{e}=\varepsilon and 𝒩FP0\mathcal{N}_{FP}\equiv 0 as ε0\varepsilon\to 0:

tρ¯+x(ρ¯u¯)=0,ρ¯u¯=ρ¯(xV+xK~ρ¯)\partial_{t}\bar{\rho}+\nabla_{x}\cdot(\bar{\rho}\bar{u})=0,\quad\bar{\rho}\bar{u}=-\bar{\rho}(\nabla_{x}V+\nabla_{x}\tilde{K}\star\bar{\rho}) (1.7)

under suitable regularity assumptions on K~\tilde{K} and VV. In [20, 22], types of compactness arguments are employed, and thus the convergences of solutions are only qualitatively investigated. On the other hand, in a recent paper [3] a new idea of establishing the large friction limit of kinetic equation to (1.7) is proposed. Introducing an intermediate system, which is pressureless Euler-type system, and employing the second order Wasserstein distance between ρ\rho and ρ¯\bar{\rho}, a quantified high force-field limit for the equation (1.1) with me=εm_{e}=\varepsilon, τe=ε\tau_{e}=\varepsilon, and σe=0\sigma_{e}=0 is provided. However, a rather strong regularity for the interaction potential WW is still imposed, and thus singular interactions cannot be covered.

In case σe>0\sigma_{e}>0, i.e. with diffusive force, there are several works on the overdamped limit for the linear Vlasov–Fokker–Planck-type equation, i.e. it is a form of (1.1) with the linear diffusion Δvf\Delta_{v}f instead of 𝒩FP(f)\mathcal{N}_{FP}(f) on the right hand side. In [18, 19], the rigorous passage from the linear Vlasov–Fokker–Planck-type equation to the diffusive model is established by using a variational technique. In particular, in [18], a quantitative estimate is obtained when there is no nonlocal interaction forces. On the other hand, nonlocal interactions are considered in [19], and a qualitative error estimate is investigated based on compactness arguments combined with duality methods. Very recently, in [15] the quantified overdamped limit for the Vlasov–Poisson–Fokker–Planck equation with linear diffusion and singular interaction forces is analysed. In particular, either attractive or repulsive Coulomb potential can be taken into account, and the equation (1.3) with the associated velocity fields is rigorously derived. The argument used in [15] relies on the evolution-variational-like inequality for Wasserstein gradient flows, and the convergence of ρ\rho towards ρ¯\bar{\rho} is obtained in the Wasserstein distance of order 2. We also refer to [5, 9, 16, 27, 28] for the rigorous derivation of aggregation-diffusion-type equations from Euler-type system through a singular limit and [21, 29, 30] for the other singular limits of the VPFP system with linear diffusion, namely the parabolic and hyperbolic scalings.

1.3. Contribution

In the present work, we establish quantitative strong convergences of the kinetic equation (1.2) towards the aggregation-diffusion equation (1.3) when ε0\varepsilon\to 0. We identify the asymptotic regime where the VPFP system (1.1) well approximates the aggregation-diffusion equation (1.3); weak solutions to the system (1.2) strongly converge towards the unique strong solution to the equation (1.3). We would like to emphasize that these quantitative strong convergences have not been investigated so far, up to our best knowledge. We show the L1L^{1}-convergence of (ρ,ρu)(\rho,\rho u) towards (ρ¯,ρ¯u¯)(\bar{\rho},\bar{\rho}\bar{u}), in particular this implies the almost everywhere convergence up to a subsequence. We also would like to emphasize that the asymptotic regime treated here differs substantially from the high field or hyperbolic scaling regime typically considered to study the asymptotic analysis of Vlasov–Fokker–Planck-type equations with linear relaxation operator [21, 29]. To the best of our knowledge, up to now it has not been observed that the derivation of aggregation-diffusion equations can be established via the combined large friction-high force-field limit.

In [15, 18, 19], an intermediate equation via a coarse-graining map is introduced for the overdamped limit for the Vlasov–Fokker–Planck-type equation. On the other hand, in case σe=0\sigma_{e}=0 as mentioned above the pressureless Euler-type system is considered as the associated intermediate system for (1.1) in [3]. However, these strategies are not available for our problem due to the nonlocal alignment force term in the operator 𝒩FP\mathcal{N}_{FP}. One can think of the isothermal Euler-type system as the intermediate system and follow the methodology proposed in [3]. In this case, it seems hard to have the uniform-in-ε\varepsilon Lipschitz estimate on the velocity field due to the presence of pressure, see [3, Section 3.1]. For that reason, we directly estimate the error between two equations (1.2) and (1.3) without introducing the intermediate system. We first reformulate the limiting equation (1.3) as the conservative form, isothermal Euler-type system, with an error term. We then employ the relative entropy method. This resembles the strategy used in [9, 27, 28], however careful analysis on the entropy and the error term is needed. In particular, in [9, 28], LL^{\infty}-bound assumption on u¯\bar{u} is used, and this makes some estimates regarding the error term comfortable. However, due to the presence of the confinement potential VV, it seems impossible to impose that assumption in our case when we consider the quadratic confinement potential V=|x|2/2V=|x|^{2}/2. In order to handle this issue, we observe a cancellation structure and use the entropy inequality, which requires higher-order regularity of solutions. See Lemma 3.1 and Remark 3.1 below for more detailed discussion. Moreover, we provide the required existence theories and needed estimates for the equations (1.2) and (1.3) to make all of our results self-contained and fully rigorous.

1.4. Notation

Let us introduce several notations used throughout the present work. For functions, f(x,v)f(x,v) and g(x)g(x), fLp\|f\|_{L^{p}} and gLp\|g\|_{L^{p}} represent the usual Lp(d×d)L^{p}(\mathbb{R}^{d}\times\mathbb{R}^{d})-and Lp(d)L^{p}(\mathbb{R}^{d})-norms, respectively. We denote by CC a generic positive constant which is independent of ε\varepsilon. C=C(α,β,)C=C(\alpha,\beta,\cdots) stands for a positive constant depending on α,β,\alpha,\beta,\cdots. fgf\lesssim g represents that there exists a positive constant C>0C>0 such that fCgf\leq Cg. For simplicity of notation, we often drop xx-dependence of differential operators, i.e. f:=xf\nabla f:=\nabla_{x}f and Δf:=Δxf\Delta f:=\Delta_{x}f whenever there is no any confusion. For any nonnegative integer kk and p[1,]p\in[1,\infty], Wk,p=Wk,p(d)W^{k,p}=W^{k,p}(\mathbb{R}^{d}) stands for the kk-th order LpL^{p} Sobolev space. In particular, if p=2p=2, we denote by Hk=Hk(d)=Wk,2(d)H^{k}=H^{k}(\mathbb{R}^{d})=W^{k,2}(\mathbb{R}^{d}). 𝒞k([0,T];E)\mathcal{C}^{k}([0,T];E) is the set of kk-times continuously differentiable functions from an interval [0,T][0,T]\subset\mathbb{R} into a Banach space EE, and Lp(0,T;E)L^{p}(0,T;E) is the set of measurable functions from an interval (0,T)(0,T) to a Banach space EE, whose pp-th power of the EE-norm is Lebesgue measurable. k\nabla^{k} stands for any partial derivative α\partial^{\alpha} with multi-index α\alpha, |α|=k|\alpha|=k.

1.5. Main result

We first introduce notions of solutions to the VPFP equations (1.2) and the aggregation-diffusion equations (1.3) below.

Definition 1.1.

For given T(0,)T\in(0,\infty), ff is a weak solution of (1.2) on the time interval [0,T][0,T] if and only if the following conditions are satisfied:

  • (i)

    fL(0,T;(L+1L)(d×d))f\in L^{\infty}(0,T;(L^{1}_{+}\cap L^{\infty})(\mathbb{R}^{d}\times\mathbb{R}^{d})),

  • (ii)

    for any φ𝒞c(d×d×[0,T])\varphi\in\mathcal{C}^{\infty}_{c}(\mathbb{R}^{d}\times\mathbb{R}^{d}\times[0,T]),

    0td×df(sφ+vφ1ε(v+V+Φ)vφ)𝑑x𝑑v𝑑s\displaystyle\int_{0}^{t}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f\left(\partial_{s}\varphi+v\cdot\nabla\varphi-\frac{1}{\varepsilon}(v+\nabla V+\nabla\Phi)\cdot\nabla_{v}\varphi\right)dxdvds
    +1ε2+δ0td×df((uv)vφ+1εΔvφ)𝑑x𝑑v𝑑s=d×df0φ(x,v,0)𝑑x𝑑v.\displaystyle\quad+\frac{1}{\varepsilon^{2+\delta}}\int_{0}^{t}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f\left((u-v)\cdot\nabla_{v}\varphi+\frac{1}{\varepsilon}\Delta_{v}\varphi\right)\,dxdvds=\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f_{0}\varphi(x,v,0)\,dxdv.

Here L+1(d×d)L^{1}_{+}(\mathbb{R}^{d}\times\mathbb{R}^{d}) denotes a set of nonnegative L1(d×d)L^{1}(\mathbb{R}^{d}\times\mathbb{R}^{d})-functions.

Definition 1.2.

For given T(0,)T\in(0,\infty), ρ¯\bar{\rho} is a strong solution of (1.3) on the time interval [0,T][0,T] if and only if the following conditions are satisfied:

  • (i)

    ρ¯𝒞([0,T];LV1(d))L(0,T;W1,1W1,(d))\bar{\rho}\in\mathcal{C}([0,T];L^{1}_{V}(\mathbb{R}^{d}))\cap L^{\infty}(0,T;W^{1,1}\cap W^{1,\infty}(\mathbb{R}^{d})), logρ¯L(0,T;W2,(d))\nabla\log\bar{\rho}\in L^{\infty}(0,T;W^{2,\infty}(\mathbb{R}^{d})),

  • (ii)

    (ρ¯,u¯)(\bar{\rho},\bar{u}) satisfies (1.3) in the sense of distributions.

Here LV1(d)L^{1}_{V}(\mathbb{R}^{d}) denotes the space of weighted measurable functions by 1+V1+V with the norm

ρ¯LV1:=d(1+V)ρ¯𝑑x.\|\bar{\rho}\|_{L^{1}_{V}}:=\int_{\mathbb{R}^{d}}(1+V)\bar{\rho}\,dx.
Remark 1.2.

The regularity assumption on ρ¯\bar{\rho} in Definition 1.2 (i) can be replaced by

ρ¯𝒞([0,T];LV1(d))L(0,T;W3,1W3,(d)),\bar{\rho}\in\mathcal{C}([0,T];L^{1}_{V}(\mathbb{R}^{d}))\cap L^{\infty}(0,T;W^{3,1}\cap W^{3,\infty}(\mathbb{R}^{d})),

see Section 3.4 for details.

Since the global-in-time existence of weak solution for the nonlinear VPFP equations (1.1) in the sense of Definition 1.1 is obtained in [4], we do not give any details on that here. We refer to [12, 24] for the global-in-time existence of weak and strong solutions for (1.1) with V0V\equiv 0 and Φ0\Phi\equiv 0. For the existence theory for the Vlasov–Poisson equation with the linear diffusion, i.e. (1.1) with Δvf\Delta_{v}f instead of 𝒩FP(f)\mathcal{N}_{FP}(f), we refer to [2, 10, 31, 32]. Qualitative properties and existence of solutions to aggregation-diffusion equations have been studied under different assumptions [8, 1, 6] and the references therein.

We next define a modulated energy 𝖧ε=𝖧ε((ρ,u)|(ρ¯,u¯))\mathsf{H}_{\varepsilon}=\mathsf{H}_{\varepsilon}((\rho,u)|(\bar{\rho},\bar{u})) by

𝖧ε((ρ,u)|(ρ¯,u¯)):=12dρ|uu¯|2𝑑x+1εdρ¯ρρzz𝑑z𝑑x+12εd|(ΦΦ¯)|2𝑑x.\mathsf{H}_{\varepsilon}((\rho,u)|(\bar{\rho},\bar{u})):=\frac{1}{2}\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dx+\frac{1}{\varepsilon}\int_{\mathbb{R}^{d}}\int_{\bar{\rho}}^{\rho}\frac{\rho-z}{z}\,dzdx+\frac{1}{2\varepsilon}\int_{\mathbb{R}^{d}}|\nabla(\Phi-\bar{\Phi})|^{2}\,dx.

We now state our main theorem.

Theorem 1.1.

Let T>0T>0. Let ff be the solution to the equation (1.2) in the sense of Definition 1.1 and ρ¯\bar{\rho} be the strong solution to the equation (1.3) in the sense of Definition 1.2 on the time interval [0,T][0,T]. Denote by

M0=d(df0|v|22𝑑vρ0|u0|2)𝑑xM_{0}=\int_{\mathbb{R}^{d}}\left(\int_{\mathbb{R}^{d}}f_{0}\frac{|v|^{2}}{2}\,dv-\rho_{0}|u_{0}|^{2}\right)dx

and

M¯0=d(df0logf0dvρ0logρ0)𝑑x,\bar{M}_{0}=\int_{\mathbb{R}^{d}}\left(\int_{\mathbb{R}^{d}}f_{0}\log f_{0}\,dv-\rho_{0}\log\rho_{0}\right)dx\,,

the nonnegative difference between the mesoscopic and the macroscopic initial kinetic energy and entropy for (1.2). Then for ε(0,1)\varepsilon\in(0,1) small enough we have the following estimates.

  • (i)

    Case with the confinement (V(x)=|x|2/2V(x)=|x|^{2}/2):

    𝖧ε(t)+12ε0tdρ|uu¯|2𝑑x𝑑s\displaystyle\mathsf{H}_{\varepsilon}(t)+\frac{1}{2\varepsilon}\int_{0}^{t}\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dxds\leq C𝖧ε(0)+CM0+Cεδε(f0)+Cεδd/21+Cε+CεM¯0.\displaystyle\,C\mathsf{H}_{\varepsilon}(0)+CM_{0}+C\varepsilon^{\delta}\mathcal{F}_{\varepsilon}(f_{0})+C\varepsilon^{\delta-d/2-1}+C\varepsilon+\frac{C}{\varepsilon}\bar{M}_{0}.

    In particular, this implies

    dρ¯ρρzz𝑑z𝑑x+d|(ΦΦ¯)|2𝑑x+0tdρ|uu¯|2𝑑x𝑑sCε𝖧ε(0)+CεM0+Cεδ+1e2εδTε(f0)+Cεδd/2+Cε2+CM¯0.\displaystyle\begin{aligned} &\int_{\mathbb{R}^{d}}\int_{\bar{\rho}}^{\rho}\frac{\rho-z}{z}\,dz\,dx+\int_{\mathbb{R}^{d}}|\nabla(\Phi-\bar{\Phi})|^{2}\,dx+\int_{0}^{t}\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dxds\cr&\qquad\qquad\leq C\varepsilon\mathsf{H}_{\varepsilon}(0)+C\varepsilon M_{0}+C\varepsilon^{\delta+1}e^{2\varepsilon^{\delta}T}\mathcal{F}_{\varepsilon}(f_{0})+C\varepsilon^{\delta-d/2}+C\varepsilon^{2}+C\bar{M}_{0}.\end{aligned}
  • (ii)

    Case without the confinement (V0V\equiv 0):

    𝖧ε(t)+14ε0tdρ|uu¯|2𝑑x𝑑sC𝖧ε(0)+CM0+Cε2+2δEε(f0)+Cε+Cεδ1+CεM¯0.\displaystyle\mathsf{H}_{\varepsilon}(t)+\frac{1}{4\varepsilon}\int_{0}^{t}\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dxds\leq C\mathsf{H}_{\varepsilon}(0)+CM_{0}+C\varepsilon^{2+2\delta}E_{\varepsilon}(f_{0})+C\varepsilon+C\varepsilon^{\delta-1}+\frac{C}{\varepsilon}\bar{M}_{0}.

    In particular, this implies

    dρ¯ρρzz𝑑z𝑑x+d|(ΦΦ¯)|2𝑑x+0tdρ|uu¯|2𝑑x𝑑sCε𝖧ε(0)+CεM0+Cε3+2δEε(f0)+Cε2+Cεδ+CM¯0.\displaystyle\begin{aligned} &\int_{\mathbb{R}^{d}}\int_{\bar{\rho}}^{\rho}\frac{\rho-z}{z}\,dz\,dx+\int_{\mathbb{R}^{d}}|\nabla(\Phi-\bar{\Phi})|^{2}\,dx+\int_{0}^{t}\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dxds\cr&\qquad\qquad\leq C\varepsilon\mathsf{H}_{\varepsilon}(0)+C\varepsilon M_{0}+C\varepsilon^{3+2\delta}E_{\varepsilon}(f_{0})+C\varepsilon^{2}+C\varepsilon^{\delta}+C\bar{M}_{0}.\end{aligned}

Here C>0C>0 is independent of ε>0\varepsilon>0,

ε(f0):=1εd×df0logf0dxdv+d×df0|v|22𝑑x𝑑v+1εd(V+12Φ0)ρ0𝑑x,\mathcal{F}_{\varepsilon}(f_{0}):=\frac{1}{\varepsilon}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f_{0}\log f_{0}\,dxdv+\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f_{0}\frac{|v|^{2}}{2}\,dxdv+\frac{1}{\varepsilon}\int_{\mathbb{R}^{d}}\left(V+\frac{1}{2}\Phi_{0}\right)\rho_{0}\,dx,

and

Eε(f0):=d×df0|v|22𝑑x𝑑v+12εdρ0Φ0𝑑x.E_{\varepsilon}(f_{0}):=\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f_{0}\frac{|v|^{2}}{2}dxdv+\frac{1}{2\varepsilon}\int_{\mathbb{R}^{d}}\rho_{0}\Phi_{0}\,dx. (1.8)
Corollary 1.1.

Suppose that all the assumptions in Theorem 1.1 hold. Furthermore we assume that

d(df0|v|22𝑑vρ0|u0|2)𝑑xC\int_{\mathbb{R}^{d}}\left(\int_{\mathbb{R}^{d}}f_{0}\frac{|v|^{2}}{2}\,dv-\rho_{0}|u_{0}|^{2}\right)dx\leq C

and

dρ¯0ρ0ρ0zz𝑑z𝑑x+d(df0logf0dvρ0logρ0)𝑑x+d|(Φ0Φ¯0)|2𝑑xCεζ\int_{\mathbb{R}^{d}}\int_{\bar{\rho}_{0}}^{\rho_{0}}\frac{\rho_{0}-z}{z}\,dzdx+\int_{\mathbb{R}^{d}}\left(\int_{\mathbb{R}^{d}}f_{0}\log f_{0}\,dv-\rho_{0}\log\rho_{0}\right)dx+\int_{\mathbb{R}^{d}}|\nabla(\Phi_{0}-\bar{\Phi}_{0})|^{2}\,dx\leq C\varepsilon^{\zeta}

for some C>0C>0 independent of ε>0\varepsilon>0, where ζ\zeta is given by

ζ:={min{1,δd/2}for V(x)=|x|2/2min{1,δ}for V0.\zeta:=\left\{\begin{array}[]{ll}\min\{1,\delta-d/2\}&\textrm{for $V(x)=|x|^{2}/2$}\\[5.69054pt] \min\{1,\delta\}&\textrm{for $V\equiv 0$}\end{array}\right..

Then we have

dρ¯ρρzz𝑑z𝑑x+d|(ΦΦ¯)|2𝑑x+0tdρ|uu¯|2𝑑x𝑑sCεζ\int_{\mathbb{R}^{d}}\int_{\bar{\rho}}^{\rho}\frac{\rho-z}{z}\,dz\,dx+\int_{\mathbb{R}^{d}}|\nabla(\Phi-\bar{\Phi})|^{2}\,dx+\int_{0}^{t}\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dxds\leq C\varepsilon^{\zeta}

for t<Tt<T, where C>0C>0 is independent of ε\varepsilon. In particular, if ζ>0\zeta>0, then the following strong convergences hold:

ρρ¯inL(0,T;L1H1(d))\rho\to\bar{\rho}\quad\mbox{in}\quad L^{\infty}(0,T;L^{1}\cap H^{-1}(\mathbb{R}^{d}))

and

ρuρ¯u¯inL2(0,T;L1(d))\rho u\to\bar{\rho}\bar{u}\quad\mbox{in}\quad L^{2}(0,T;L^{1}(\mathbb{R}^{d}))

as ε0\varepsilon\to 0.

Remark 1.3.

One may extend our result to the case with confinement potentials VV satisfying

eVL1(d),|2V(x)|c1,and|V(x)|2c2(1+V(x))for xd,e^{-V}\in L^{1}(\mathbb{R}^{d}),\quad|\nabla^{2}V(x)|\leq c_{1},\quad\mbox{and}\quad|\nabla V(x)|^{2}\leq c_{2}(1+V(x))\quad\mbox{for }x\in\mathbb{R}^{d},

for some ci>0c_{i}>0, i=1,2i=1,2. It is clear that the quadratic confinement potential V(x)=|x|2/2V(x)=|x|^{2}/2 satisfies the above inequality with c1=1c_{1}=1 and c2=2c_{2}=2. Our strategy can be easily extended to the case that Φ\Phi is given as K~ρ\nabla\tilde{K}\star\rho, where the interaction potential K~\tilde{K} satisfying K~L(d)\nabla\tilde{K}\in L^{\infty}(\mathbb{R}^{d}). See Remark 3.2 for details.

1.6. Organization of paper

The rest of this paper is organized as follows. In Section 2, we provide a free energy estimate and introduce our main functional, relative entropy functional. Section 3 is devoted to prove our main results; Theorem 1.1 and Corollary 1.1. Finally, in Section 4, we present some bound estimates for the solution ρ¯\bar{\rho} to (1.3) which are required for our quantitative error estimates.

2. Preliminaries

2.1. Free energy estimate

Let us first introduce free energy ε\mathcal{F}_{\varepsilon} for the system (1.2) and its associated dissipation 𝒟ε\mathcal{D}_{\varepsilon}:

ε(f):=1εd×dflogfdxdv+d×df|v|22𝑑x𝑑v+1εdρV𝑑x+12εdΦρ𝑑x\mathcal{F}_{\varepsilon}(f):=\frac{1}{\varepsilon}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f\log f\,dxdv+\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f\frac{|v|^{2}}{2}\,dxdv+\frac{1}{\varepsilon}\int_{\mathbb{R}^{d}}\rho V\,dx+\frac{1}{2\varepsilon}\int_{\mathbb{R}^{d}}\Phi\rho\,dx

and

𝒟ε(f):=d×d1f|1εvf(uv)f|2𝑑x𝑑v.\mathcal{D}_{\varepsilon}(f):=\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\frac{1}{f}\left|\frac{1}{\varepsilon}\nabla_{v}f-(u-v)f\right|^{2}\,dxdv.
Lemma 2.1.

Let T>0T>0 and ff be a solution of (1.2) with sufficient integrability on the time interval [0,T][0,T]. Then we have

ε(f)+0t(1ε2+δ𝒟ε(f)+1εd×df|v|2𝑑x𝑑v)𝑑sε(f0)+dε2t.\mathcal{F}_{\varepsilon}(f)+\int_{0}^{t}\left(\frac{1}{\varepsilon^{2+\delta}}\mathcal{D}_{\varepsilon}(f)+\frac{1}{\varepsilon}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|v|^{2}dxdv\right)ds\leq\mathcal{F}_{\varepsilon}(f_{0})+d\varepsilon^{-2}t.

for t[0,T]t\in[0,T]. Furthermore, we obtain

ε(f)+0t(1ε2+δ𝒟ε(f)+1εdρ|u|2𝑑x)𝑑sε(f0)+εδ20td×df|v|2𝑑x𝑑v𝑑s.\mathcal{F}_{\varepsilon}(f)+\int_{0}^{t}\left(\frac{1}{\varepsilon^{2+\delta}}\mathcal{D}_{\varepsilon}(f)+\frac{1}{\varepsilon}\int_{\mathbb{R}^{d}}\rho|u|^{2}\,dx\right)ds\leq\mathcal{F}_{\varepsilon}(f_{0})+\frac{\varepsilon^{\delta}}{2}\int_{0}^{t}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|v|^{2}\,dxdvds.

Here ρ\rho and uu are defined as in (1.4).

Proof.

Using equation (1.2), we can directly compute

ddtd×dflogfdxdv=d×d(1+logf)tfdxdv=1εd×d1fvf(v+(V+Φ))f𝑑x𝑑v1ε2+δd×d1fvf((vu)f+1εvf)𝑑x𝑑v=dε1ε2+δd×d1fvf((vu)f+1εvf)𝑑x𝑑v.\displaystyle\begin{aligned} \frac{d}{dt}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f\log f\,dxdv&=\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}(1+\log f)\,\partial_{t}f\,dxdv\cr&=-\frac{1}{\varepsilon}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\frac{1}{f}\nabla_{v}f\cdot\Big{(}v+(\nabla V+\nabla\Phi)\Big{)}\,f\,dxdv\cr&\quad\,-\frac{1}{\varepsilon^{2+\delta}}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\frac{1}{f}\nabla_{v}f\cdot\Big{(}(v-u)f+\frac{1}{\varepsilon}\nabla_{v}f\Big{)}\,dxdv\cr&=\frac{d}{\varepsilon}-\frac{1}{\varepsilon^{2+\delta}}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\frac{1}{f}\nabla_{v}f\cdot\Big{(}(v-u)f+\frac{1}{\varepsilon}\nabla_{v}f\Big{)}\,dxdv.\end{aligned} (2.1)

Similarly, we deduce

ddtd×df|v|22𝑑x𝑑v\displaystyle\frac{d}{dt}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f\frac{|v|^{2}}{2}dxdv =d×dtf|v|22dxdv\displaystyle=\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\partial_{t}f\frac{|v|^{2}}{2}dxdv
=1εd×df|v|2𝑑x𝑑vddt(1εdρV𝑑x+12εdΦρ𝑑x)\displaystyle=-\frac{1}{\varepsilon}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|v|^{2}dxdv-\frac{d}{dt}\left(\frac{1}{\varepsilon}\int_{\mathbb{R}^{d}}\rho Vdx+\frac{1}{2\varepsilon}\int_{\mathbb{R}^{d}}\Phi\rho\,dx\right)
1ε2+δd×df|vu|2𝑑x𝑑v1ε3+δd×d(vu)vfdxdv,\displaystyle\quad-\frac{1}{\varepsilon^{2+\delta}}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|v-u|^{2}dxdv-\frac{1}{\varepsilon^{3+\delta}}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}(v-u)\cdot\nabla_{v}f\,dxdv,

which subsequently implies

ddt(d×df|v|22𝑑x𝑑v+1εdρV𝑑x+12εdΦρ𝑑x)=1εd×df|v|2𝑑x𝑑v1ε2+δd×df|vu|2𝑑x𝑑v1ε3+δd×d(vu)vfdxdv.\displaystyle\begin{aligned} &\frac{d}{dt}\left(\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f\frac{|v|^{2}}{2}\,dxdv+\frac{1}{\varepsilon}\int_{\mathbb{R}^{d}}\rho V\,dx+\frac{1}{2\varepsilon}\int_{\mathbb{R}^{d}}\Phi\rho\,dx\right)\cr&\quad=-\frac{1}{\varepsilon}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|v|^{2}\,dxdv-\frac{1}{\varepsilon^{2+\delta}}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|v-u|^{2}\,dxdv\cr&\qquad-\frac{1}{\varepsilon^{3+\delta}}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}(v-u)\cdot\nabla_{v}f\,dxdv.\end{aligned} (2.2)

Dividing (2.1) by ε\varepsilon and adding the resulting equation to (2.2), we get

ddtε(f)+1εd×df|v|2𝑑x𝑑v+1ε2+δd×d1f|1εvf(uv)f|2𝑑x𝑑v=dε2.\frac{d}{dt}\mathcal{F}_{\varepsilon}(f)+\frac{1}{\varepsilon}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|v|^{2}dxdv+\frac{1}{\varepsilon^{2+\delta}}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\frac{1}{f}\left|\frac{1}{\varepsilon}\nabla_{v}f-(u-v)f\right|^{2}dxdv=\frac{d}{\varepsilon^{2}}. (2.3)

Integrating the above with respect to time gives the first assertion.

One can further estimate

dε2=1εd×dv((uv)f1εvf)𝑑x𝑑v1εd×dv(uv)f𝑑x𝑑v1ε(d×df|v|2𝑑x𝑑v)1/2(d×d1f|1εvf(uv)f|2𝑑x𝑑v)1/21εdρ|u|2𝑑x+1εd×df|v|2𝑑x𝑑v12ε2+δ𝒟ε(f)+εδ2d×df|v|2𝑑x𝑑v1εdρ|u|2𝑑x+1εd×df|v|2𝑑x𝑑v.\displaystyle\begin{aligned} \frac{d}{\varepsilon^{2}}&=\frac{1}{\varepsilon}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}v\cdot\left((u-v)f-\frac{1}{\varepsilon}\nabla_{v}f\right)dxdv-\frac{1}{\varepsilon}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}v\cdot(u-v)f\,dxdv\cr&\leq\frac{1}{\varepsilon}\left(\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|v|^{2}\,dxdv\right)^{1/2}\left(\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\frac{1}{f}\left|\frac{1}{\varepsilon}\nabla_{v}f-(u-v)f\right|^{2}dxdv\right)^{1/2}\cr&\quad-\frac{1}{\varepsilon}\int_{\mathbb{R}^{d}}\rho|u|^{2}\,dx+\frac{1}{\varepsilon}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|v|^{2}\,dxdv\cr&\leq\frac{1}{2\varepsilon^{2+\delta}}\mathcal{D}_{\varepsilon}(f)+\frac{\varepsilon^{\delta}}{2}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|v|^{2}\,dxdv-\frac{1}{\varepsilon}\int_{\mathbb{R}^{d}}\rho|u|^{2}dx+\frac{1}{\varepsilon}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|v|^{2}\,dxdv.\end{aligned} (2.4)

Substituting (2.4) into (2.3), we have

ddtε(f)+12ε2+δ𝒟ε(f)+1εdρ|u|2𝑑xεδ2d×df|v|2𝑑x𝑑v\displaystyle\frac{d}{dt}\mathcal{F}_{\varepsilon}(f)+\frac{1}{2\varepsilon^{2+\delta}}\mathcal{D}_{\varepsilon}(f)+\frac{1}{\varepsilon}\int_{\mathbb{R}^{d}}\rho|u|^{2}\,dx\leq\frac{\varepsilon^{\delta}}{2}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|v|^{2}\,dxdv

from which, we obtain the second assertion. ∎

Remark 2.1.

In [4], it is showed that the weak solutions ff of (1.2) in the sense of Definition 1.1 satisfies the entropy inequality appeared in Lemma 2.1.

2.2. Relative entropy

Note that the equation (1.3) can be also rewritten as

tρ¯+(ρ¯u¯)=0,t(ρ¯u¯)+(ρ¯u¯u¯)+1ερ¯=1ερ¯u¯1ερ¯(V+Φ¯)+e¯,ΔΦ¯=ρ¯,\displaystyle\begin{aligned} &\partial_{t}\bar{\rho}+\nabla\cdot(\bar{\rho}\bar{u})=0,\cr&\partial_{t}(\bar{\rho}\bar{u})+\nabla\cdot(\bar{\rho}\bar{u}\otimes\bar{u})+\frac{1}{\varepsilon}\nabla\bar{\rho}=-\frac{1}{\varepsilon}\bar{\rho}\bar{u}-\frac{1}{\varepsilon}\bar{\rho}(\nabla V+\nabla\bar{\Phi})+\bar{e},\cr&-\Delta\bar{\Phi}=\bar{\rho},\end{aligned} (2.5)

where e¯=ρ¯(tu¯+u¯u¯)\bar{e}=\bar{\rho}(\partial_{t}\bar{u}+\bar{u}\cdot\nabla\bar{u}). Let us rewrite the system (2.5) as a conservative form:

tU¯+Aε(U¯)=Fε(U¯),\partial_{t}\bar{U}+\nabla\cdot A_{\varepsilon}(\bar{U})=F_{\varepsilon}(\bar{U}),

where

m¯=ρ¯u¯,U¯:=(ρ¯m¯),Aε(U¯):=(m¯0(m¯m¯)/ρ¯ρ¯/ε),\bar{m}=\bar{\rho}\bar{u},\quad\bar{U}:=\begin{pmatrix}\bar{\rho}\\ \bar{m}\end{pmatrix},\quad A_{\varepsilon}(\bar{U}):=\begin{pmatrix}\bar{m}&0\\ (\bar{m}\otimes\bar{m})/\bar{\rho}&\bar{\rho}/\varepsilon\end{pmatrix},

and

Fε(U¯):=(01ερ¯u¯1ερ¯(V+Φ¯)+e¯).F_{\varepsilon}(\bar{U}):=\begin{pmatrix}0\\ \displaystyle-\frac{1}{\varepsilon}\bar{\rho}\bar{u}-\frac{1}{\varepsilon}\bar{\rho}(\nabla V+\nabla\bar{\Phi})+\bar{e}\end{pmatrix}.

Then the above system has the following macroscopic entropy form:

ε(U¯):=|m¯|22ρ¯+1ερ¯logρ¯.\mathcal{E}_{\varepsilon}(\bar{U}):=\frac{|\bar{m}|^{2}}{2\bar{\rho}}+\frac{1}{\varepsilon}\bar{\rho}\log\bar{\rho}.

We now define the relative entropy functional ε\mathcal{H}_{\varepsilon} as follows.

ε(U|U¯):=ε(U)ε(U¯)Dε(U¯)(UU¯)withU:=(ρm),m=ρu,\mathcal{H}_{\varepsilon}(U|\bar{U}):=\mathcal{E}_{\varepsilon}(U)-\mathcal{E}_{\varepsilon}(\bar{U})-D\mathcal{E}_{\varepsilon}(\bar{U})(U-\bar{U})\quad\mbox{with}\quad U:=\begin{pmatrix}\rho\\ m\\ \end{pmatrix},\quad m=\rho u,

where Dε(U¯)D\mathcal{E}_{\varepsilon}(\bar{U}) denotes the derivation of ε\mathcal{E}_{\varepsilon} with respect to ρ¯,m¯\bar{\rho},\bar{m}, and we find

Dε(U¯)(UU¯)\displaystyle-D\mathcal{E}_{\varepsilon}(\bar{U})(U-\bar{U}) =(|m¯|22ρ¯2+1ε(logρ¯+1)m¯ρ¯)(ρρ¯mm¯)\displaystyle=-\begin{pmatrix}\displaystyle-\frac{|\bar{m}|^{2}}{2\bar{\rho}^{2}}+\frac{1}{\varepsilon}(\log\bar{\rho}+1)\\[8.53581pt] \displaystyle\frac{\bar{m}}{\bar{\rho}}\end{pmatrix}\begin{pmatrix}\rho-\bar{\rho}\\ m-\bar{m}\end{pmatrix}
=ρ|u¯|22ρ¯|u¯|22+1ε(ρ¯ρ)(logρ¯+1)+ρ¯|u¯|2ρu¯u.\displaystyle=\frac{\rho|\bar{u}|^{2}}{2}-\frac{\bar{\rho}|\bar{u}|^{2}}{2}+\frac{1}{\varepsilon}(\bar{\rho}-\rho)(\log\bar{\rho}+1)+\bar{\rho}|\bar{u}|^{2}-\rho\bar{u}\cdot u.

This yields

ε(U|U¯)=ρ2|uu¯|2+1εp(ρ|ρ¯),\mathcal{H}_{\varepsilon}(U|\bar{U})=\frac{\rho}{2}|u-\bar{u}|^{2}+\frac{1}{\varepsilon}p(\rho|\bar{\rho}),

where pp represents the relative entropy which is defined by

p(ρ|ρ¯):=ρ¯ρρzz𝑑z.p(\rho|\bar{\rho}):=\int_{\bar{\rho}}^{\rho}\frac{\rho-z}{z}\,dz.

3. Proofs of Theorem 1.1 & Corollary 1.1

3.1. Relative entropy estimate

Proposition 3.1.

Let T>0T>0. Let ff be the solution to the equation (1.2) in the sense of Definition 1.1 and ρ¯\bar{\rho} be the strong solution to the equation (1.3) in the sense of Definition 1.2 on the time interval [0,T][0,T]. Then for ε(0,1)\varepsilon\in(0,1) small enough we have

12dρ|uu¯|2𝑑x+1εdp(ρ|ρ¯)𝑑x+12ε0tdρ|uu¯|2𝑑x𝑑s+12εd|(ΦΦ¯)|2𝑑xCdρ0|u0u¯0|2𝑑x+Cεdp(ρ0|ρ¯0)𝑑x+Cd(𝒦ε(f0)ε(U0))𝑑x+Cεδ0td×df|v|2𝑑x𝑑v𝑑s+Cε+Cεd|(Φ0Φ¯0)|2𝑑x,\displaystyle\begin{aligned} &\frac{1}{2}\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dx+\frac{1}{\varepsilon}\int_{\mathbb{R}^{d}}p(\rho|\bar{\rho})\,dx+\frac{1}{2\varepsilon}\int_{0}^{t}\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dxds+\frac{1}{2\varepsilon}\int_{\mathbb{R}^{d}}|\nabla(\Phi-\bar{\Phi})|^{2}\,dx\cr&\quad\leq C\int_{\mathbb{R}^{d}}\rho_{0}|u_{0}-\bar{u}_{0}|^{2}\,dx+\frac{C}{\varepsilon}\int_{\mathbb{R}^{d}}p(\rho_{0}|\bar{\rho}_{0})\,dx+C\int_{\mathbb{R}^{d}}\left(\mathcal{K}_{\varepsilon}(f_{0})-\mathcal{E}_{\varepsilon}(U_{0})\right)dx\cr&\qquad\,\,+C\varepsilon^{\delta}\int_{0}^{t}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|v|^{2}\,dxdvds+C\varepsilon+\frac{C}{\varepsilon}\int_{\mathbb{R}^{d}}|\nabla(\Phi_{0}-\bar{\Phi}_{0})|^{2}\,dx,\end{aligned} (3.1)

where C>0C>0 is independent of ε\varepsilon.

Proof.

Straightforward computations yield

dε(U|U¯)𝑑x=dε(U0|U¯0)𝑑x+dε(U)𝑑xdε(U0)𝑑x0td(Dε(U¯)):Aε(U|U¯)dxds0tdD2ε(U¯)Fε(U¯)(UU¯)+Dε(U¯)Fε(U)dxds=:i=15Ii,\displaystyle\begin{aligned} \int_{\mathbb{R}^{d}}\mathcal{H}_{\varepsilon}(U|\bar{U})\,dx&=\int_{\mathbb{R}^{d}}\mathcal{H}_{\varepsilon}(U_{0}|\bar{U}_{0})\,dx+\int_{\mathbb{R}^{d}}\mathcal{E}_{\varepsilon}(U)\,dx-\int_{\mathbb{R}^{d}}\mathcal{E}_{\varepsilon}(U_{0})\,dx\cr&\quad-\int_{0}^{t}\int_{\mathbb{R}^{d}}\nabla(D\mathcal{E}_{\varepsilon}(\bar{U})):A_{\varepsilon}(U|\bar{U})\,dxds\cr&\quad-\int_{0}^{t}\int_{\mathbb{R}^{d}}D^{2}\mathcal{E}_{\varepsilon}(\bar{U})F_{\varepsilon}(\bar{U})(U-\bar{U})+D\mathcal{E}_{\varepsilon}(\bar{U})F_{\varepsilon}(U)\,dxds\cr&=:\sum_{i=1}^{5}I_{i},\end{aligned}

where we easily estimate I4I_{4} as

I4u¯L0tdρ|uu¯|2𝑑x𝑑s.I_{4}\leq\|\nabla\bar{u}\|_{L^{\infty}}\int_{0}^{t}\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dxds.

Set 𝒦ε\mathcal{K}_{\varepsilon} the mesoscopic entropy:

𝒦ε(f):=df|v|22𝑑v+1εdflogfdv.\mathcal{K}_{\varepsilon}(f):=\int_{\mathbb{R}^{d}}f\frac{|v|^{2}}{2}\,dv+\frac{1}{\varepsilon}\int_{\mathbb{R}^{d}}f\log f\,dv.

By the classical minimization principle, see [26], we find

dε(U)𝑑xd𝒦ε(f)𝑑x,\int_{\mathbb{R}^{d}}\mathcal{E}_{\varepsilon}(U)\,dx\leq\int_{\mathbb{R}^{d}}\mathcal{K}_{\varepsilon}(f)\,dx,

and this gives

I2+I3\displaystyle I_{2}+I_{3} =dε(U)𝑑xd𝒦ε(f)𝑑x+d𝒦ε(f)𝑑xd𝒦ε(f0)𝑑x\displaystyle=\int_{\mathbb{R}^{d}}\mathcal{E}_{\varepsilon}(U)\,dx-\int_{\mathbb{R}^{d}}\mathcal{K}_{\varepsilon}(f)\,dx+\int_{\mathbb{R}^{d}}\mathcal{K}_{\varepsilon}(f)\,dx-\int_{\mathbb{R}^{d}}\mathcal{K}_{\varepsilon}(f_{0})\,dx
+d𝒦ε(f0)𝑑xdε(U0)𝑑x\displaystyle\quad+\int_{\mathbb{R}^{d}}\mathcal{K}_{\varepsilon}(f_{0})\,dx-\int_{\mathbb{R}^{d}}\mathcal{E}_{\varepsilon}(U_{0})\,dx
d𝒦ε(f)𝑑xd𝒦ε(f0)𝑑x+d𝒦ε(f0)𝑑xdε(U0)𝑑x.\displaystyle\leq\int_{\mathbb{R}^{d}}\mathcal{K}_{\varepsilon}(f)\,dx-\int_{\mathbb{R}^{d}}\mathcal{K}_{\varepsilon}(f_{0})\,dx+\int_{\mathbb{R}^{d}}\mathcal{K}_{\varepsilon}(f_{0})\,dx-\int_{\mathbb{R}^{d}}\mathcal{E}_{\varepsilon}(U_{0})\,dx.

On the other hand, since

Fε(U)=(01ερu1ερ(V+Φ)+e),F_{\varepsilon}(U)=\begin{pmatrix}0\\ \displaystyle-\frac{1}{\varepsilon}\rho u-\frac{1}{\varepsilon}\rho(\nabla V+\nabla\Phi)+e\end{pmatrix},

where ee is appeared in (1.6), we get

dD2ε(U¯)Fε(U¯)(UU¯)+Dε(U¯)Fε(U)dx=1εdρ|uu¯|2𝑑x+1εdρ|u|2𝑑x+1εdρuVdx+1εdρuΦdx+1εdρ(uu¯)(Φ¯Φ)dxdρ(uu¯)e¯ρ¯𝑑xdu¯e𝑑x=1εdρ|uu¯|2𝑑x+1εdρ|u|2𝑑x+1εddtdρV𝑑x+12εddtdρΦ𝑑x+1εdρ(uu¯)(Φ¯Φ)dxdρ(uu¯)e¯ρ¯𝑑xdu¯e𝑑x.\displaystyle\begin{aligned} &-\int_{\mathbb{R}^{d}}D^{2}\mathcal{E}_{\varepsilon}(\bar{U})F_{\varepsilon}(\bar{U})(U-\bar{U})+D\mathcal{E}_{\varepsilon}(\bar{U})F_{\varepsilon}(U)\,dx\cr&\quad=-\frac{1}{\varepsilon}\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dx+\frac{1}{\varepsilon}\int_{\mathbb{R}^{d}}\rho|u|^{2}\,dx+\frac{1}{\varepsilon}\int_{\mathbb{R}^{d}}\rho u\cdot\nabla V\,dx+\frac{1}{\varepsilon}\int_{\mathbb{R}^{d}}\rho u\cdot\nabla\Phi\,dx\cr&\qquad+\frac{1}{\varepsilon}\int_{\mathbb{R}^{d}}\rho(u-\bar{u})\cdot\nabla(\bar{\Phi}-\Phi)\,dx-\int_{\mathbb{R}^{d}}\rho(u-\bar{u})\cdot\frac{\bar{e}}{\bar{\rho}}\,dx-\int_{\mathbb{R}^{d}}\bar{u}\cdot e\,dx\cr&\quad=-\frac{1}{\varepsilon}\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dx+\frac{1}{\varepsilon}\int_{\mathbb{R}^{d}}\rho|u|^{2}\,dx+\frac{1}{\varepsilon}\frac{d}{dt}\int_{\mathbb{R}^{d}}\rho V\,dx+\frac{1}{2\varepsilon}\frac{d}{dt}\int_{\mathbb{R}^{d}}\rho\Phi\,dx\cr&\qquad+\frac{1}{\varepsilon}\int_{\mathbb{R}^{d}}\rho(u-\bar{u})\cdot\nabla(\bar{\Phi}-\Phi)\,dx-\int_{\mathbb{R}^{d}}\rho(u-\bar{u})\cdot\frac{\bar{e}}{\bar{\rho}}\,dx-\int_{\mathbb{R}^{d}}\bar{u}\cdot e\,dx.\end{aligned} (3.2)

We next recall from [13, Lemma 5.1], see also [7, 27, 28] that

12ddtd|(ΦΦ¯)|2𝑑x=d(ΦΦ¯)(ρuρ¯u¯)𝑑x.\frac{1}{2}\frac{d}{dt}\int_{\mathbb{R}^{d}}\left|\nabla(\Phi-\bar{\Phi})\right|^{2}dx=\int_{\mathbb{R}^{d}}\nabla(\Phi-\bar{\Phi})\cdot(\rho u-\bar{\rho}\bar{u})\,dx.

Thus, we obtain

1εdρ(uu¯)(Φ¯Φ)dx=1εd(ΦΦ¯)(ρuρ¯u¯)𝑑x+1εd(ΦΦ¯)u¯(ρρ¯)𝑑x=12εddtd|(ΦΦ¯)|2𝑑x+1εd(ΦΦ¯)u¯(ρρ¯)𝑑x.\displaystyle\begin{aligned} &\frac{1}{\varepsilon}\int_{\mathbb{R}^{d}}\rho(u-\bar{u})\cdot\nabla(\bar{\Phi}-\Phi)\,dx\cr&\quad=-\frac{1}{\varepsilon}\int_{\mathbb{R}^{d}}\nabla(\Phi-\bar{\Phi})\cdot(\rho u-\bar{\rho}\bar{u})\,dx+\frac{1}{\varepsilon}\int_{\mathbb{R}^{d}}\nabla(\Phi-\bar{\Phi})\cdot\bar{u}(\rho-\bar{\rho})\,dx\cr&\quad=-\frac{1}{2\varepsilon}\frac{d}{dt}\int_{\mathbb{R}^{d}}\left|\nabla(\Phi-\bar{\Phi})\right|^{2}dx+\frac{1}{\varepsilon}\int_{\mathbb{R}^{d}}\nabla(\Phi-\bar{\Phi})\cdot\bar{u}(\rho-\bar{\rho})\,dx.\end{aligned} (3.3)

For the second term on the right-hand-side of (3.3), we compute that

1εd(ΦΦ¯)u¯(ρρ¯)𝑑x=1εd(ΦΦ¯)u¯Δ(ΦΦ¯)𝑑x=12εd|(ΦΦ¯)|2u¯𝑑x+1εd(ΦΦ¯)(ΦΦ¯):u¯dx32εu¯Ld|(ΦΦ¯)|2𝑑x.\displaystyle\begin{aligned} &\frac{1}{\varepsilon}\int_{\mathbb{R}^{d}}\nabla(\Phi-\bar{\Phi})\cdot\bar{u}(\rho-\bar{\rho})\,dx\cr&\quad=-\frac{1}{\varepsilon}\int_{\mathbb{R}^{d}}\nabla(\Phi-\bar{\Phi})\cdot\bar{u}\Delta(\Phi-\bar{\Phi})\,dx\cr&\quad=-\frac{1}{2\varepsilon}\int_{\mathbb{R}^{d}}\left|\nabla(\Phi-\bar{\Phi})\right|^{2}\nabla\cdot\bar{u}\,dx+\frac{1}{\varepsilon}\int_{\mathbb{R}^{d}}\nabla(\Phi-\bar{\Phi})\otimes\nabla(\Phi-\bar{\Phi}):\nabla\bar{u}\,dx\cr&\quad\leq\frac{3}{2\varepsilon}\|\nabla\bar{u}\|_{L^{\infty}}\int_{\mathbb{R}^{d}}\left|\nabla(\Phi-\bar{\Phi})\right|^{2}\,dx.\end{aligned} (3.4)

Substituting (3.4) into (3.3), we see that

1εdρ(uu¯)(Φ¯Φ)dx12εddtd|(ΦΦ¯)|2𝑑x+32εu¯Ld|(ΦΦ¯)|2𝑑x.\displaystyle\frac{1}{\varepsilon}\int_{\mathbb{R}^{d}}\rho(u-\bar{u})\cdot\nabla(\bar{\Phi}-\Phi)\,dx\leq-\frac{1}{2\varepsilon}\frac{d}{dt}\int_{\mathbb{R}^{d}}\left|\nabla(\Phi-\bar{\Phi})\right|^{2}dx+\frac{3}{2\varepsilon}\|\nabla\bar{u}\|_{L^{\infty}}\int_{\mathbb{R}^{d}}\left|\nabla(\Phi-\bar{\Phi})\right|^{2}dx.

For the estimate of the sixth term on the right hand side of (3.2), we give the following lemma. As mentioned in Introduction, we cannot use the assumption u¯L(d×(0,T))\bar{u}\in L^{\infty}(\mathbb{R}^{d}\times(0,T)) when we consider V=|x|2/2V=|x|^{2}/2, and thus we estimate it differently from [9, 28]. For the sake of the reader, we provide the details of proof at the end of this subsection.

Lemma 3.1.

There exists a constant C>0C>0 depending only on the regularity estimates ρ¯L(0,T;W1,1W1,)\|\bar{\rho}\|_{L^{\infty}(0,T;W^{1,1}\cap W^{1,\infty})} and logρ¯L(0,T;W2,)\|\nabla\log\bar{\rho}\|_{L^{\infty}(0,T;W^{2,\infty})} such that

dρ(uu¯)e¯ρ¯𝑑xdρ|uu¯|2𝑑x+C(dρ|uu¯|2𝑑x)1/2.\int_{\mathbb{R}^{d}}\rho(u-\bar{u})\cdot\frac{\bar{e}}{\bar{\rho}}\,dx\leq\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dx+C\left(\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dx\right)^{1/2}.

In particular, for ε>0\varepsilon>0 small enough we have

dρ(uu¯)e¯ρ¯𝑑x12εdρ|uu¯|2𝑑x+Cε\int_{\mathbb{R}^{d}}\rho(u-\bar{u})\cdot\frac{\bar{e}}{\bar{\rho}}\,dx\leq\frac{1}{2\varepsilon}\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dx+C\varepsilon

for some C>0C>0 independent of ε>0\varepsilon>0.

We next estimate

|du¯e𝑑x|\displaystyle\left|\int_{\mathbb{R}^{d}}\bar{u}\cdot e\,dx\right| =|du¯((d(uuvv+1ε𝕀)f𝑑v))𝑑x|\displaystyle=\left|\int_{\mathbb{R}^{d}}\bar{u}\cdot\left(\nabla\cdot\left(\int_{\mathbb{R}^{d}}\left(u\otimes u-v\otimes v+\frac{1}{\varepsilon}\mathbb{I}\right)f\,dv\right)\right)dx\right|
u¯Ld|d(uuvv+1ε𝕀)f𝑑v|𝑑x\displaystyle\leq\|\nabla\bar{u}\|_{L^{\infty}}\int_{\mathbb{R}^{d}}\left|\int_{\mathbb{R}^{d}}\left(u\otimes u-v\otimes v+\frac{1}{\varepsilon}\mathbb{I}\right)f\,dv\right|dx
Cu¯L(d×df|v|2𝑑x𝑑v)1/2(d×d1f|1εvf(uv)f|2𝑑x𝑑v)1/2\displaystyle\leq C\|\nabla\bar{u}\|_{L^{\infty}}\left(\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|v|^{2}\,dxdv\right)^{1/2}\left(\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\frac{1}{f}\left|\frac{1}{\varepsilon}\nabla_{v}f-(u-v)f\right|^{2}dxdv\right)^{1/2}
Cu¯Lε2+δd×df|v|2𝑑x𝑑v+12ε2+δ𝒟ε(f)\displaystyle\leq C\|\nabla\bar{u}\|_{L^{\infty}}\varepsilon^{2+\delta}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|v|^{2}\,dxdv+\frac{1}{2\varepsilon^{2+\delta}}\mathcal{D}_{\varepsilon}(f)

due to [26, Proof of Lemma 4.4]. This yields

I512ε0tdρ|uu¯|2𝑑x𝑑s12εd|(ΦΦ¯)|2𝑑x+12εd|(Φ0Φ¯0)|2𝑑x+CTε+Cε0td|(ΦΦ¯)|2𝑑x𝑑s+1ε0tdρ|u|2𝑑x𝑑s+Cε2+δ0td×df|v|2𝑑x𝑑v𝑑s+12ε2+δ0t𝒟ε(f)𝑑s+1ε(dρV𝑑xdρ0V𝑑x)+12ε(dρΦ𝑑xdρ0Φ0𝑑x).\displaystyle\begin{aligned} I_{5}&\leq-\frac{1}{2\varepsilon}\int_{0}^{t}\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dxds-\frac{1}{2\varepsilon}\int_{\mathbb{R}^{d}}\left|\nabla(\Phi-\bar{\Phi})\right|^{2}dx+\frac{1}{2\varepsilon}\int_{\mathbb{R}^{d}}\left|\nabla(\Phi_{0}-\bar{\Phi}_{0})\right|^{2}dx\cr&\quad+CT\varepsilon+\frac{C}{\varepsilon}\int_{0}^{t}\int_{\mathbb{R}^{d}}\left|\nabla(\Phi-\bar{\Phi})\right|^{2}dxds+\frac{1}{\varepsilon}\int_{0}^{t}\int_{\mathbb{R}^{d}}\rho|u|^{2}\,dxds\\ &\quad+C\varepsilon^{2+\delta}\int_{0}^{t}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|v|^{2}\,dxdvds+\frac{1}{2\varepsilon^{2+\delta}}\int_{0}^{t}\mathcal{D}_{\varepsilon}(f)\,ds\cr&\quad+\frac{1}{\varepsilon}\left(\int_{\mathbb{R}^{d}}\rho V\,dx-\int_{\mathbb{R}^{d}}\rho_{0}V\,dx\right)+\frac{1}{2\varepsilon}\left(\int_{\mathbb{R}^{d}}\rho\Phi\,dx-\int_{\mathbb{R}^{d}}\rho_{0}\Phi_{0}\,dx\right).\end{aligned}

Thus we obtain

I2+I3+I5\displaystyle I_{2}+I_{3}+I_{5} 12ε0tdρ|uu¯|2𝑑x𝑑s12εd|(ΦΦ¯)|2𝑑x+12εd|(Φ0Φ¯0)|2𝑑x\displaystyle\leq-\frac{1}{2\varepsilon}\int_{0}^{t}\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dxds-\frac{1}{2\varepsilon}\int_{\mathbb{R}^{d}}\left|\nabla(\Phi-\bar{\Phi})\right|^{2}dx+\frac{1}{2\varepsilon}\int_{\mathbb{R}^{d}}\left|\nabla(\Phi_{0}-\bar{\Phi}_{0})\right|^{2}dx
+Cε0td|(ΦΦ¯)|2𝑑x𝑑s+d𝒦ε(f0)𝑑xdε(U0)𝑑x+CTε\displaystyle\quad+\frac{C}{\varepsilon}\int_{0}^{t}\int_{\mathbb{R}^{d}}\left|\nabla(\Phi-\bar{\Phi})\right|^{2}dxds+\int_{\mathbb{R}^{d}}\mathcal{K}_{\varepsilon}(f_{0})\,dx-\int_{\mathbb{R}^{d}}\mathcal{E}_{\varepsilon}(U_{0})\,dx+CT\varepsilon
+Cεδ0td×df|v|2𝑑x𝑑v𝑑s.\displaystyle\quad+C\varepsilon^{\delta}\int_{0}^{t}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|v|^{2}\,dxdvds.

Combining all of the above estimates provides

12dρ|uu¯|2𝑑x+1εdp(ρ|ρ¯)𝑑x+12ε0tdρ|uu¯|2𝑑x𝑑s+12εd|(ΦΦ¯)|2𝑑x12dρ0|u0u¯0|2𝑑x+1εdp(ρ0|ρ¯0)𝑑x+d(𝒦ε(f0)ε(U0))𝑑x+Cε+Cεd|(Φ0Φ¯0)|2𝑑x+Cε0td|(ΦΦ¯)|2𝑑x𝑑s+Cεδ0td×df|v|2𝑑x𝑑v𝑑s\displaystyle\begin{aligned} &\frac{1}{2}\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dx+\frac{1}{\varepsilon}\int_{\mathbb{R}^{d}}p(\rho|\bar{\rho})\,dx+\frac{1}{2\varepsilon}\int_{0}^{t}\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dxds+\frac{1}{2\varepsilon}\int_{\mathbb{R}^{d}}|\nabla(\Phi-\bar{\Phi})|^{2}\,dx\cr&\qquad\leq\frac{1}{2}\int_{\mathbb{R}^{d}}\rho_{0}|u_{0}-\bar{u}_{0}|^{2}\,dx+\frac{1}{\varepsilon}\int_{\mathbb{R}^{d}}p(\rho_{0}|\bar{\rho}_{0})\,dx+\int_{\mathbb{R}^{d}}\left(\mathcal{K}_{\varepsilon}(f_{0})-\mathcal{E}_{\varepsilon}(U_{0})\right)dx\cr&\quad\qquad+C\varepsilon+\frac{C}{\varepsilon}\int_{\mathbb{R}^{d}}|\nabla(\Phi_{0}-\bar{\Phi}_{0})|^{2}\,dx+\frac{C}{\varepsilon}\int_{0}^{t}\int_{\mathbb{R}^{d}}|\nabla(\Phi-\bar{\Phi})|^{2}\,dxds\\ &\quad\qquad\,\,+C\varepsilon^{\delta}\int_{0}^{t}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|v|^{2}\,dxdvds\end{aligned} (3.5)

for ε(0,1)\varepsilon\in(0,1), which implies that

d|(ΦΦ¯)|2𝑑xεdρ0|u0u¯0|2𝑑x+2εd(𝒦ε(f0)ε(U0))𝑑x+Cε2+Cεδ+10td×df|v|2𝑑x𝑑v𝑑s+2dp(ρ0|ρ¯0)𝑑x+Cd|(Φ0Φ¯0)|2𝑑x+C0td|(ΦΦ¯)|2𝑑x𝑑s.\displaystyle\begin{aligned} &\int_{\mathbb{R}^{d}}|\nabla(\Phi-\bar{\Phi})|^{2}\,dx\cr&\quad\leq\varepsilon\int_{\mathbb{R}^{d}}\rho_{0}|u_{0}-\bar{u}_{0}|^{2}\,dx+2\varepsilon\int_{\mathbb{R}^{d}}\left(\mathcal{K}_{\varepsilon}(f_{0})-\mathcal{E}_{\varepsilon}(U_{0})\right)dx+C\varepsilon^{2}+C\varepsilon^{\delta+1}\int_{0}^{t}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|v|^{2}\,dxdvds\\ &\qquad+2\int_{\mathbb{R}^{d}}p(\rho_{0}|\bar{\rho}_{0})\,dx+C\int_{\mathbb{R}^{d}}|\nabla(\Phi_{0}-\bar{\Phi}_{0})|^{2}\,dx+C\int_{0}^{t}\int_{\mathbb{R}^{d}}|\nabla(\Phi-\bar{\Phi})|^{2}\,dxds.\end{aligned}

Applying Grönwall’s lemma entails that

d|(ΦΦ¯)|2𝑑x\displaystyle\int_{\mathbb{R}^{d}}|\nabla(\Phi-\bar{\Phi})|^{2}\,dx
Cεdρ0|u0u¯0|2𝑑x+Cdp(ρ0|ρ¯0)𝑑x+Cεd(𝒦ε(f0)ε(U0))𝑑x\displaystyle\quad\leq C\varepsilon\int_{\mathbb{R}^{d}}\rho_{0}|u_{0}-\bar{u}_{0}|^{2}\,dx+C\int_{\mathbb{R}^{d}}p(\rho_{0}|\bar{\rho}_{0})\,dx+C\varepsilon\int_{\mathbb{R}^{d}}\left(\mathcal{K}_{\varepsilon}(f_{0})-\mathcal{E}_{\varepsilon}(U_{0})\right)dx
+Cεδ+10td×df|v|2𝑑x𝑑v𝑑s+Cε2+Cd|(Φ0Φ¯0)|2𝑑x.\displaystyle\qquad\,\,+C\varepsilon^{\delta+1}\int_{0}^{t}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|v|^{2}\,dxdvds+C\varepsilon^{2}+C\int_{\mathbb{R}^{d}}|\nabla(\Phi_{0}-\bar{\Phi}_{0})|^{2}\,dx.

Putting this into (3.5), we have

12dρ|uu¯|2𝑑x+1εdp(ρ|ρ¯)𝑑x+12ε0tdρ|uu¯|2𝑑x𝑑s+12εd|(ΦΦ¯)|2𝑑x\displaystyle\frac{1}{2}\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dx+\frac{1}{\varepsilon}\int_{\mathbb{R}^{d}}p(\rho|\bar{\rho})\,dx+\frac{1}{2\varepsilon}\int_{0}^{t}\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dxds+\frac{1}{2\varepsilon}\int_{\mathbb{R}^{d}}|\nabla(\Phi-\bar{\Phi})|^{2}\,dx
Cdρ0|u0u¯0|2𝑑x+Cεdp(ρ0|ρ¯0)𝑑x+Cd(𝒦ε(f0)ε(U0))𝑑x\displaystyle\quad\leq C\int_{\mathbb{R}^{d}}\rho_{0}|u_{0}-\bar{u}_{0}|^{2}\,dx+\frac{C}{\varepsilon}\int_{\mathbb{R}^{d}}p(\rho_{0}|\bar{\rho}_{0})\,dx+C\int_{\mathbb{R}^{d}}\left(\mathcal{K}_{\varepsilon}(f_{0})-\mathcal{E}_{\varepsilon}(U_{0})\right)dx
+Cεδ0td×df|v|2𝑑x𝑑v𝑑s+Cε+Cεd|(Φ0Φ¯0)|2𝑑x\displaystyle\qquad\,\,+C\varepsilon^{\delta}\int_{0}^{t}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|v|^{2}\,dxdvds+C\varepsilon+\frac{C}{\varepsilon}\int_{\mathbb{R}^{d}}|\nabla(\Phi_{0}-\bar{\Phi}_{0})|^{2}\,dx

for ε(0,1)\varepsilon\in(0,1), where C>0C>0 is independent of ε\varepsilon. This completes the proof. ∎

Proof of Lemma 3.1.

We first notice that

u¯j=xjjΦ¯jlogρ¯forj=1,,d,\bar{u}_{j}=-x_{j}-\partial_{j}\bar{\Phi}-\partial_{j}\log\bar{\rho}\quad\mbox{for}\quad j=1,\dots,d,

and thus

u¯iiu¯j=u¯i(δij+ijΦ¯+ijlogρ¯)fori,j=1,,d.\bar{u}_{i}\partial_{i}\bar{u}_{j}=-\bar{u}_{i}\left(\delta_{ij}+\partial_{ij}\bar{\Phi}+\partial_{ij}\log\bar{\rho}\right)\quad\mbox{for}\quad i,j=1,\dots,d.

For notational simplicity, for the rest of this proof, we omit the summation, i.e., uivi=i=1duiviu_{i}v_{i}=\sum_{i=1}^{d}u_{i}v_{i} and we denote by i=xi\partial_{i}=\partial_{x_{i}} for i=1,,di=1,\dots,d. Then we obtain

dρ(uu¯)(u¯u¯)𝑑x=dρ(uju¯j)u¯j𝑑xdρ(uju¯j)u¯i(ijΦ¯)𝑑xdρ(uju¯j)u¯iijlogρ¯dx.\displaystyle\begin{aligned} \int_{\mathbb{R}^{d}}\rho(u-\bar{u})\cdot(\bar{u}\cdot\nabla\bar{u})\,dx&=-\int_{\mathbb{R}^{d}}\rho(u_{j}-\bar{u}_{j})\bar{u}_{j}\,dx-\int_{\mathbb{R}^{d}}\rho(u_{j}-\bar{u}_{j})\bar{u}_{i}(\partial_{ij}\bar{\Phi})\,dx\cr&\quad-\int_{\mathbb{R}^{d}}\rho(u_{j}-\bar{u}_{j})\bar{u}_{i}\partial_{ij}\log\bar{\rho}\,dx.\end{aligned} (3.6)

On the other hand, we also find (see Remark 1.1)

jtΦ¯=jKtρ¯=ijK(ρ¯u¯i)=dijK(xy)(ρ¯u¯i)(y)dy-\partial_{j}\partial_{t}\bar{\Phi}=-\partial_{j}K\star\partial_{t}\bar{\rho}=\partial_{ij}K\star(\bar{\rho}\bar{u}_{i})=\int_{\mathbb{R}^{d}}\partial_{ij}K(x-y)(\bar{\rho}\bar{u}_{i})(y)\,dy

and

jtlogρ¯=j(u¯iilogρ¯+iu¯i)=(ju¯i)ilogρ¯+u¯iijlogρ¯+iju¯i-\partial_{j}\partial_{t}\log\bar{\rho}=\partial_{j}(\bar{u}_{i}\partial_{i}\log\bar{\rho}+\partial_{i}\bar{u}_{i})=(\partial_{j}\bar{u}_{i})\partial_{i}\log\bar{\rho}+\bar{u}_{i}\partial_{ij}\log\bar{\rho}+\partial_{ij}\bar{u}_{i}

for j=1,,dj=1,\dots,d. This implies

dρ(uu¯)tu¯dx\displaystyle\int_{\mathbb{R}^{d}}\rho(u-\bar{u})\cdot\partial_{t}\bar{u}\,dx =d×dρ(x)(uju¯j)(x)ijK(xy)(ρ¯u¯i)(y)dxdy\displaystyle=\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\rho(x)(u_{j}-\bar{u}_{j})(x)\partial_{ij}K(x-y)(\bar{\rho}\bar{u}_{i})(y)\,dxdy
+dρ(uju¯j)((ju¯i)ilogρ¯+u¯iijlogρ¯+iju¯i)𝑑x.\displaystyle\quad+\int_{\mathbb{R}^{d}}\rho(u_{j}-\bar{u}_{j})\left((\partial_{j}\bar{u}_{i})\partial_{i}\log\bar{\rho}+\bar{u}_{i}\partial_{ij}\log\bar{\rho}+\partial_{ij}\bar{u}_{i}\right)dx.

We now combine this with (3.6) to get

dρ(uu¯)e¯ρ¯𝑑x\displaystyle\int_{\mathbb{R}^{d}}\rho(u-\bar{u})\cdot\frac{\bar{e}}{\bar{\rho}}\,dx =dρ(uu¯)(tu¯+u¯u¯)𝑑x\displaystyle=\int_{\mathbb{R}^{d}}\rho(u-\bar{u})\cdot(\partial_{t}\bar{u}+\bar{u}\cdot\nabla\bar{u})\,dx
=dρ(uju¯j)u¯j𝑑xdρ(uju¯j)u¯i(ijΦ¯)𝑑x\displaystyle=-\int_{\mathbb{R}^{d}}\rho(u_{j}-\bar{u}_{j})\bar{u}_{j}\,dx-\int_{\mathbb{R}^{d}}\rho(u_{j}-\bar{u}_{j})\bar{u}_{i}(\partial_{ij}\bar{\Phi})\,dx
+d×dρ(x)(uju¯j)(x)ijK(xy)(ρ¯u¯i)(y)dxdy\displaystyle\quad+\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\rho(x)(u_{j}-\bar{u}_{j})(x)\partial_{ij}K(x-y)(\bar{\rho}\bar{u}_{i})(y)\,dxdy
+dρ(uju¯j)((ju¯i)ilogρ¯+iju¯i)𝑑x\displaystyle\quad+\int_{\mathbb{R}^{d}}\rho(u_{j}-\bar{u}_{j})\left((\partial_{j}\bar{u}_{i})\partial_{i}\log\bar{\rho}+\partial_{ij}\bar{u}_{i}\right)dx
=:i=14Ji,\displaystyle=:\sum_{i=1}^{4}J_{i},

where Ji,i=1,4J_{i},i=1,4 can be easily estimated as

J1\displaystyle J_{1} =dρ|uu¯|2𝑑xdρ(uu¯)u𝑑x\displaystyle=\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dx-\int_{\mathbb{R}^{d}}\rho(u-\bar{u})\cdot u\,dx
dρ|uu¯|2𝑑x+(dρ|uu¯|2𝑑x)1/2(dρ|u|2𝑑x)1/2\displaystyle\leq\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dx+\left(\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dx\right)^{1/2}\left(\int_{\mathbb{R}^{d}}\rho|u|^{2}\,dx\right)^{1/2}
dρ|uu¯|2𝑑x+C(dρ|uu¯|2𝑑x)1/2\displaystyle\leq\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dx+C\left(\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dx\right)^{1/2}

and

J4\displaystyle J_{4} ((u¯)L+u¯Llogρ¯L)dρ|uu¯|𝑑x\displaystyle\leq\left(\|\nabla(\nabla\cdot\bar{u})\|_{L^{\infty}}+\|\nabla\bar{u}\|_{L^{\infty}}\|\nabla\log\bar{\rho}\|_{L^{\infty}}\right)\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|\,dx
((u¯)L+u¯Llogρ¯L)(dρ|uu¯|2𝑑x)1/2.\displaystyle\leq\left(\|\nabla(\nabla\cdot\bar{u})\|_{L^{\infty}}+\|\nabla\bar{u}\|_{L^{\infty}}\|\nabla\log\bar{\rho}\|_{L^{\infty}}\right)\left(\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dx\right)^{1/2}.

Here we used

dρ𝑑x=1anddρ|u|2𝑑xd×df|v|2𝑑x𝑑vC\int_{\mathbb{R}^{d}}\rho\,dx=1\quad\mbox{and}\quad\int_{\mathbb{R}^{d}}\rho|u|^{2}\,dx\leq\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|v|^{2}\,dxdv\leq C

for some C>0C>0 independent of solutions (ρ¯,u¯)(\bar{\rho},\bar{u}) and ε\varepsilon due to Lemma 2.1. We next estimate

|J2+J3|\displaystyle\left|J_{2}+J_{3}\right| =|d×dρ(x)(uju¯j)(x)ijK(xy)ρ¯(y)(u¯i(y)u¯i(x))dxdy|\displaystyle=\left|\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\rho(x)(u_{j}-\bar{u}_{j})(x)\partial_{ij}K(x-y)\bar{\rho}(y)(\bar{u}_{i}(y)-\bar{u}_{i}(x))\,dxdy\right|
u¯Ldρ(x)|(uu¯)(x)|(d1|xy|d1ρ¯(y)𝑑y)𝑑x\displaystyle\leq\|\nabla\bar{u}\|_{L^{\infty}}\int_{\mathbb{R}^{d}}\rho(x)|(u-\bar{u})(x)|\left(\int_{\mathbb{R}^{d}}\frac{1}{|x-y|^{d-1}}\bar{\rho}(y)\,dy\right)\,dx
Cu¯Lρ¯L1Ldρ|uu¯|𝑑x\displaystyle\leq C\|\nabla\bar{u}\|_{L^{\infty}}\|\bar{\rho}\|_{L^{1}\cap L^{\infty}}\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|\,dx
Cu¯Lρ¯L1L(dρ|uu¯|2𝑑x)1/2,\displaystyle\leq C\|\nabla\bar{u}\|_{L^{\infty}}\|\bar{\rho}\|_{L^{1}\cap L^{\infty}}\left(\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dx\right)^{1/2},

where C>0C>0 is independent of solutions (ρ¯,u¯)(\bar{\rho},\bar{u}) and ε\varepsilon.

We next rewrite the assumption on the regularity of solutions in terms of ρ¯\bar{\rho} only, not u¯\bar{u}. In fact, this is not that hard, simply we find

u¯=IdKρ¯2logρ¯\nabla\bar{u}=-I_{d}-\nabla K\star\nabla\bar{\rho}-\nabla^{2}\log\bar{\rho}

and

u¯=d+ρ¯Δlogρ¯.\nabla\cdot\bar{u}=-d+\bar{\rho}-\Delta\log\bar{\rho}.

This yields

u¯LC(1+ρ¯L1L+2logρ¯L)\|\nabla\bar{u}\|_{L^{\infty}}\leq C\left(1+\|\nabla\bar{\rho}\|_{L^{1}\cap L^{\infty}}+\|\nabla^{2}\log\bar{\rho}\|_{L^{\infty}}\right)

and

(u¯)Lρ¯L+Δlogρ¯L.\|\nabla(\nabla\cdot\bar{u})\|_{L^{\infty}}\leq\|\nabla\bar{\rho}\|_{L^{\infty}}+\|\nabla\Delta\log\bar{\rho}\|_{L^{\infty}}.

This completes the proof. ∎

3.2. Proof of Theorem 1.1

In order to close the relative entropy estimate in Proposition 3.1, we need to handle the kinetic energy term on the right hand side of (3.1).

3.2.1. Case with the confinement

In this case, we show that the kinetic energy can be controlled by the free energy ε(f)\mathcal{F}_{\varepsilon}(f). For this, we need to estimate the negative part of the entropy term.

Note that there exists a positive constant CC such that the following estimate

g|logg|=glogg2gloggχ0g1glogg+2(ωg+Ceω/2)\displaystyle g|\log g|=g\log g-2g\log g\,\chi_{0\leq g\leq 1}\leq g\log g+2\left(\omega g+Ce^{-\omega/2}\right) (3.7)

holds for g,ω0g,\omega\geq 0, where χ\chi is a characteristic function. We then take g=fg=f and ω=ε|v|2+|x|28\omega=\frac{\varepsilon|v|^{2}+|x|^{2}}{8} in (3.7) to have

d×df|logf|𝑑x𝑑vd×dflogfdxdv+d×dε|v|2+|x|24f𝑑x𝑑v+Cεd/2.\displaystyle\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|\log f|\,dxdv\leq\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f\log f\,dxdv+\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\frac{\varepsilon|v|^{2}+|x|^{2}}{4}f\,dxdv+\frac{C}{\varepsilon^{d/2}}.

This implies

1εd×df|logf|𝑑x𝑑v\displaystyle\frac{1}{\varepsilon}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|\log f|\,dxdv 1εd×dflogfdxdv+d×d|v|24f𝑑x𝑑v\displaystyle\leq\frac{1}{\varepsilon}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f\log f\,dxdv+\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\frac{|v|^{2}}{4}f\,dxdv
+1εd×d|x|24f𝑑x𝑑v+Cεd/2+1.\displaystyle\quad+\frac{1}{\varepsilon}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\frac{|x|^{2}}{4}f\,dxdv+\frac{C}{\varepsilon^{d/2+1}}.

Thus we obtain

ε(f)=1εd×dflogfdxdv+d×df|v|22𝑑x𝑑v+1εd×df|x|22𝑑x+12εdΦρ𝑑x1εd×df|logf|𝑑x𝑑v+14d×df|v|2𝑑x𝑑v+14εd×df|x|2𝑑x𝑑v+12εdΦρ𝑑xCεd/2+1,\displaystyle\begin{aligned} \mathcal{F}_{\varepsilon}(f)&=\frac{1}{\varepsilon}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f\log f\,dxdv+\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f\frac{|v|^{2}}{2}\,dxdv+\frac{1}{\varepsilon}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f\frac{|x|^{2}}{2}\,dx+\frac{1}{2\varepsilon}\int_{\mathbb{R}^{d}}\Phi\rho\,dx\\ &\geq\frac{1}{\varepsilon}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|\log f|\,dxdv+\frac{1}{4}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|v|^{2}\,dxdv\\ &\quad+\frac{1}{4\varepsilon}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|x|^{2}\,dxdv+\frac{1}{2\varepsilon}\int_{\mathbb{R}^{d}}\Phi\rho\,dx-\frac{C}{\varepsilon^{d/2+1}},\end{aligned}

and subsequently this together with Lemma 2.1 yields

d×df|v|2𝑑x𝑑v4ε(f)+Cεd/2+1(4ε(f0)+Cεd/2+1)+2εδ0td×df|v|2𝑑x𝑑v𝑑s.\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|v|^{2}\,dxdv\leq 4\mathcal{F}_{\varepsilon}(f)+\frac{C}{\varepsilon^{d/2+1}}\leq\left(4\mathcal{F}_{\varepsilon}(f_{0})+\frac{C}{\varepsilon^{d/2+1}}\right)+2\varepsilon^{\delta}\int_{0}^{t}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|v|^{2}\,dxdvds.

We now apply Grönwall’s lemma to the above to have

0td×df|v|2𝑑x𝑑v𝑑se2εδ(4ε(f0)+Cεd/2+1)(e2εδt1)Cε(f0)+Cεd/21,\int_{0}^{t}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|v|^{2}\,dxdvds\leq e^{-2\varepsilon^{\delta}}\left(4\mathcal{F}_{\varepsilon}(f_{0})+\frac{C}{\varepsilon^{d/2+1}}\right)\left(e^{2\varepsilon^{\delta}t}-1\right)\leq C\mathcal{F}_{\varepsilon}(f_{0})+C\varepsilon^{-d/2-1},

where we used 4ε(f0)+Cεd/2104\mathcal{F}_{\varepsilon}(f_{0})+C\varepsilon^{-d/2-1}\geq 0. This combined with Proposition 3.1 concludes the desired result.

Remark 3.1.

If we further assume u¯L(d×(0,T))\bar{u}\in L^{\infty}(\mathbb{R}^{d}\times(0,T)) i.e. e¯=tu¯+u¯u¯L(d×(0,T))\bar{e}=\partial_{t}\bar{u}+\bar{u}\cdot\nabla\bar{u}\in L^{\infty}(\mathbb{R}^{d}\times(0,T)), by the arguments used in [9, 28], then we can easily estimate

dρ(uu¯)e¯ρ¯𝑑x\displaystyle\int_{\mathbb{R}^{d}}\rho(u-\bar{u})\cdot\frac{\bar{e}}{\bar{\rho}}\,dx 18εdρ|uu¯|2𝑑x+Cεdρ|tu¯+u¯u¯|2𝑑x\displaystyle\leq\frac{1}{8\varepsilon}\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dx+C\varepsilon\int_{\mathbb{R}^{d}}\rho\left|\partial_{t}\bar{u}+\bar{u}\cdot\nabla\bar{u}\right|^{2}dx
18εdρ|uu¯|2𝑑x+Cεe¯L2ρL1,\displaystyle\leq\frac{1}{8\varepsilon}\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dx+C\varepsilon\|\bar{e}\|_{L^{\infty}}^{2}\|\rho\|_{L^{1}},

where C>0C>0 is independent of ε>0\varepsilon>0. However, in the case with confinement, the velocity fields u¯\bar{u} is given by u¯=xΦ¯logρ¯\bar{u}=-x-\nabla\bar{\Phi}-\nabla\log\bar{\rho}, and thus it seems impossible to assume the LL^{\infty}-bound on u¯\bar{u}.

3.2.2. Case without the confinement

Differently from the case with the confinement, in this case, we first control the (mesoscopic) kinetic energy by using the (macroscopic) kinetic and interaction energies.

Following a similar way as in (2.2), we deduce

ddt(d×df|v|22𝑑x𝑑v+12εdΦρ𝑑x)+1εd×df|v|2𝑑x𝑑v+1ε2+δd×df|vu|2𝑑x𝑑v=dε3+δ.\displaystyle\begin{aligned} &\frac{d}{dt}\left(\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f\frac{|v|^{2}}{2}\,dxdv+\frac{1}{2\varepsilon}\int_{\mathbb{R}^{d}}\Phi\rho\,dx\right)\cr&\quad+\frac{1}{\varepsilon}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|v|^{2}\,dxdv+\frac{1}{\varepsilon^{2+\delta}}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|v-u|^{2}\,dxdv=\frac{d}{\varepsilon^{3+\delta}}.\end{aligned} (3.8)

Integrating (3.8) from 0 to tt, one has

0td×df|uv|2𝑑x𝑑v𝑑sε2+δEε(f0)+dTε,\displaystyle\begin{aligned} \int_{0}^{t}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|u-v|^{2}\,dxdvds\leq\varepsilon^{2+\delta}E_{\varepsilon}(f_{0})+\frac{dT}{\varepsilon},\end{aligned}

where Eε(f0)E_{\varepsilon}(f_{0}) is given as in (1.8). Since

d×df|uv|2𝑑x𝑑v=d×df|v|2𝑑x𝑑vdρ|u|2𝑑x,\displaystyle\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|u-v|^{2}\,dxdv=\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|v|^{2}\,dxdv-\int_{\mathbb{R}^{d}}\rho|u|^{2}\,dx,

we further have

0td×df|v|2𝑑x𝑑v𝑑sε2+δEε(f0)+Cε+0tdρ|u|2𝑑x𝑑s.\int_{0}^{t}\iint_{\mathbb{R}^{d}\times\mathbb{R}^{d}}f|v|^{2}\,dxdvds\leq\varepsilon^{2+\delta}E_{\varepsilon}(f_{0})+\frac{C}{\varepsilon}+\int_{0}^{t}\int_{\mathbb{R}^{d}}\rho|u|^{2}\,dxds.

On the other hand, the (macroscopic) kinetic energy can be estimated as

dρ|u|2𝑑x\displaystyle\int_{\mathbb{R}^{d}}\rho|u|^{2}\,dx 2dρ|uu¯|2𝑑x+2dρ|u¯|2𝑑x\displaystyle\leq 2\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dx+2\int_{\mathbb{R}^{d}}\rho|\bar{u}|^{2}\,dx
14εδ+1dρ|uu¯|2𝑑x+2(Φ¯+logρ¯)L\displaystyle\leq\frac{1}{4\varepsilon^{\delta+1}}\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dx+2\|\nabla(\bar{\Phi}+\log\bar{\rho})\|_{L^{\infty}}

for ε(0,1)\varepsilon\in(0,1) small enough. Combining this and Proposition 3.1 completes the proof.

3.3. Proof of Corollary 1.1

By Taylor’s theorem, we first easily find

ρ¯ρρzz𝑑z12min{1ρ¯,1ρ}(ρρ¯)2.\int_{\bar{\rho}}^{\rho}\frac{\rho-z}{z}\,dz\geq\frac{1}{2}\min\left\{\frac{1}{\bar{\rho}},\frac{1}{\rho}\right\}(\rho-\bar{\rho})^{2}.

We then estimate

(d|ρρ¯|𝑑x)2(d(ρ+ρ¯)𝑑x)(dmin{1ρ¯,1ρ}(ρρ¯)2𝑑x)Cdρ¯ρρzz𝑑z𝑑x,\displaystyle\begin{aligned} \left(\int_{\mathbb{R}^{d}}|\rho-\bar{\rho}|\,dx\right)^{2}&\leq\left(\int_{\mathbb{R}^{d}}(\rho+\bar{\rho})\,dx\right)\left(\int_{\mathbb{R}^{d}}\min\left\{\frac{1}{\bar{\rho}},\frac{1}{\rho}\right\}(\rho-\bar{\rho})^{2}\,dx\right)\cr&\leq C\int_{\mathbb{R}^{d}}\int_{\bar{\rho}}^{\rho}\frac{\rho-z}{z}\,dzdx,\end{aligned} (3.9)

where C>0C>0 is independent of ε>0\varepsilon>0 and we used 1(x+y)min{x1,y1}1\leq(x+y)\min\{x^{-1},y^{-1}\} for x,y>0x,y>0. This asserts the convergence of ρρ¯\rho\to\bar{\rho} in L(0,T;L1(d))L^{\infty}(0,T;L^{1}(\mathbb{R}^{d})) as ε0\varepsilon\to 0.

Moreover, for any ψH1(d)\psi\in H^{1}(\mathbb{R}^{d}) with ψH11\|\psi\|_{H^{1}}\leq 1, we obtain

|dψ(x)(ρρ¯)(x)𝑑x|\displaystyle\left|\int_{\mathbb{R}^{d}}\psi(x)(\rho-\bar{\rho})(x)\,dx\right| =|dψ(x)Δ(ΦΦ¯)(x)𝑑x|\displaystyle=\left|\int_{\mathbb{R}^{d}}\psi(x)\Delta(\Phi-\bar{\Phi})(x)\,dx\right|
=|dψ(x)(ΦΦ¯)(x)𝑑x|\displaystyle=\left|\int_{\mathbb{R}^{d}}\nabla\psi(x)\cdot\nabla(\Phi-\bar{\Phi})(x)\,dx\right|
(ΦΦ¯)L2,\displaystyle\leq\|\nabla(\Phi-\bar{\Phi})\|_{L^{2}},

and this yields

ρρ¯H1(ΦΦ¯)L2.\|\rho-\bar{\rho}\|_{H^{-1}}\leq\|\nabla(\Phi-\bar{\Phi})\|_{L^{2}}.

Thus we have the convergence of ρρ¯\rho\to\bar{\rho} in L(0,T;H1(d))L^{\infty}(0,T;H^{-1}(\mathbb{R}^{d})) as ε0\varepsilon\to 0.

For the convergence of ρu\rho u towards ρ¯u¯\bar{\rho}\bar{u}, we estimate

d|ρρ¯||u¯|𝑑x\displaystyle\int_{\mathbb{R}^{d}}|\rho-\bar{\rho}||\bar{u}|\,dx d|ρρ¯||V|𝑑x+(Φ¯+logρ¯)Ld|ρρ¯|𝑑x\displaystyle\leq\int_{\mathbb{R}^{d}}|\rho-\bar{\rho}||\nabla V|\,dx+\|\nabla(\bar{\Phi}+\log\bar{\rho})\|_{L^{\infty}}\int_{\mathbb{R}^{d}}|\rho-\bar{\rho}|\,dx
C(d|ρρ¯|𝑑x)1/2(d(ρ+ρ¯)(1+V)𝑑x)1/2+Cd|ρρ¯|𝑑x,\displaystyle\leq C\left(\int_{\mathbb{R}^{d}}|\rho-\bar{\rho}|\,dx\right)^{1/2}\left(\int_{\mathbb{R}^{d}}(\rho+\bar{\rho})(1+V)\,dx\right)^{1/2}+C\int_{\mathbb{R}^{d}}|\rho-\bar{\rho}|\,dx,

where C>0C>0 depends on (Φ¯+logρ¯)L\|\nabla(\bar{\Phi}+\log\bar{\rho})\|_{L^{\infty}}. This deduces

d|ρuρ¯u¯|𝑑x\displaystyle\int_{\mathbb{R}^{d}}|\rho u-\bar{\rho}\bar{u}|\,dx dρ|uu¯|𝑑x+d|ρρ¯||u¯|𝑑x\displaystyle\leq\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|\,dx+\int_{\mathbb{R}^{d}}|\rho-\bar{\rho}||\bar{u}|\,dx
(dρ𝑑x)1/2(dρ|uu¯|2𝑑x)1/2+Cd|ρρ¯|𝑑x\displaystyle\leq\left(\int_{\mathbb{R}^{d}}\rho\,dx\right)^{1/2}\left(\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dx\right)^{1/2}+C\int_{\mathbb{R}^{d}}|\rho-\bar{\rho}|\,dx
+C(d|ρρ¯|𝑑x)1/2(d(ρ+ρ¯)(1+V)𝑑x)1/2.\displaystyle\quad+C\left(\int_{\mathbb{R}^{d}}|\rho-\bar{\rho}|\,dx\right)^{1/2}\left(\int_{\mathbb{R}^{d}}(\rho+\bar{\rho})(1+V)\,dx\right)^{1/2}.

We finally combine this with (3.9) to have

ρuρ¯u¯L12Cdρ|uu¯|2𝑑x+C(dρ¯ρρzz𝑑z𝑑x)1/2+Cdρ¯ρρzz𝑑z𝑑x,\|\rho u-\bar{\rho}\bar{u}\|_{L^{1}}^{2}\leq C\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dx+C\left(\int_{\mathbb{R}^{d}}\int_{\bar{\rho}}^{\rho}\frac{\rho-z}{z}\,dzdx\right)^{1/2}+C\int_{\mathbb{R}^{d}}\int_{\bar{\rho}}^{\rho}\frac{\rho-z}{z}\,dzdx,

where C>0C>0 depends on (Φ¯+logρ¯)L\|\nabla(\bar{\Phi}+\log\bar{\rho})\|_{L^{\infty}}. Integrating the above inequality over the time interval [0,T][0,T] concludes the desired convergence estimate.

Remark 3.2.

If Φ\Phi is given as Φ=K~ρ\Phi=\tilde{K}\star\rho with K~L(d)\nabla\tilde{K}\in L^{\infty}(\mathbb{R}^{d}), then we can estimate the fifth term on the right hand side of (3.2) as

1ε|dρ(uu¯)K~(ρ¯ρ)𝑑x|\displaystyle\frac{1}{\varepsilon}\left|\int_{\mathbb{R}^{d}}\rho(u-\bar{u})\cdot\nabla\tilde{K}\star(\bar{\rho}-\rho)\,dx\right| 1εK~Lρρ¯L1dρ|uu¯|𝑑x\displaystyle\leq\frac{1}{\varepsilon}\|\nabla\tilde{K}\|_{L^{\infty}}\|\rho-\bar{\rho}\|_{L^{1}}\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|\,dx
Cε(dρ¯ρρzz𝑑z𝑑x)1/2(dρ|uu¯|2𝑑x)1/2\displaystyle\leq\frac{C}{\varepsilon}\left(\int_{\mathbb{R}^{d}}\int_{\bar{\rho}}^{\rho}\frac{\rho-z}{z}\,dzdx\right)^{1/2}\left(\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dx\right)^{1/2}
Cεdρ¯ρρzz𝑑z𝑑x+14εdρ|uu¯|2𝑑x.\displaystyle\leq\frac{C}{\varepsilon}\int_{\mathbb{R}^{d}}\int_{\bar{\rho}}^{\rho}\frac{\rho-z}{z}\,dzdx+\frac{1}{4\varepsilon}\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dx.

This combined with almost the same arguments as before concludes the same convergence estimates appeared in Corollary 1.1.

3.4. Remarks on the regularity assumptions on the limiting system

In this part, we provide some estimates on logρ¯L(0,T;W2,)\|\nabla\log\bar{\rho}\|_{L^{\infty}(0,T;W^{2,\infty})} when V=|x|2/2V=|x|^{2}/2. The estimates directly cover the case V0V\equiv 0. Let us start with the estimate of logρ¯L\|\nabla\log\bar{\rho}\|_{L^{\infty}}. For notational simplicity, we set u~:=VΦ¯\tilde{u}:=-\nabla V-\nabla\bar{\Phi}, we again omit the summation, and we denote by i=xi\partial_{i}=\partial_{x_{i}} for i=1,,di=1,\dots,d.

We first show the LL^{\infty}-bound on logρ¯\nabla\log\bar{\rho} in the lemma below.

Lemma 3.2.

There exists C>0C>0 depends only on ρ¯L(0,T;L1L)\|\nabla\bar{\rho}\|_{L^{\infty}(0,T;L^{1}\cap L^{\infty})} and T>0T>0 such that

sup0tTlogρ¯(,t)LC(1+logρ¯0L).\sup_{0\leq t\leq T}\|\nabla\log\bar{\rho}(\cdot,t)\|_{L^{\infty}}\leq C(1+\|\nabla\log\bar{\rho}_{0}\|_{L^{\infty}}).
Proof.

Note that ilogρ¯\partial_{i}\log\bar{\rho} satisfies

tilogρ¯+u~ilogρ¯=iu~iu~logρ¯+Δilogρ¯+2ilogρ¯logρ¯\partial_{t}\partial_{i}\log\bar{\rho}+\tilde{u}\cdot\nabla\partial_{i}\log\bar{\rho}=-\nabla\cdot\partial_{i}\tilde{u}-\partial_{i}\tilde{u}\cdot\nabla\log\bar{\rho}+\Delta\partial_{i}\log\bar{\rho}+2\nabla\partial_{i}\log\bar{\rho}\cdot\nabla\log\bar{\rho}

for i=1,,di=1,\dots,d. For a given tt, at any local Maximum point of ilogρ¯\partial_{i}\log\bar{\rho}, we get

Δilogρ¯0andilogρ¯=0.\Delta\partial_{i}\log\bar{\rho}\leq 0\quad\mbox{and}\quad\nabla\partial_{i}\log\bar{\rho}=0.

Using this observation together with an elementary estimate yields

ddtlogρ¯LC((u~)L+u~Llogρ¯L),\frac{d}{dt}\|\nabla\log\bar{\rho}\|_{L^{\infty}}\leq C(\|\nabla(\nabla\cdot\tilde{u})\|_{L^{\infty}}+\|\nabla\tilde{u}\|_{L^{\infty}}\|\nabla\log\bar{\rho}\|_{L^{\infty}}),

and applying Grönwall’s lemma gives

logρ¯LC(1+logρ¯0L),\|\nabla\log\bar{\rho}\|_{L^{\infty}}\leq C(1+\|\nabla\log\bar{\rho}_{0}\|_{L^{\infty}}),

where C>0C>0 depends only on (u~)L\|\nabla(\nabla\cdot\tilde{u})\|_{L^{\infty}}, u~L\|\nabla\tilde{u}\|_{L^{\infty}}, and T>0T>0. On the other hand, similarly as in the proof of Lemma 3.1, we easily get

u~LC(1+ρ¯L1L)\|\nabla\tilde{u}\|_{L^{\infty}}\leq C\left(1+\|\nabla\bar{\rho}\|_{L^{1}\cap L^{\infty}}\right)

and

(u~)Lρ¯L.\|\nabla(\nabla\cdot\tilde{u})\|_{L^{\infty}}\leq\|\nabla\bar{\rho}\|_{L^{\infty}}.

This completes the proof. ∎

We next provide higher-order estimates on logρ¯\nabla\log\bar{\rho}.

Lemma 3.3.

There exists 0<TT0<T_{*}\leq T such that

sup0tT2logρ¯(,t)LC1\sup_{0\leq t\leq T_{*}}\|\nabla^{2}\log\bar{\rho}(\cdot,t)\|_{L^{\infty}}\leq C_{1}

and

sup0tT3logρ¯(,t)LC2,\sup_{0\leq t\leq T_{*}}\|\nabla^{3}\log\bar{\rho}(\cdot,t)\|_{L^{\infty}}\leq C_{2},

where Ck>0,k=1,2C_{k}>0,k=1,2 depends only on logρ¯0Wk,\|\nabla\log\bar{\rho}_{0}\|_{W^{k,\infty}} and ρ¯L(0,T;Wk,1Wk,)\|\nabla\bar{\rho}\|_{L^{\infty}(0,T;W^{k,1}\cap W^{k,\infty})}.

Proof.

For i,j=1,,di,j=1,\dots,d, we get

tijlogρ¯+u~ijlogρ¯\displaystyle\partial_{t}\partial_{ij}\log\bar{\rho}+\tilde{u}\cdot\nabla\partial_{ij}\log\bar{\rho} =ju~ilogρ¯iju~iju~logρ¯iu~jlogρ¯\displaystyle=-\partial_{j}\tilde{u}\cdot\nabla\partial_{i}\log\bar{\rho}-\nabla\cdot\partial_{ij}\tilde{u}-\partial_{ij}\tilde{u}\cdot\nabla\log\bar{\rho}-\partial_{i}\tilde{u}\cdot\nabla\partial_{j}\log\bar{\rho}
+Δijlogρ¯+2ijlogρ¯logρ¯+2ilogρ¯jlogρ¯.\displaystyle\quad+\Delta\partial_{ij}\log\bar{\rho}+2\nabla\partial_{ij}\log\bar{\rho}\cdot\nabla\log\bar{\rho}+2\nabla\partial_{i}\log\bar{\rho}\cdot\nabla\partial_{j}\log\bar{\rho}.

Then similarly as before, we obtain

ddt2logρ¯(,t)L\displaystyle\frac{d}{dt}\|\nabla^{2}\log\bar{\rho}(\cdot,t)\|_{L^{\infty}}
u~L2logρ¯L+2(u~)L+2u~Llogρ¯L+2logρ¯L2\displaystyle\quad\lesssim\|\nabla\tilde{u}\|_{L^{\infty}}\|\nabla^{2}\log\bar{\rho}\|_{L^{\infty}}+\|\nabla^{2}(\nabla\cdot\tilde{u})\|_{L^{\infty}}+\|\nabla^{2}\tilde{u}\|_{L^{\infty}}\|\nabla\log\bar{\rho}\|_{L^{\infty}}+\|\nabla^{2}\log\bar{\rho}\|_{L^{\infty}}^{2}
C(1+ρ¯W1,1W1,)(1+logρ¯L+2logρ¯L+2logρ¯L2).\displaystyle\quad\leq C(1+\|\nabla\bar{\rho}\|_{W^{1,1}\cap W^{1,\infty}})(1+\|\nabla\log\bar{\rho}\|_{L^{\infty}}+\|\nabla^{2}\log\bar{\rho}\|_{L^{\infty}}+\|\nabla^{2}\log\bar{\rho}\|_{L^{\infty}}^{2}).

This gives the first assertion.

We next estimate 3logρ¯L\|\nabla^{3}\log\bar{\rho}\|_{L^{\infty}}. For i,j,k=1,,di,j,k=1,\dots,d, we find

tijklogρ¯+u~ijklogρ¯\displaystyle\partial_{t}\partial_{ijk}\log\bar{\rho}+\tilde{u}\cdot\nabla\partial_{ijk}\log\bar{\rho} =ku~ijlogρ¯jku~ilogρ¯ju~iklogρ¯\displaystyle=-\partial_{k}\tilde{u}\cdot\nabla\partial_{ij}\log\bar{\rho}-\partial_{jk}\tilde{u}\cdot\nabla\partial_{i}\log\bar{\rho}-\partial_{j}\tilde{u}\cdot\nabla\partial_{ik}\log\bar{\rho}
ijku~ijku~logρ¯iju~klogρ¯\displaystyle\quad-\nabla\cdot\partial_{ijk}\tilde{u}-\partial_{ijk}\tilde{u}\cdot\nabla\log\bar{\rho}-\partial_{ij}\tilde{u}\cdot\nabla\partial_{k}\log\bar{\rho}
iku~jlogρ¯iu~jklogρ¯+Δijklogρ¯\displaystyle\quad-\partial_{ik}\tilde{u}\cdot\nabla\partial_{j}\log\bar{\rho}-\partial_{i}\tilde{u}\cdot\nabla\partial_{jk}\log\bar{\rho}+\Delta\partial_{ijk}\log\bar{\rho}
+2ijklogρ¯logρ¯+2ijlogρ¯klogρ¯\displaystyle\quad+2\nabla\partial_{ijk}\log\bar{\rho}\cdot\nabla\log\bar{\rho}+2\nabla\partial_{ij}\log\bar{\rho}\cdot\nabla\partial_{k}\log\bar{\rho}
+2iklogρ¯jlogρ¯+2ilogρ¯jklogρ¯.\displaystyle\quad+2\nabla\partial_{ik}\log\bar{\rho}\cdot\nabla\partial_{j}\log\bar{\rho}+2\nabla\partial_{i}\log\bar{\rho}\cdot\nabla\partial_{jk}\log\bar{\rho}.

Then by a similar fashion as above, we obtain

ddt3logρ¯(,t)L\displaystyle\frac{d}{dt}\|\nabla^{3}\log\bar{\rho}(\cdot,t)\|_{L^{\infty}} 3logρ¯L(u~L+2logρ¯L)\displaystyle\lesssim\|\nabla^{3}\log\bar{\rho}\|_{L^{\infty}}\left(\|\nabla\tilde{u}\|_{L^{\infty}}+\|\nabla^{2}\log\bar{\rho}\|_{L^{\infty}}\right)
+2logρ¯L2u~L+logρ¯L3u~L+3(u~)L.\displaystyle\quad+\|\nabla^{2}\log\bar{\rho}\|_{L^{\infty}}\|\nabla^{2}\tilde{u}\|_{L^{\infty}}+\|\nabla\log\bar{\rho}\|_{L^{\infty}}\|\nabla^{3}\tilde{u}\|_{L^{\infty}}+\|\nabla^{3}(\nabla\cdot\tilde{u})\|_{L^{\infty}}.

Since 3u~L+3(u~)LC3ρ¯L1L\|\nabla^{3}\tilde{u}\|_{L^{\infty}}+\|\nabla^{3}(\nabla\cdot\tilde{u})\|_{L^{\infty}}\leq C\|\nabla^{3}\bar{\rho}\|_{L^{1}\cap L^{\infty}}, we conclude the second assertion. ∎

The above two lemmas imply

sup0tTlogρ¯(,t)W2,C,\sup_{0\leq t\leq T_{*}}\|\nabla\log\bar{\rho}(\cdot,t)\|_{W^{2,\infty}}\leq C,

for some T>0T_{*}>0 and C>0C>0 depends only on logρ¯0W2,\|\nabla\log\bar{\rho}_{0}\|_{W^{2,\infty}} and ρ¯L(0,T;W2,1W2,)\|\nabla\bar{\rho}\|_{L^{\infty}(0,T;W^{2,1}\cap W^{2,\infty})}.

Combining all of the above discussion yields that Lemma 3.1 can be restated as

Lemma 3.4.

There exists C>0C>0 depending only on ρ¯L(0,T;W3,1W3,)\|\bar{\rho}\|_{L^{\infty}(0,T;W^{3,1}\cap W^{3,\infty})} and logρ¯0W2,\|\nabla\log\bar{\rho}_{0}\|_{W^{2,\infty}} such that

dρ(uu¯)e¯ρ¯𝑑xdρ|uu¯|2𝑑x+C(dρ|uu¯|2𝑑x)1/2.\int_{\mathbb{R}^{d}}\rho(u-\bar{u})\cdot\frac{\bar{e}}{\bar{\rho}}\,dx\leq\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dx+C\left(\int_{\mathbb{R}^{d}}\rho|u-\bar{u}|^{2}\,dx\right)^{1/2}.
Remark 3.3.

For the periodic domain case 𝕋d\mathbb{T}^{d}, the bound on ρ¯L(0,T;W3,1)\|\bar{\rho}\|_{L^{\infty}(0,T;W^{3,1})} is not required, thus the constant C>0C>0 appeared in Lemma 3.4 only depends on ρ¯L(0,T;W3,)\|\bar{\rho}\|_{L^{\infty}(0,T;W^{3,\infty})} and logρ¯0W2,\|\nabla\log\bar{\rho}_{0}\|_{W^{2,\infty}}. Thus if one can establish a solution ρ¯\bar{\rho} in L(0,T;Hs(𝕋d))L^{\infty}(0,T;H^{s}(\mathbb{T}^{d})) with s>d/2+3s>d/2+3, then we have the bound ρ¯L(0,T;W3,)<\|\bar{\rho}\|_{L^{\infty}(0,T;W^{3,\infty})}<\infty due to the Sobolev embedding.

4. Regularity Estimates for ρ¯\bar{\rho}

In this section, we provide the regularity estimates for ρ¯\bar{\rho} used in the arguments in the previous section. To this end, we first simplify the aggregation-diffusion equation (1.3) by substituting its second equation into the first one, we have

tρ¯=Δρ¯+(ρ¯(Φ¯+V)),ΔΦ¯=ρ¯.\partial_{t}\bar{\rho}=\Delta\bar{\rho}+\nabla\cdot(\bar{\rho}(\nabla\bar{\Phi}+\nabla V)),\quad-\Delta\bar{\Phi}=\bar{\rho}. (4.1)

Since we consider the quadratic confinement V(x)=|x|2/2V(x)=|x|^{2}/2, (4.1) thus reduces to

tρ¯=Δρ¯+(ρ¯Φ¯+xρ¯),ΔΦ¯=ρ¯.\partial_{t}\bar{\rho}=\Delta\bar{\rho}+\nabla\cdot(\bar{\rho}\nabla\bar{\Phi}+x\bar{\rho}),\quad-\Delta\bar{\Phi}=\bar{\rho}. (4.2)

Taking the change of variables, motivated from [11, Section 3],

ρ¯(x,t)=edtn(etx,12(e2t1))=:edtn(x¯,t¯)\bar{\rho}(x,t)=e^{dt}n(e^{t}x,\frac{1}{2}(e^{2t}-1))=:e^{dt}n(\bar{x},\bar{t})

with x¯=etx\bar{x}=e^{t}x and t¯=12(e2t1)\bar{t}=\frac{1}{2}(e^{2t}-1) in (4.2), then we can compute

tρ¯(x,t)\displaystyle\partial_{t}\bar{\rho}(x,t) =dedtn(x¯,t¯)+e(d+1)txx¯n(x¯,t¯)+e(d+2)tt¯n(x¯,t¯),\displaystyle=de^{dt}n(\bar{x},\bar{t})+e^{(d+1)t}x\cdot\nabla_{\bar{x}}n(\bar{x},\bar{t})+e^{(d+2)t}\partial_{\bar{t}}n(\bar{x},\bar{t}),
xρ¯(x,t)\displaystyle\nabla_{x}\bar{\rho}(x,t) =e(d+1)tx¯n(x¯,t¯),\displaystyle=e^{(d+1)t}\nabla_{\bar{x}}n(\bar{x},\bar{t}),
Δxρ¯(x,t)\displaystyle\Delta_{x}\bar{\rho}(x,t) =x(xρ¯(x,t))=x(e(d+1)tx¯n(x¯,t¯))=e(d+2)tΔx¯n(x¯,t¯),\displaystyle=\nabla_{x}\cdot(\nabla_{x}\bar{\rho}(x,t))=\nabla_{x}\cdot(e^{(d+1)t}\nabla_{\bar{x}}n(\bar{x},\bar{t}))=e^{(d+2)t}\Delta_{\bar{x}}n(\bar{x},\bar{t}),

and

x(xρ¯(x,t))=x(xedtn(x¯,t¯))=dedtn(x¯,t¯)+xe(d+1)tx¯n(x¯,t¯).\nabla_{x}\cdot(x\bar{\rho}(x,t))=\nabla_{x}\cdot(xe^{dt}n(\bar{x},\bar{t}))=de^{dt}n(\bar{x},\bar{t})+xe^{(d+1)t}\cdot\nabla_{\bar{x}}n(\bar{x},\bar{t}).

From the Poisson equation in (4.2), one can express Φ¯(x,t)\bar{\Phi}(x,t) uniquely as Φ¯(x,t)=Kρ¯\bar{\Phi}(x,t)=K\star\bar{\rho}, where KK is defined as in Remark 1.1. Then we deduce that

x(ρ¯(x,t)xΦ¯(x,t))=xρ¯(x,t)xΦ¯(x,t)+ρ¯(x,t)ΔxΦ¯(x,t)=xρ¯(x,t)xΦ¯(x,t)ρ¯2(x,t).\displaystyle\begin{aligned} \nabla_{x}\cdot\left(\bar{\rho}(x,t)\nabla_{x}\bar{\Phi}(x,t)\right)&=\nabla_{x}\bar{\rho}(x,t)\cdot\nabla_{x}\bar{\Phi}(x,t)+\bar{\rho}(x,t)\Delta_{x}\bar{\Phi}(x,t)\\ &=\nabla_{x}\bar{\rho}(x,t)\cdot\nabla_{x}\bar{\Phi}(x,t)-\bar{\rho}^{2}(x,t).\end{aligned} (4.3)

The property of convolution entails that

xΦ¯(x,t)=dxK(xy)ρ¯(y,t)𝑑y\displaystyle\nabla_{x}\bar{\Phi}(x,t)=\int_{\mathbb{R}^{d}}\nabla_{x}K(x-y)\bar{\rho}(y,t)\,dy

and

xK(xy)=xy|xy|d=et(x¯y¯)edt|x¯y¯|d=e(d1)tx¯y¯|x¯y¯|d=e(d1)tx¯K(x¯y¯)\displaystyle\nabla_{x}K(x-y)=\frac{x-y}{|x-y|^{d}}=\frac{e^{-t}(\bar{x}-\bar{y})}{e^{-dt}|\bar{x}-\bar{y}|^{d}}=e^{(d-1)t}\frac{\bar{x}-\bar{y}}{|\bar{x}-\bar{y}|^{d}}=e^{(d-1)t}\nabla_{\bar{x}}K(\bar{x}-\bar{y})

with y¯=ety\bar{y}=e^{t}y. Hence, we have

xΦ¯(x,t)=e(d1)tdx¯K(x¯y¯)n(y¯,t¯)𝑑y¯=e(d1)t(x¯Kn)(x¯,t¯).\displaystyle\nabla_{x}\bar{\Phi}(x,t)=e^{(d-1)t}\int_{\mathbb{R}^{d}}\nabla_{\bar{x}}K(\bar{x}-\bar{y})n(\bar{y},\bar{t})\,d\bar{y}=e^{(d-1)t}(\nabla_{\bar{x}}K\star n)(\bar{x},\bar{t}). (4.4)

Substituting (4.4) into (4.3), one has

x(ρ¯(x,t)xΦ¯(x,t))\displaystyle\nabla_{x}\cdot\left(\bar{\rho}(x,t)\nabla_{x}\bar{\Phi}(x,t)\right) =e(d+1)tx¯n(x¯,t¯)e(d1)t(x¯Kn)(x¯,t¯)e2dtn2(x¯,t¯)\displaystyle=e^{(d+1)t}\nabla_{\bar{x}}n(\bar{x},\bar{t})\cdot e^{(d-1)t}(\nabla_{\bar{x}}K\star n)(\bar{x},\bar{t})-e^{2dt}n^{2}(\bar{x},\bar{t})
=e2dt(x¯n(x¯,t¯)(x¯Kn)(x¯,t¯)n2(x¯,t¯))\displaystyle=e^{2dt}\left(\nabla_{\bar{x}}n(\bar{x},\bar{t})\cdot(\nabla_{\bar{x}}K\star n)(\bar{x},\bar{t})-n^{2}(\bar{x},\bar{t})\right)
=:e2dtx¯(n(x¯,t¯)x¯Ψ(x¯,t¯)),\displaystyle=:e^{2dt}\nabla_{\bar{x}}\cdot(n(\bar{x},\bar{t})\nabla_{\bar{x}}\Psi(\bar{x},\bar{t})),

where we denote Ψ(x¯,t¯):=(Kn)(x¯,t¯)\Psi(\bar{x},\bar{t}):=(K\star n)(\bar{x},\bar{t}). Substituting the above equalities into (4.2) and using the fact that e(d2)t=(2t¯+1)(d2)/2e^{(d-2)t}=(2\bar{t}+1)^{(d-2)/2}, then we obtain the equation for n(x¯,t¯)n(\bar{x},\bar{t}) as the following form, for simplicity, we still use the notation xx and tt,

tn=Δn+(2t+1)(d2)/2(nΨ),ΔΨ=n\partial_{t}n=\Delta n+(2t+1)^{(d-2)/2}\nabla\cdot(n\nabla\Psi),\quad-\Delta\Psi=n (4.5)

with initial data n0(x):=n(x,0)=ρ¯(x,0)=ρ¯0(x)n_{0}(x):=n(x,0)=\bar{\rho}(x,0)=\bar{\rho}_{0}(x). For the local-in-time existence and uniqueness of smooth solutions to (4.5), we refer to [14] where Riesz interaction potential is considered, but it can be easily extended to the Coulomb one. With (4.5) at hand, nn can therefore be represented by the following Duhamel integral equation

n(t)=etΔn0+0te(ts)Δ((2s+1)(d2)/2(n(s)Ψ(s)))𝑑sfor   allt(0,T)n(t)=e^{t\Delta}n_{0}+\int_{0}^{t}e^{(t-s)\Delta}\left((2s+1)^{(d-2)/2}\nabla\cdot\left(n(s)\nabla\Psi(s)\right)\right)ds\qquad\textrm{for\,\, all}\,\,t\in(0,T)

with Ψ=Kn\Psi=K\star n and KK defined by Remark 1.1. Here {etΔ}t0\{e^{t\Delta}\}_{t\geq 0} denotes the semigroup generated by the heat equation:

tn=Δn,xd.\partial_{t}n=\Delta n,\quad x\in\mathbb{R}^{d}.

In the proposition below, we provide some bound estimates for the equation (4.5). For this, we introduce weighted norms for a function f=f(x)f=f(x)

fLr:=esssupxd(1+|x|2)r/2f(x),fWrk,:=j=0kjfLr.\|f\|_{L^{\infty}_{r}}:=\operatorname*{ess\,sup}_{x\in\mathbb{R}^{d}}(1+|x|^{2})^{r/2}f(x),\quad\|f\|_{W^{k,\infty}_{r}}:=\sum_{j=0}^{k}\|\nabla^{j}f\|_{L^{\infty}_{r}}.

Lr(d)L^{\infty}_{r}(\mathbb{R}^{d}) and Wrk,(d)W^{k,\infty}_{r}(\mathbb{R}^{d}) are functions spaces with finite corresponding norms.

Proposition 4.1.

Let T>0T>0 and nn be a solution to the equation (4.5) on the time interval [0,T)[0,T) with sufficient regularity. Suppose that the initial data n0n_{0} satisfies

n0Wr3,(d)withr>d.n_{0}\in W^{3,\infty}_{r}(\mathbb{R}^{d})\quad\mbox{with}\quad r>d.

Then there exists T(0,T]T_{*}\in(0,T] such that

sup0tTn(,t)Wr3,<.\sup_{0\leq t\leq T_{*}}\|n(\cdot,t)\|_{W^{3,\infty}_{r}}<\infty.
Remark 4.1.

It is clear that Proposition 4.1 implies

sup0tTρ¯(,t)Wr3,<.\sup_{0\leq t\leq T_{*}}\|\bar{\rho}(\cdot,t)\|_{W^{3,\infty}_{r}}<\infty.

Furthermore, we deduce

ρ¯L(0,T;W3,1(d))\bar{\rho}\in L^{\infty}(0,T_{*};W^{3,1}(\mathbb{R}^{d}))

since

d|kρ¯|𝑑x\displaystyle\int_{\mathbb{R}^{d}}|\nabla^{k}\bar{\rho}|\,dx =d1(1+|x|2)r/2(1+|x|2)r/2|kρ¯(x)|𝑑x\displaystyle=\int_{\mathbb{R}^{d}}\frac{1}{(1+|x|^{2})^{r/2}}(1+|x|^{2})^{r/2}|\nabla^{k}\bar{\rho}(x)|\,dx
kρ¯Lrd1(1+|x|2)r/2𝑑x\displaystyle\leq\|\nabla^{k}\bar{\rho}\|_{L^{\infty}_{r}}\int_{\mathbb{R}^{d}}\frac{1}{(1+|x|^{2})^{r/2}}\,dx
<\displaystyle<\infty

due to r>dr>d, for k=0,1,2,3k=0,1,2,3. Hence, by combining this observation and the discussion in Section 3.4, under the assumption that

logρ¯0W3,(d)andρ¯0Wr3,(d)\log\bar{\rho}_{0}\in W^{3,\infty}(\mathbb{R}^{d})\quad\mbox{and}\quad\bar{\rho}_{0}\in W^{3,\infty}_{r}(\mathbb{R}^{d})

we have ρ¯L(0,T;W3,1W3,(d))\bar{\rho}\in L^{\infty}(0,T_{*};W^{3,1}\cap W^{3,\infty}(\mathbb{R}^{d})), which is the sufficient condition for the estimate appeared in Lemma 3.1 (see also Lemma 3.4).

For the rest of this section, we devote ourselves to prove Proposition 4.1. We first start with the Lp(d)L^{p}(\mathbb{R}^{d})-estimate of n(x,t)n(x,t). For any 1p<1\leq p<\infty, multiplying the first equation in (4.5) by pnp1pn^{p-1}, integrating the resulting equation with respect to xx, and using integration by parts, one has

ddtnLpp+4(p1)pnp2L22\displaystyle\frac{d}{dt}\|n\|_{L^{p}}^{p}+\frac{4(p-1)}{p}\|\nabla n^{\frac{p}{2}}\|_{L^{2}}^{2} =p(2t+1)(d2)/2dnp1(nΨ)𝑑x\displaystyle=p(2t+1)^{(d-2)/2}\int_{\mathbb{R}^{d}}n^{p-1}\nabla\cdot(n\nabla\Psi)\,dx
=p(p1)(2t+1)(d2)/2dnp1nΨdx\displaystyle=-p(p-1)(2t+1)^{(d-2)/2}\int_{\mathbb{R}^{d}}n^{p-1}\nabla n\cdot\nabla\Psi\,dx
=(p1)(2t+1)(d2)/2dnpΨdx\displaystyle=-(p-1)(2t+1)^{(d-2)/2}\int_{\mathbb{R}^{d}}\nabla n^{p}\cdot\nabla\Psi\,dx
=(p1)(2t+1)(d2)/2dnpΔΨ𝑑x\displaystyle=(p-1)(2t+1)^{(d-2)/2}\int_{\mathbb{R}^{d}}n^{p}\Delta\Psi\,dx
=(p1)(2t+1)(d2)/2dnp+1𝑑x0,\displaystyle=-(p-1)(2t+1)^{(d-2)/2}\int_{\mathbb{R}^{d}}n^{p+1}\,dx\leq 0,

which implies

sup0tTn(,t)Lpn0Lpfor  all  1p<\sup_{0\leq t\leq T}\|n(\cdot,t)\|_{L^{p}}\leq\|n_{0}\|_{L^{p}}\quad\textrm{for\,\,all}\,\,1\leq p<\infty (4.6)

and

sup0tTn(,t)Ln0L\sup_{0\leq t\leq T}\|n(\cdot,t)\|_{L^{\infty}}\leq\|n_{0}\|_{L^{\infty}} (4.7)

by letting pp\rightarrow\infty in (4.6).

With the above bound estimate at hand, we first show Wr1,(d)W^{1,\infty}_{r}(\mathbb{R}^{d})-estimate of nn in the following lemma.

Lemma 4.1.

Let T>0T>0 and nn be a solution to the equation (4.5) on the time interval [0,T)[0,T) with sufficient regularity. Assume n0Wr1,(d)n_{0}\in W^{1,\infty}_{r}(\mathbb{R}^{d}) with r>dr>d. Then there exists T>0T_{*}>0 such that the following estimates hold:

sup0tTn(,t)LrCandsup0tTn(,t)LrC,\displaystyle\sup_{0\leq t\leq T}\|n(\cdot,t)\|_{L^{\infty}_{r}}\leq C\quad\mbox{and}\quad\sup_{0\leq t\leq T_{*}}\|\nabla n(\cdot,t)\|_{L^{\infty}_{r}}\leq C,

where C>0C>0 only depends on d,rd,r, TT, and n0n_{0}.

Proof.

We first introduce simplified notations:

Y(x,t):=(1+|x|2)r/2n(x,t)andZ(x,t):=(1+|x|2)r/2n(x,t).Y(x,t):=(1+|x|^{2})^{r/2}n(x,t)\quad\mbox{and}\quad Z(x,t):=(1+|x|^{2})^{r/2}\nabla n(x,t).

In order to obtain the estimate YL\|Y\|_{L^{\infty}}, we multiply the first equation of (4.5) by (1+|x2|)r/2(1+|x^{2}|)^{r/2} to get

tY(2t+1)(d2)/2ΨYΔY=r(r2)(1+|x|2)(r4)/2|x|2nrd(1+|x|2)(r2)/2n2r(1+|x|2)(r2)/2xn+(2t+1)(d2)/2YΔΨr(2t+1)(d2)/2(1+|x|2)(r2)/2xnΨ.\displaystyle\begin{aligned} &\partial_{t}Y-(2t+1)^{(d-2)/2}\nabla\Psi\cdot\nabla Y-\Delta Y\cr&\quad=-r(r-2)(1+|x|^{2})^{(r-4)/2}|x|^{2}n-rd(1+|x|^{2})^{(r-2)/2}n-2r(1+|x|^{2})^{(r-2)/2}x\cdot\nabla n\\ &\qquad+(2t+1)^{(d-2)/2}Y\Delta\Psi-r(2t+1)^{(d-2)/2}(1+|x|^{2})^{(r-2)/2}xn\cdot\nabla\Psi.\end{aligned} (4.8)

Since we have

2r(1+|x|2)(r2)/2xn=2r1+|x|2xY+2r2(1+|x|2)(r4)/2|x|2n,-2r(1+|x|^{2})^{(r-2)/2}x\cdot\nabla n=-\frac{2r}{1+|x|^{2}}x\cdot\nabla Y+2r^{2}(1+|x|^{2})^{(r-4)/2}|x|^{2}n,

the equation (4.8) can thus be rewritten as

tY+(2rx1+|x|2(2t+1)(d2)/2Ψ)YΔY=R1+R2,\displaystyle\begin{aligned} &\partial_{t}Y+\left(2r\frac{x}{1+|x|^{2}}-(2t+1)^{(d-2)/2}\nabla\Psi\right)\cdot\nabla Y-\Delta Y=R_{1}+R_{2},\end{aligned}

where

R1:=(r(r+2)|x|21+|x|2rd)(1+|x|2)(r2)/2nR_{1}:=\left(r(r+2)\frac{|x|^{2}}{1+|x|^{2}}-rd\right)(1+|x|^{2})^{(r-2)/2}n

and

R2:=(2t+1)(d2)/2nYr(2t+1)(d2)/2(1+|x|2)(r2)/2xnΨ.R_{2}:=-(2t+1)^{(d-2)/2}nY-r(2t+1)^{(d-2)/2}(1+|x|^{2})^{(r-2)/2}xn\cdot\nabla\Psi.

For R1R_{1} and R2R_{2}, we have the following estimates:

R1LC(r,d)(1+|x|2)(r2)/2nLC(r,d)YL\|R_{1}\|_{L^{\infty}}\leq C(r,d)\|(1+|x|^{2})^{(r-2)/2}n\|_{L^{\infty}}\leq C(r,d)\|Y\|_{L^{\infty}}

and

R2LC(d,T)nLYL+C(r,d,T)ΨL(1+|x|2)(r1)/2nLC(d,T)n0LYL+C(r,d,T)nL1LYLC(r,d,T,n0)YL.\displaystyle\begin{aligned} \|R_{2}\|_{L^{\infty}}&\leq C(d,T)\|n\|_{L^{\infty}}\|Y\|_{L^{\infty}}+C(r,d,T)\|\nabla\Psi\|_{L^{\infty}}\|(1+|x|^{2})^{(r-1)/2}n\|_{L^{\infty}}\\ &\leq C(d,T)\|n_{0}\|_{L^{\infty}}\|Y\|_{L^{\infty}}+C(r,d,T)\|n\|_{L^{1}\cap L^{\infty}}\|Y\|_{L^{\infty}}\\ &\leq C(r,d,T,n_{0})\|Y\|_{L^{\infty}}.\end{aligned}

So, by a similar result in [17, Proposition A.3], we have the following LL^{\infty} estimate for YY

Y(,t)LY(,0)L+0t(R1(,s)L+R2(,s)L)𝑑sC1+C20tY(,s)L𝑑s.\displaystyle\|Y(\cdot,t)\|_{L^{\infty}}\leq\|Y(\cdot,0)\|_{L^{\infty}}+\int_{0}^{t}\left(\|R_{1}(\cdot,s)\|_{L^{\infty}}+\|R_{2}(\cdot,s)\|_{L^{\infty}}\right)ds\leq C_{1}+C_{2}\int_{0}^{t}\|Y(\cdot,s)\|_{L^{\infty}}\,ds.

By Grönwall’s inequality, we have

Y(,t)LC1(1+C2TeC2T)for a.e. 0tT.\displaystyle\|Y(\cdot,t)\|_{L^{\infty}}\leq C_{1}\left(1+C_{2}Te^{C_{2}T}\right)\qquad\textrm{for a.e. }0\leq t\leq T. (4.9)

Now, differentiating the first equation of (4.5) with respect to xx, we have

t(n)(2t+1)(d2)/2Ψ(n)Δ(n)=(2t+1)(d2)/22Ψn2(2t+1)(d2)/2nn.\displaystyle\begin{aligned} &\partial_{t}(\nabla n)-(2t+1)^{(d-2)/2}\nabla\Psi\cdot\nabla(\nabla n)-\Delta(\nabla n)\\ &\quad=(2t+1)^{(d-2)/2}\nabla^{2}\Psi\cdot\nabla n-2(2t+1)^{(d-2)/2}n\nabla n.\end{aligned} (4.10)

Then multiplying the obtained equation (4.10) by (1+|x|2)r/2(1+|x|^{2})^{r/2}, one has

tZ(2t+1)(d2)/2ΨZΔZ=r(2t+1)(d2)/2(1+|x|2)(r2)/2(xΨ)nr(r2)(1+|x|2)(r4)/2|x|2nrd(1+|x|2)(r2)/2n2r(1+|x|2)(r2)/2x(n)+(2t+1)(d2)/22ΨZ2(2t+1)(d2)/2nZ.\displaystyle\begin{aligned} &\partial_{t}Z-(2t+1)^{(d-2)/2}\nabla\Psi\cdot\nabla Z-\Delta Z\cr&\quad=-r(2t+1)^{(d-2)/2}(1+|x|^{2})^{(r-2)/2}(x\cdot\nabla\Psi)\nabla n-r(r-2)(1+|x|^{2})^{(r-4)/2}|x|^{2}\nabla n\\ &\qquad-rd(1+|x|^{2})^{(r-2)/2}\nabla n-2r(1+|x|^{2})^{(r-2)/2}x\cdot\nabla(\nabla n)+(2t+1)^{(d-2)/2}\nabla^{2}\Psi\cdot Z\\ &\qquad-2(2t+1)^{(d-2)/2}nZ.\end{aligned} (4.11)

Analogously, we have

2r(1+|x|2)(r2)/2x(n)=2r1+|x|2xZ+2r2(1+|x|2)(r4)/2|x|2n,-2r(1+|x|^{2})^{(r-2)/2}x\cdot\nabla(\nabla n)=-\frac{2r}{1+|x|^{2}}x\cdot\nabla Z+2r^{2}(1+|x|^{2})^{(r-4)/2}|x|^{2}\nabla n,

which together with (4.11) leads to

tZ+(2rx1+|x|2(2t+1)(d2)/2Ψ)ZΔZ=R3+R4,\displaystyle\partial_{t}Z+\left(2r\frac{x}{1+|x|^{2}}-(2t+1)^{(d-2)/2}\nabla\Psi\right)\cdot\nabla Z-\Delta Z=R_{3}+R_{4}, (4.12)

where

R3:=(r(r+2)|x|21+|x|2rd)(1+|x|2)(r2)/2nR_{3}:=\left(r(r+2)\frac{|x|^{2}}{1+|x|^{2}}-rd\right)(1+|x|^{2})^{(r-2)/2}\nabla n

and

R4\displaystyle R_{4} :=r(2t+1)(d2)/2(1+|x|2)(r2)/2(xΨ)n+(2t+1)(d2)/22ΨZ2(2t+1)(d2)/2nZ.\displaystyle:=-r(2t+1)^{(d-2)/2}(1+|x|^{2})^{(r-2)/2}(x\cdot\nabla\Psi)\nabla n+(2t+1)^{(d-2)/2}\nabla^{2}\Psi\cdot Z-2(2t+1)^{(d-2)/2}nZ.

Hence, we can easily deduce that

R3LC(r,d)(1+|x|2)(r2)/2nLC(r,d)ZL\displaystyle\|R_{3}\|_{L^{\infty}}\leq C(r,d)\|(1+|x|^{2})^{(r-2)/2}\nabla n\|_{L^{\infty}}\leq C(r,d)\|Z\|_{L^{\infty}} (4.13)

and

R4LC(r,d,T)ΨL(1+|x|2)(r1)/2nL+C(d,T)2ΨZL+C(d,T)nZLC(r,d,T)nL1LZL+C(d,T)nL1LZLC(r,d,T,n0)(ZL+ZL2).\displaystyle\begin{aligned} \|R_{4}\|_{L^{\infty}}&\leq C(r,d,T)\|\nabla\Psi\|_{L^{\infty}}\|(1+|x|^{2})^{(r-1)/2}\nabla n\|_{L^{\infty}}\\ &\quad+C(d,T)\|\nabla^{2}\Psi\cdot Z\|_{L^{\infty}}+C(d,T)\|nZ\|_{L^{\infty}}\\ &\leq C(r,d,T)\|n\|_{L^{1}\cap L^{\infty}}\|Z\|_{L^{\infty}}+C(d,T)\|\nabla n\|_{L^{1}\cap L^{\infty}}\|Z\|_{L^{\infty}}\\ &\leq C(r,d,T,n_{0})\left(\|Z\|_{L^{\infty}}+\|Z\|_{L^{\infty}}^{2}\right).\end{aligned} (4.14)

Here we used

kΨLC(r,d)k1nL1L,k\|\nabla^{k}\Psi\|_{L^{\infty}}\leq C(r,d)\|\nabla^{k-1}n\|_{L^{1}\cap L^{\infty}},\quad k\in\mathbb{N}

for some C>0C>0. Considering the equation (4.12) and using the above estimates (4.13) and (4.14), we obtain that

Z(,t)LZ(,0)L+0t(R3(,s)L+R4(,s)L)𝑑sC3+C40tZ(,s)L2𝑑s,\displaystyle\|Z(\cdot,t)\|_{L^{\infty}}\leq\|Z(\cdot,0)\|_{L^{\infty}}+\int_{0}^{t}\left(\|R_{3}(\cdot,s)\|_{L^{\infty}}+\|R_{4}(\cdot,s)\|_{L^{\infty}}\right)ds\leq C_{3}+C_{4}\int_{0}^{t}\|Z(\cdot,s)\|^{2}_{L^{\infty}}\,ds,

which further implies

Z(,t)LZ(,0)L1C4Z(,0)Ltfor a.e. 0t<1C4Z(,0)L.\displaystyle\|Z(\cdot,t)\|_{L^{\infty}}\leq\frac{\|Z(\cdot,0)\|_{L^{\infty}}}{1-C_{4}\|Z(\cdot,0)\|_{L^{\infty}}t}\qquad\textrm{for a.e. }0\leq t<\frac{1}{C_{4}\|Z(\cdot,0)\|_{L^{\infty}}}. (4.15)

The desired results can be easily concluded from (4.9) and (4.15). ∎

Next, we will devote ourselves to estimates for higher-order derivatives.

Lemma 4.2.

Let T>0T>0 and nn be a solution to the equation (4.5) on the time interval [0,T)[0,T) with sufficient regularity. Assume 2n0Wr1,(d)\nabla^{2}n_{0}\in W^{1,\infty}_{r}(\mathbb{R}^{d}) with r>dr>d. Then there exists T>0T_{*}>0 such that

sup0tT2n(,t)Wr1,C,\sup_{0\leq t\leq T_{*}}\|\nabla^{2}n(\cdot,t)\|_{W^{1,\infty}_{r}}\leq C,

where C>0C>0 only depends on d,rd,r, TT, and n0n_{0}.

Proof.

For simplicity of notation, we denote

G(x,t):=(1+|x|2)r/22n(x,t)andQ(x,t):=(1+|x|2)r/23n(x,t).G(x,t):=(1+|x|^{2})^{r/2}\nabla^{2}n(x,t)\quad\mbox{and}\quad Q(x,t):=(1+|x|^{2})^{r/2}\nabla^{3}n(x,t).

We apply \nabla to the equation (4.10) to deduce

t(2n)(2t+1)(d2)/2Ψ(2n)Δ(2n)=(2t+1)(d2)/23Ψn+2(2t+1)(d2)/22Ψ2n2(2t+1)(d2)/2nn2(2t+1)(d2)/2n2n.\displaystyle\begin{aligned} &\partial_{t}(\nabla^{2}n)-(2t+1)^{(d-2)/2}\nabla\Psi\cdot\nabla(\nabla^{2}n)-\Delta(\nabla^{2}n)\cr&\quad=(2t+1)^{(d-2)/2}\nabla^{3}\Psi\cdot\nabla n+2(2t+1)^{(d-2)/2}\nabla^{2}\Psi\cdot\nabla^{2}n\cr&\qquad-2(2t+1)^{(d-2)/2}\nabla n\nabla n-2(2t+1)^{(d-2)/2}n\nabla^{2}n.\end{aligned} (4.16)

Multiplying (4.16) by (1+|x|2)r/2(1+|x|^{2})^{r/2} and using the definition of G(x,t)G(x,t), we get

tG(2t+1)(d2)/2ΨGΔG=r(2t+1)(d2)/2Ψ(1+|x|2)(r2)/2x2nr(r2)(1+|x|2)(r4)/2|x|22nrd(1+|x|2)(r2)/22n2r(1+|x|2)(r2)/2x(2n)+(2t+1)(d2)/23ΨZ+2(2t+1)(d2)/22ΨG2(2t+1)(d2)/2Zn2(2t+1)(d2)/2nG.\displaystyle\begin{aligned} &\partial_{t}G-(2t+1)^{(d-2)/2}\nabla\Psi\cdot\nabla G-\Delta G\cr&\quad=-r(2t+1)^{(d-2)/2}\nabla\Psi\cdot(1+|x|^{2})^{(r-2)/2}x\nabla^{2}n\cr&\qquad-r(r-2)(1+|x|^{2})^{(r-4)/2}|x|^{2}\nabla^{2}n-rd(1+|x|^{2})^{(r-2)/2}\nabla^{2}n\cr&\qquad-2r(1+|x|^{2})^{(r-2)/2}x\cdot\nabla(\nabla^{2}n)+(2t+1)^{(d-2)/2}\nabla^{3}\Psi\cdot Z\cr&\qquad+2(2t+1)^{(d-2)/2}\nabla^{2}\Psi\cdot G-2(2t+1)^{(d-2)/2}Z\nabla n-2(2t+1)^{(d-2)/2}nG.\end{aligned} (4.17)

For the fourth term on the right-hand-side of (4.17), we further deduce that

2r(1+|x|2)(r2)/2x(2n)=2r1+|x|2xG+2r2(1+|x|2)(r4)/2|x|22n.\displaystyle-2r(1+|x|^{2})^{(r-2)/2}x\cdot\nabla(\nabla^{2}n)=-\frac{2r}{1+|x|^{2}}x\cdot\nabla G+2r^{2}(1+|x|^{2})^{(r-4)/2}|x|^{2}\nabla^{2}n. (4.18)

Substituting (4.18) into (4.17) leads to

tG+(2rx1+|x|2(2t+1)(d2)/2Ψ)GΔG=R5+R6,\displaystyle\partial_{t}G+\left(2r\frac{x}{1+|x|^{2}}-(2t+1)^{(d-2)/2}\nabla\Psi\right)\cdot\nabla G-\Delta G=R_{5}+R_{6}, (4.19)

where

R5:=(r(r+2)|x|21+|x|2rd)(1+|x|2)(r2)/22nR_{5}:=\left(r(r+2)\frac{|x|^{2}}{1+|x|^{2}}-rd\right)(1+|x|^{2})^{(r-2)/2}\nabla^{2}n

and

R6\displaystyle R_{6} :=r(2t+1)(d2)/2Ψ(1+|x|2)(r2)/2x2n+(2t+1)(d2)/23ΨZ\displaystyle:=-r(2t+1)^{(d-2)/2}\nabla\Psi\cdot(1+|x|^{2})^{(r-2)/2}x\nabla^{2}n+(2t+1)^{(d-2)/2}\nabla^{3}\Psi\cdot Z
+2(2t+1)(d2)/22ΨG2(2t+1)(d2)/2Zn2(2t+1)(d2)/2nG.\displaystyle\quad+2(2t+1)^{(d-2)/2}\nabla^{2}\Psi\cdot G-2(2t+1)^{(d-2)/2}Z\nabla n-2(2t+1)^{(d-2)/2}nG.

Similarly, we have that

R5LC(r,d)(1+|x|2)(r2)/22nLC(r,d)GL.\displaystyle\|R_{5}\|_{L^{\infty}}\leq C(r,d)\|(1+|x|^{2})^{(r-2)/2}\nabla^{2}n\|_{L^{\infty}}\leq C(r,d)\|G\|_{L^{\infty}}. (4.20)

By Lemma 4.1 and (4.7), one can bound R6R_{6} as

R6LC(r,d,T)ΨL(1+|x|2)(r1)/22nL+C(d,T)3ΨLZL+C(d,T)2ΨLGL+C(d,T)ZLnL+C(d,T)nLGLC(r,d,T)nL1LGL+C(d,T)2nL1LZL+C(d,T)nL1LGL+C(d,T)ZL2C(r,d,T)nL1LGL+C(d,T)GLZL+C(d,T)ZL2C(r,d,T,n0)+C(r,d,T,n0)GL\displaystyle\begin{aligned} \|R_{6}\|_{L^{\infty}}&\leq C(r,d,T)\|\nabla\Psi\|_{L^{\infty}}\|(1+|x|^{2})^{(r-1)/2}\nabla^{2}n\|_{L^{\infty}}+C(d,T)\|\nabla^{3}\Psi\|_{L^{\infty}}\|Z\|_{L^{\infty}}\cr&\quad+C(d,T)\|\nabla^{2}\Psi\|_{L^{\infty}}\|G\|_{L^{\infty}}+C(d,T)\|Z\|_{L^{\infty}}\|\nabla n\|_{L^{\infty}}+C(d,T)\|n\|_{L^{\infty}}\|G\|_{L^{\infty}}\cr&\leq C(r,d,T)\|n\|_{L^{1}\cap L^{\infty}}\|G\|_{L^{\infty}}+C(d,T)\|\nabla^{2}n\|_{L^{1}\cap L^{\infty}}\|Z\|_{L^{\infty}}\cr&\quad+C(d,T)\|\nabla n\|_{L^{1}\cap L^{\infty}}\|G\|_{L^{\infty}}+C(d,T)\|Z\|_{L^{\infty}}^{2}\cr&\leq C(r,d,T)\|n\|_{L^{1}\cap L^{\infty}}\|G\|_{L^{\infty}}+C(d,T)\|G\|_{L^{\infty}}\|Z\|_{L^{\infty}}+C(d,T)\|Z\|_{L^{\infty}}^{2}\cr&\leq C(r,d,T,n_{0})+C(r,d,T,n_{0})\|G\|_{L^{\infty}}\end{aligned} (4.21)

for tTt\leq T_{*}, where we used the boundedness of ZL\|Z\|_{L^{\infty}}. With (4.20) and (4.21) at hand, we can thus infer from the equation (4.19) that

G(,t)LG(,0)L+0t(R5(,s)L+R6(,s)L)𝑑sC5+C60tG(,s)L𝑑s,\displaystyle\|G(\cdot,t)\|_{L^{\infty}}\leq\|G(\cdot,0)\|_{L^{\infty}}+\int_{0}^{t}(\|R_{5}(\cdot,s)\|_{L^{\infty}}+\|R_{6}(\cdot,s)\|_{L^{\infty}})\,ds\leq C_{5}+C_{6}\int_{0}^{t}\|G(\cdot,s)\|_{L^{\infty}}ds,

which leads to

sup0tTG(,t)LC5(1+C6TeC6T).\sup_{0\leq t\leq T_{*}}\|G(\cdot,t)\|_{L^{\infty}}\leq C_{5}(1+C_{6}T_{*}e^{C_{6}T_{*}}).

Next, we estimate QL\|Q\|_{L^{\infty}}. To this end, we apply \nabla to (4.16) to obtain that

t(3n)(2t+1)(d2)/2Ψ(3n)Δ(3n)=3(2t+1)(d2)/22Ψ3n+(2t+1)(d2)/24Ψn+3(2t+1)(d2)/23Ψ2n6(2t+1)(d2)/2n2n2(2t+1)(d2)/2n3n.\displaystyle\begin{aligned} &\partial_{t}(\nabla^{3}n)-(2t+1)^{(d-2)/2}\nabla\Psi\cdot\nabla(\nabla^{3}n)-\Delta(\nabla^{3}n)\cr&\quad=3(2t+1)^{(d-2)/2}\nabla^{2}\Psi\cdot\nabla^{3}n+(2t+1)^{(d-2)/2}\nabla^{4}\Psi\cdot\nabla n+3(2t+1)^{(d-2)/2}\nabla^{3}\Psi\cdot\nabla^{2}n\cr&\qquad-6(2t+1)^{(d-2)/2}\nabla n\nabla^{2}n-2(2t+1)^{(d-2)/2}n\nabla^{3}n.\end{aligned} (4.22)

It follows from multiplying (4.22) by (1+|x|2)r/2(1+|x|^{2})^{r/2} that

tQ(2t+1)(d2)/2ΨQΔQ=r(2t+1)(d2)/2Ψ(1+|x|2)(r2)/2x3nr(r2)(1+|x|2)(r4)/2|x|23nrd(1+|x|2)(r2)/23n2r(1+|x|2)(r2)/2x(3n)+3(2t+1)(d2)/22ΨQ+(2t+1)(d2)/24ΨZ+3(2t+1)(d2)/23ΨG6(2t+1)(d2)/2nG2(2t+1)(d2)/2nQ.\displaystyle\begin{aligned} &\partial_{t}Q-(2t+1)^{(d-2)/2}\nabla\Psi\cdot\nabla Q-\Delta Q\cr&\quad=-r(2t+1)^{(d-2)/2}\nabla\Psi\cdot(1+|x|^{2})^{(r-2)/2}x\nabla^{3}n-r(r-2)(1+|x|^{2})^{(r-4)/2}|x|^{2}\nabla^{3}n\cr&\quad-rd(1+|x|^{2})^{(r-2)/2}\nabla^{3}n-2r(1+|x|^{2})^{(r-2)/2}x\cdot\nabla(\nabla^{3}n)+3(2t+1)^{(d-2)/2}\nabla^{2}\Psi\cdot Q\cr&\quad+(2t+1)^{(d-2)/2}\nabla^{4}\Psi\cdot Z+3(2t+1)^{(d-2)/2}\nabla^{3}\Psi\cdot G\cr&\quad-6(2t+1)^{(d-2)/2}\nabla n\,G-2(2t+1)^{(d-2)/2}nQ.\end{aligned} (4.23)

Similarly as before, the forth term on he right-hand-side of (4.23) can be rewritten as

2r(1+|x|2)(r2)/2x(3n)=2rx1+|x|2Q+2r2(1+|x|2)(r4)/2|x|23n.\displaystyle-2r(1+|x|^{2})^{(r-2)/2}x\cdot\nabla(\nabla^{3}n)=-2r\frac{x}{1+|x|^{2}}\cdot\nabla Q+2r^{2}(1+|x|^{2})^{(r-4)/2}|x|^{2}\nabla^{3}n. (4.24)

Substituting (4.24) into (4.22) and rearranging the resulting equality, one has

tQ+(2rx1+|x|2(2t+1)(d2)/2Ψ)QΔQ=R7+R8\displaystyle\partial_{t}Q+\left(2r\frac{x}{1+|x|^{2}}-(2t+1)^{(d-2)/2}\nabla\Psi\right)\cdot\nabla Q-\Delta Q=R_{7}+R_{8}

with

R7:=(r(r+2)|x|21+|x|2rd)(1+|x|2)(r2)/23n\displaystyle R_{7}:=\left(r(r+2)\frac{|x|^{2}}{1+|x|^{2}}-rd\right)(1+|x|^{2})^{(r-2)/2}\nabla^{3}n

and

R8:=r(2t+1)(d2)/2Ψ(1+|x|2)(r2)/2x3n+3(2t+1)(d2)/22ΨQ+(2t+1)(d2)/24ΨZ+3(2t+1)(d2)/23ΨG6(2t+1)(d2)/2nG2(2t+1)(d2)/2nQ.\displaystyle\begin{aligned} R_{8}&:=-r(2t+1)^{(d-2)/2}\nabla\Psi\cdot(1+|x|^{2})^{(r-2)/2}x\nabla^{3}n+3(2t+1)^{(d-2)/2}\nabla^{2}\Psi\cdot Q\cr&\quad+(2t+1)^{(d-2)/2}\nabla^{4}\Psi\cdot Z+3(2t+1)^{(d-2)/2}\nabla^{3}\Psi\cdot G\cr&\quad-6(2t+1)^{(d-2)/2}\nabla n\,G-2(2t+1)^{(d-2)/2}nQ.\end{aligned}

We can then bound R7R_{7} and R8R_{8} by using the obtained boundedness of YY, ZZ, GG and (4.7). Indeed, we have

R7LC(r,d)(1+|x|2)(r2)/23nLC(r,d)QL\displaystyle\|R_{7}\|_{L^{\infty}}\leq C(r,d)\|(1+|x|^{2})^{(r-2)/2}\nabla^{3}n\|_{L^{\infty}}\leq C(r,d)\|Q\|_{L^{\infty}}

and

R8LC(r,d,T)ΨL(1+|x|2)(r1)/23nL+C(d,T)2ΨLQL+C(d,T)4ΨLZL+C(d,T)3ΨLGL+C(d,T)nLGL+C(d,T)nLQLC(r,d,T)nL1LQL+C(d,T)nL1LQL+C(d,T)3nL1LZL+C(d,T)2nL1LGL+C(d,T)ZLGLC(r,d,T)nL1LQL+C(d,T)ZLQL+C(d,T)GL2+C(d,T)ZLGLC(r,d,T,n0)+C(r,d,T,n0)QL\displaystyle\begin{aligned} \|R_{8}\|_{L^{\infty}}&\leq C(r,d,T)\|\nabla\Psi\|_{L^{\infty}}\|(1+|x|^{2})^{(r-1)/2}\nabla^{3}n\|_{L^{\infty}}+C(d,T)\|\nabla^{2}\Psi\|_{L^{\infty}}\|Q\|_{L^{\infty}}\cr&\quad+C(d,T)\|\nabla^{4}\Psi\|_{L^{\infty}}\|Z\|_{L^{\infty}}+C(d,T)\|\nabla^{3}\Psi\|_{L^{\infty}}\|G\|_{L^{\infty}}\cr&\quad+C(d,T)\|\nabla n\|_{L^{\infty}}\|G\|_{L^{\infty}}+C(d,T)\|n\|_{L^{\infty}}\|Q\|_{L^{\infty}}\cr&\leq C(r,d,T)\|n\|_{L^{1}\cap L^{\infty}}\|Q\|_{L^{\infty}}+C(d,T)\|\nabla n\|_{L^{1}\cap L^{\infty}}\|Q\|_{L^{\infty}}\cr&\quad+C(d,T)\|\nabla^{3}n\|_{L^{1}\cap L^{\infty}}\|Z\|_{L^{\infty}}+C(d,T)\|\nabla^{2}n\|_{L^{1}\cap L^{\infty}}\|G\|_{L^{\infty}}+C(d,T)\|Z\|_{L^{\infty}}\|G\|_{L^{\infty}}\cr&\leq C(r,d,T)\|n\|_{L^{1}\cap L^{\infty}}\|Q\|_{L^{\infty}}+C(d,T)\|Z\|_{L^{\infty}}\|Q\|_{L^{\infty}}\cr&\quad+C(d,T)\|G\|^{2}_{L^{\infty}}+C(d,T)\|Z\|_{L^{\infty}}\|G\|_{L^{\infty}}\cr&\leq C(r,d,T,n_{0})+C(r,d,T,n_{0})\|Q\|_{L^{\infty}}\end{aligned}

for tTt\leq T_{*}, where we again used the boundedness of Z(,t)L\|Z(\cdot,t)\|_{L^{\infty}} for all tTt\leq T_{*}. These estimates yield

Q(,t)LQ(,0)L+0t(R7(,s)L+R8(,s)L)𝑑sC7+C80tQ(,s)L𝑑s,\displaystyle\|Q(\cdot,t)\|_{L^{\infty}}\leq\|Q(\cdot,0)\|_{L^{\infty}}+\int_{0}^{t}(\|R_{7}(\cdot,s)\|_{L^{\infty}}+\|R_{8}(\cdot,s)\|_{L^{\infty}})\,ds\leq C_{7}+C_{8}\int_{0}^{t}\|Q(\cdot,s)\|_{L^{\infty}}\,ds,

which implies that

sup0tTQ(,t)LC7(1+C8TeC8T).\sup_{0\leq t\leq T_{*}}\|Q(\cdot,t)\|_{L^{\infty}}\leq C_{7}(1+C_{8}T_{*}e^{C_{8}T_{*}}).

This completes the proof. ∎

Proof of Proposition 4.1.

The proof follows from a simple combination of Lemmas 4.1 and 4.2. ∎

Acknowledgments

The research of JAC was supported by the Advanced Grant Nonlocal-CPD (Nonlocal PDEs for Complex Particle Dynamics: Phase Transitions, Patterns and Synchronization) of the European Research Council Executive Agency (ERC) under the European Union’s Horizon 2020 research and innovation programme (grant agreement No. 883363) and also partially supported by EPSRC grant number EP/T022132/1. The work of YPC is supported by NRF grant (No. 2017R1C1B2012918), POSCO Science Fellowship of POSCO TJ Park Foundation, and Yonsei University Research Fund of 2020-22-0505. YP is partially supported by the Applied Fundamental Research Program of Sichuan Province (No. 2020YJ0264).

References

  • [1] L. Ambrosio, N. Gigli, and G. Savaré, Gradient flows in metric spaces and in the space of probability measures, Lectures in Mathematics ETH Zürich. Birkhäuser Verlag, Basel, 2005.
  • [2] F. Bouchut, Existence and uniqueness of a global smooth solution for the Vlasov–Poisson–Fokker–Planck system in three dimensions, J. Funct. Anal., 111, (1993), 239–258.
  • [3] J. A. Carrillo and Y.-P. Choi, Quantitative error estimates for the large friction limit of Vlasov equation with nonlocal forces, Ann. Inst. H. Poincaré Anal. Non Linéaire, 37, (2020), 925–954.
  • [4] J. A. Carrillo, Y.-P. Choi, and J. Jung, Quantifying the hydrodynamic limit of Vlasov-type equations with alignment and nonlocal forces, Math. Models Methods Appl. Sci., 31, (2021), 327–408.
  • [5] J. A. Carrillo, Y.-P. Choi, and O. Tse, Convergence to equilibrium in Wasserstein distance for damped Euler equations with interaction forces, Commun. Math. Phys, 365, (2019), 329–361.
  • [6] J. A. Carrillo, K. Craig, and Y. Yao, Aggregation-diffusion equations: dynamics, asymptotics, and singular limits, in N. Bellomo, P. Degond, and E. Tadmor (Eds.), Active Particles Vol. II: Advances in Theory, Models, and Applications, Series: Modelling and Simulation in Science and Technology, Birkhäuser Basel, 65–108, 2019.
  • [7] J. A. Carrillo, E. Feireisl, P. Gwiazda and A. Świerczewska-Gwiazda, Weak solutions for Euler systems with non-local interactions, J. London Math. Soc., 95, (2017), 705–724.
  • [8] J. A. Carrillo, R. J. McCann, C. Villani, Kinetic equilibration rates for granular media and related equations: entropy dissipation and mass transportation estimates, Rev. Mat. Iberoamericana, 19, (2003), 971–1018.
  • [9] J. A. Carrillo, Y. Peng, and A. Wróblewska-Kamińska, Relative entropy method for the relaxation limit of hydrodynamic models, Netw. Heterog. Media, 15, (2020), 369–387.
  • [10] J. A. Carrillo and J. Soler, On the initial value problem for the Vlasov–Poisson–Fokker–Planck system with initial data in LpL^{p} spaces, Math. Methods Appl. Sci., 18, (1995), 825–839.
  • [11] J. A. Carrillo and G. Toscani, Exponential convergence toward equilibrium for homogeneous Fokker–Planck-type equations, Math. Methods Appl. Sci., 21, (1998), 1269–1286.
  • [12] Y.-P. Choi, Global classical solutions of the Vlasov–Fokker–Planck equation with local alignment forces, Nonlinearity, 29, (2016), 1887–1916.
  • [13] Y.-P. Choi, Large friction limit of pressureless Euler equations with nonlocal forces, preprint, arXiv:2002.01691.
  • [14] Y.-P. Choi and I.-J. Jeong, Classical solutions to the fractional porous medium flow, preprint, arXiv:2102.01816.
  • [15] Y.-P. Choi and O. Tse, Quantified overdamped limit for kinetic Vlasov–Fokker–Planck equations with singular interaction forces, preprint, arXiv:2012.00422.
  • [16] J. F. Coulombel and T. Goudon, The strong relaxation limit of the multidimensional isothermal Euler equations, Trans. Amer. Math. Soc., 359, (2007), 637–648.
  • [17] P. Degond, Global existence of smooth solutions for the Vlasov–Fokker–Planck equation in 1 and 2 space dimensions, Ann Sci. École Norm. Sup., 19, (1986), 519–542.
  • [18] M. H. Duong, A. Lamacz, M. A. Peletier, A. Schlichting, and U. Sharma, Quantification of coarse-graining error in Langevin and overdamped Langevin dynamics, Nonlinearity, 31, (2018), 4517–4566.
  • [19] M. H. Duong, A. Lamacz, M. A. Peletier, and U. Sharma, Variational approach to coarse-graining of generalized gradient flows, Calc. Var. Partial Differential Equations, 56, (2017), 100.
  • [20] R. Fetecau and W. Sun, First-order aggregation models and zero inertia limits, J. Differential Equations, 259, (2015), 6774–6802.
  • [21] T. Goudon, J. Nieto, F. Poupaud, and J. Soler, Multidimensional high-field limit of the electrostatic Vlasov–Poisson–Fokker–Planck system, J. Differential Equations, 213, (2005), 418–442.
  • [22] P.-E. Jabin, Macroscopic limit of Vlasov type equations with friction, Ann. Inst. H. Poincaré Anal. Non Linéaire, 17, (2000), 651–672.
  • [23] R. Jordan, D. Kinderlehrer, and F. Otto, The variational formulation of the Fokker–Planck equation, SIAM J. Math. Anal., 29, (1998), 1–17.
  • [24] T. Karper, A. Mellet and K. Trivisa, Existence of weak solutions to kinetic flocking models, SIAM Math. Anal. 45, (2013), 215–243.
  • [25] T. Karper, A. Mellet and K. Trivisa, On strong local alignment in the kinetic Cucker–Smale model, Hyperbolic conservation laws and related analysis with applications, 227–242, Springer Proc. Math. Stat., 49, Springer, Heidelberg, 2014.
  • [26] T. K. Karper, A. Mellet, and K. Trivisa, Hydrodynamic limit of the kinetic Cucker–Smale flocking model, Math. Models Methods Appl. Sci., 25, (2015), 131–163.
  • [27] C. Lattanzio, A. E. Tzavaras, Relative entropy in diffusive relaxation, SIAM J. Math. Anal., 45, (2013), 1563–1584.
  • [28] C. Lattanzio, A. E. Tzavaras, From gas dynamics with large friction to gradient flows describing diffusion theories, Comm. Partial Differential Equations, 42, (2017), 261–290.
  • [29] J. Nieto, F. Poupaud, and J. Soler, High-field limit for the Vlasov–Poisson–Fokker–Planck system, Arch. Ration. Mech. Anal., 158, (2001), 29–59.
  • [30] F. Poupaud and J. Soler, Parabolic limit and stability of the Vlasov–Poisson–Fokker–Planck system, Math. Models Methods Appl. Sci., 10, (2000), 1027–1045.
  • [31] H. D. Victory, Jr, On the existence of global weak solutions for Vlasov–Poisson–Fokker–Planck systems, J. Math. Anal. Appl., 160, (1991), 525–555.
  • [32] H. D. Victory, Jr. and B. P. O’Dwyer, On classical solutions of Vlasov–Poisson–Fokker–Planck systems, Indiana Univ. Math. J., 39, (1990), 105–156.
  • [33] C. Villani, A review of mathematical topics in collisional kinetic theory Handbook of Mathematical Fluid Dynamics vol I (Amsterdam: North-Holland), 2002, pp 71–305.