Processing math: 64%

Hydrodynamic limit for a Fokker-Planck equation with coefficients in Sobolev spaces

  • Received: 01 July 2016 Revised: 01 February 2017
  • Primary: 35Q70, 35Q84; Secondary: 35Q35

  • In this paper we study the hydrodynamic (small mass approximation) limit of a Fokker-Planck equation. This equation arises in the kinetic description of the evolution of a particle system immersed in a viscous Stokes flow. We discuss two different methods of hydrodynamic convergence. The first method works with initial data in a weighted L2 space and uses weak convergence and the extraction of convergent subsequences. The second uses entropic initial data and gives an L1 convergence to the solution of the limit problem via the study of the relative entropy.

    Citation: Ioannis Markou. Hydrodynamic limit for a Fokker-Planck equation with coefficients in Sobolev spaces[J]. Networks and Heterogeneous Media, 2017, 12(4): 683-705. doi: 10.3934/nhm.2017028

    Related Papers:

    [1] Ioannis Markou . Hydrodynamic limit for a Fokker-Planck equation with coefficients in Sobolev spaces. Networks and Heterogeneous Media, 2017, 12(4): 683-705. doi: 10.3934/nhm.2017028
    [2] Karoline Disser, Matthias Liero . On gradient structures for Markov chains and the passage to Wasserstein gradient flows. Networks and Heterogeneous Media, 2015, 10(2): 233-253. doi: 10.3934/nhm.2015.10.233
    [3] José Antonio Carrillo, Yingping Peng, Aneta Wróblewska-Kamińska . Relative entropy method for the relaxation limit of hydrodynamic models. Networks and Heterogeneous Media, 2020, 15(3): 369-387. doi: 10.3934/nhm.2020023
    [4] Michael Herty, Lorenzo Pareschi, Sonja Steffensen . Mean--field control and Riccati equations. Networks and Heterogeneous Media, 2015, 10(3): 699-715. doi: 10.3934/nhm.2015.10.699
    [5] Diandian Huang, Xin Huang, Tingting Qin, Yongtao Zhou . A transformed L1 Legendre-Galerkin spectral method for time fractional Fokker-Planck equations. Networks and Heterogeneous Media, 2023, 18(2): 799-812. doi: 10.3934/nhm.2023034
    [6] Hirotada Honda . On Kuramoto-Sakaguchi-type Fokker-Planck equation with delay. Networks and Heterogeneous Media, 2024, 19(1): 1-23. doi: 10.3934/nhm.2024001
    [7] Martino Bardi . Explicit solutions of some linear-quadratic mean field games. Networks and Heterogeneous Media, 2012, 7(2): 243-261. doi: 10.3934/nhm.2012.7.243
    [8] Luis Almeida, Federica Bubba, Benoît Perthame, Camille Pouchol . Energy and implicit discretization of the Fokker-Planck and Keller-Segel type equations. Networks and Heterogeneous Media, 2019, 14(1): 23-41. doi: 10.3934/nhm.2019002
    [9] Yves Achdou, Victor Perez . Iterative strategies for solving linearized discrete mean field games systems. Networks and Heterogeneous Media, 2012, 7(2): 197-217. doi: 10.3934/nhm.2012.7.197
    [10] Shui-Nee Chow, Xiaojing Ye, Hongyuan Zha, Haomin Zhou . Influence prediction for continuous-time information propagation on networks. Networks and Heterogeneous Media, 2018, 13(4): 567-583. doi: 10.3934/nhm.2018026
  • In this paper we study the hydrodynamic (small mass approximation) limit of a Fokker-Planck equation. This equation arises in the kinetic description of the evolution of a particle system immersed in a viscous Stokes flow. We discuss two different methods of hydrodynamic convergence. The first method works with initial data in a weighted L2 space and uses weak convergence and the extraction of convergent subsequences. The second uses entropic initial data and gives an L1 convergence to the solution of the limit problem via the study of the relative entropy.



    We study the hydrodynamic limit for a Fokker-Planck equation that arises in the modeling of a system of large particles immersed in a much larger number of micromolecules. Examples of such particle systems include dilute solutions of polymers that arise often in industrial settings [2,7,8,21,24]. Typically, macromolecules (or more precisely, the monomer parts they are comprised of) are modeled by ideal spheres whose interactions are mediated by interactions with the micromolecules. We model the micromolecules as an incompressible fluid governed by Stokes flow. The interactions of these idealized particles with the fluid are modeled by admissible boundary conditions, Brownian noise and the introduction of damping.

    The dynamics of particle motion is described by a phase-space vector (x,v)Rnx×Rnv. If the statistics of the particle motion is described by the probability density f(t,x,v)0, then the evolution of f is governed by the Fokker-Planck equation

    tf+vxf+1mv(Ff)=1m2v(G(x)vf), (1)

    with the particle mass represented by m. The force F(t,x,v) in our model is chosen so that F(t,x,v)=V(x)G(x)v, where V(x) is a potential that depends on the particles' configuration and G(x)v is the damping (hydrodynamic) force term. The potential V(x) captures all interactions between beads that are not mediated by the fluid. This allows, particularly, for the incorporation in the model of any type of spring forces between beads.

    The Fokker-Planck equation (1) is naturally associated to the phase-space stochastic differential system

    {˙x(t)=v,˙v(t)=1m(G(x)v+V(x))+2mG1/2(x)˙W(t),

    where W(t) is the centered Gaussian vector in Rn with covariance E(˙W(t)˙W(t))=Iδ(tt). Here E stands for expectation with respect to Gaussian measure. The inclusion of the friction matrix G(x) in the Brownian forcing term is an instance of the fluctuation-dissipation theorem which asserts that fluctuations caused by white noise and the response to small perturbations applied to the system are in balance. This is evident by the Einstein-Smoluchowski relation [10,32,35] that states that the diffusion tensor (related to thermal motion) is proportional to friction G(x). The study of the limit m0 in the stochastic system is the celebrated Smoluchowski-Kramers approximation (see e.g. [11]).

    Equation (1) is very important in the description of polymer models when inertial effects are involved. This is reminiscent of the inertial kinetic models in the work by P. Degond and H. Liu [6]. Therein, the authors introduce novel kinetic models for Dumbell-like and rigid-rod polymers in the presence of inertial forces and show formally that when inertial effects vanish the limit is consistent with well accepted macroscopic models in polymer rheology. A direct quote from [6] reasons on the importance of kinetic models involving inertial effects in describing polymer sedimentation: ''In current kinetic theory models for polymers, the inertia of molecules is often neglected. However, neglect of inertia in some cases leads to incorrect predictions of the behavior of polymers. The forgoing considerations indicate that the inertial effects are of importance in practical applications, e.g., for short time characteristics of materials based on the relevant underlying phenomena''.

    One of the differences with the theory in the Degond & Liu work is that we take into account hydrodynamic interactions between N particles, with the use of the symmetric, non negative, 3N×3N friction tensor G(x) that contains all the information for these interactions. These hydrodynamic interactions are the result of a particle's motion that perturbs the fluid and has an effect on other particles' movement. The constant friction case G(x)=γI (for γ>0) is interesting in its own right as it corresponds to particles that ''sink freely'' without any hydrodynamic type of interaction between them. In this trivial case, there is no account of hydrodynamic effects and the parabolic limit is derived with no difficulty as we show. In a similar spirit as in [6], our goal is to show rigorously that equation (1) leads to the derivation of a well accepted Smoluchowski type of equation when inertial effects are ignored (see Theorem 1.1).

    Before we proceed with the details of the limiting approximation, we should note the difficulties in computing the exact formula for friction G(x) (or most commonly the mobility μ=G1(x)) for every N particle configuration. In practice, this would involve solving a linear Stokes system with very complicated boundary conditions, i.e., the N particles' surface. A particular modelling problem is the appropriate way to compute these interactions for overlapping particles and particles that are almost touching. More specific, for particles that are very close, integrable singularities of the friction tensor are possible (lubrication effects). Below we give the two most important approximations of the mobility tensor used in simulations.

    The first non-trivial approximation to mobility is the Oseen tensor that corresponds to Green's kernel solution of a Stokes problem for point particles [8,23]. For N particles with centers {xi}Ni=1, radius a, in a fluid with viscosity η, the Oseen tensor μOS=[μOSij]Ni,j=1 is a 3N×3N tensor with 3×3 blocks

    μOSij={18πη|Rij|(I+ˆRijˆRij),ij16πηaI,i=j,

    where Rij=xixj and ˆRij=Rij/|Rij|. This approximation works quite well when particles are well separated (|Rij|a), but it is degenerate for particle configurations that involve particles relatively close. This implies that the Oseen tensor cannot be a meaningful choice that leads to a well-posed theory (in the sense of existence, uniqueness, macroscopic limit, …).

    The Rotne-Prager-Yamakawa approximation of the mobility tensor [33,36] is a nonnegative correction to the Oseen tensor that applies to all particle configurations. In addition, Rotne and Prager [33] obtained a way to calculate mobilities for overlapping spheres. The expression for the RPY 3N×3N mobility has blocks

    μRPYij={18πη|Rij|[(1+2a23|Rij|2)I+(12a2|Rij|2)ˆRijˆRij],|Rij|>2a16πηa[(19|Rij|32a)I+3|Rij|32aˆRijˆRij],|Rij|2a.

    Eigenvalues of the tensor depend continuously on the particles' positions, they are bounded and the RPY mobility is locally integrable in space. On the other hand, the tensor is still not strictly positive. In more detail, when two spheres (of radius a) almost coincide and their centers have distance d=|Rij|a, then the minimum eigenvalues λmin(x1,x2) of RPY are of order O(d). This in turn implies that the friction associated to the RPY tensor is of order O(1d) and hence gives an integrable singularity.

    We should note that the exact computation of the eigenvalues of the RPY mobility for N>2 is impossible and the problem of directly obtaining the best lower bounds for λmin(x1,,xN) is still open. On the other hand, the additive nature of hydrodynamic interactions suggests a bound from below that is linear with respect to particle distances. For instance, for N particles in a configuration with all interparticle distances equal to d=|Rij|ai,j, the minimum eigenvalues can be computed exactly and are once again of order O(d). Moreover, for two nearly touching spheres with dimensionless gap parameter ξ=|Rij|a2 (with ξ1), lubrication theory suggests that the leading order of the friction tensor is O(1ξ) [3,22,23,31].

    We now turn our focus to the study of the diffusion limit for kinetic equation (1) which begins by introducing the appropriate scaling to separate conservative and dissipative terms. We repeat the scaling procedure in [6] that involves the change of variables,

    m=ϵ2,v=ϵv,x=x.

    Thus, (1) becomes (after we re-introduce the notation for x, v in the place of x, v and set initial conditions) the Cauchy problem

    tfϵ+Lϵfϵ=0,fϵ(0,x,v)=f0,ϵ(x,v),withLϵ=1ϵ(vxfϵV(x)vfϵ)1ϵ2v(G(x)(vfϵ+vfϵ)). (2)

    Our main objective is to study the (zero mass) limit ϵ0, for both fϵ and the hydrodynamical density ρϵ:=fϵdv, where integration is assumed everywhere over Rnv (or Rnx×Rnv when spatial variables are also involved). The second term of Lϵ in (2) is responsible for the system approaching local equilibrium Gibbs states ρM(v), with M(v) being the standard Maxwellian distribution

    M(v)=e|v|22/(2π)n2 (3)

    and ρ the limit of ρϵ.

    We give two results of convergence which are discussed in Section 1.2. First, we show in Theorem 1.1 that ρϵ converges weakly to ρ and that the limiting distribution satisfies the Smoluchowski equation tρ=(G1(ρ+ρV(x))). The convergence is proven under very mild assumptions on the hydrodynamic tensor, i.e., local integrability for the friction and mobility. This observation implies that the RPY tensor as well as any other physically meaningful nonnegative choice of mobility satisfies the assumptions of the first result. We then ask the following question: Is it possible to achieve a stronger convergence, say in an L1 setting? We give a definite answer in Theorem 1.2. The result that we prove is a more theoretical one, in that it does not apply to the realistic examples of mobilities mentioned earlier, but it is mathematically interesting in its own right. We work with entropic initial data that are ''well prepared'' in the sense that the tails are accommodated by a Maxwellian, and we show that more stringent control on the hydrodynamic mobility is required (typically G1(W3,(Rnx))n×n). The analysis uses a relative entropy functional and solutions that are renormalized in the spirit of Le Bris & Lions [27]. As a result, a rigorous justification of the computations is completed by a regularization process explained in detail.

    We mention that similar macroscopic limits in the parabolic scaling regime have been considered by many authors in the past, and for various collision operators that lie in the fast scale ϵ2. A discussion of the literature cannot, by any means, be inclusive. We only outline here some works that are relevant [1,5,9,12,14,29,30]. For instance, in [5] this limit is considered for the linear Boltzmann equation with a collision operator of the form σ(x,v,ω)f(ω)dμ(ω)fσ(x,v,ω)dμ(ω), for a σ-finite measure dμ(ω), and under the assumption that there exists a unique stationary state F(x,v) for which

    F(x,v)σ(x,v,ω)dμ(ω)=σ(x,v,ω)F(x,ω)dμ(ω)a.e.

    The collision kernel σ(x,v,ω) is assumed measurable with σ(x,v,ω)dμ(ω)< and it does not satisfy the micro-reversibility condition σ(x,v,ω)F(x,ω)=σ(x,ω,v)F(x,v). Such models are prominent in the theory of plasmas, semiconductors, rarefied gases etc. In [30] the authors study the parabolic limit for the nonlinear Boltzmann operator M(v)(1f)fdvfM(v)(1f)dv. This operator appears in the study of semiconductors, where f(t,x,v) is the fraction of occupied states (occupancy number). The operator leads to relaxation to the Fermi-Dirac distribution fFD(μ,v)=(1+e(12|v|2μ))1, where μ is the Fermi energy that depends implicitly on ρ(μ)=fFD(μ,v)dv. When the limit is considered for ϵ=τL0 (mean free path τ is small compared to characteristic length scale L) then ffFD and the Fermi energy μ satisfies the diffusive equation tρ(μ)=x(D(μ)x(μV)), for a diffusive coefficient D(μ) with an explicit structure. The electrostatic potential V(t,x) appears in the transport term vxfxVvf, which is in scale ϵ1. The Rosseland approximation for the radiative transfer equation has been studied in [1]. Equations that lead to nonlinear diffusions in the limit have been considered in [9].

    Finally, we mention that the particle system described here without the inclusion of Brownian motion has also been studied in relevant works (see e.g. [19,20]). In this present work, the derivation of a convection-diffusion limit is carried out for a linear Fokker-Planck equation with dominating friction and Brownian forcing terms governed by an anisotropic tensor G(x). The equation is of particular importance in the theory of particles moving in Stokes flows. The limiting Smoluchowski equation that we derive is the cornerstone of the kinetic theory of polymer chains in dilute solutions [7,8,24].

    We now bring our attention to the two main results of hydrodynamic convergence. In both of the results we are about to give, we assume that the solution to equation (2) is weak (in the sense that will be explained in Section 2) thus allowing for quite irregular coefficients. We make two assumptions. First, we assume a non-degenerate, nonnegative definite friction such that G1(x) exists a.e. and we also assume that eV(x)L1(Rnx). These assumptions suggest that there exists a unique global equilibrium state explicitly given by

    Meq(x,v)=eV(x)M(v)/Z,withZ=(2π)n2eV(x)dx. (4)

    We also consider V(x) bounded from below in the sense that infV(x)>.

    In the first theorem, we establish weak convergence of the hydrodynamic variable ρϵ(t,x) based on weak compactness arguments. The proof is actually quite straightforward. We assume a solution of (2) in the mild-weak sense. Such a solution fϵ lives in C(R+,D(Rnx×Rnv)). We also make the assumption that the initial data are in the weighted L2Meq space, where L2Meq=MeqL2(Meqdvdx)=MeqL2(dμ) (for a measure μ with density Meq), i.e.,

    fϵ(0,x,v)L2Meq<Cϵ>0,for someC>0. (5)

    We prove in Section 3 the following theorem.

    Theorem 1.1. Let fϵ be a mild-weak solution to (2) with bounded initial energy fϵ(0,,)L2Meq< (uniformly in ϵ>0), and let ρϵ be the hydrodynamical density ρϵ:=fϵdv. Assume that the non-degenerate a.e. friction tensor G(x) and potential V(x) satisfy conditions : G1(x)&G(x)(L1loc(Rnx))n×n, V(x)(L2loc(Rnx))n, G1/2V(x)(L2loc(Rnx))n and eV(x)L1(Rnx). In the limit ϵ0, we have the convergence

    ρϵρinC([0,T],wL2(dx)),

    where ρ is the solution to the Smoluchowski equation

    tρ=x(G1(xρ+V(x)ρ))inC([0,T],D(Rnx)). (6)

    In the second theorem, we use the relative entropy functional to prove an L1 convergence result. The relative entropy H(f|g) between two densities f,g is defined by

    H(f|g)=flogfgdvdx, (7)

    and in the present work it will be used to control the distance of a solution fϵ of (2) from the local Gibbs state ρM(v) as ϵ0.

    The relative entropy has been used in the study of many asymptotic problems. The earliest example appears to be in the study of the hydrodynamic limit for the Ginzburg-Landau problem in [37]. In [34], the author takes a probabilistic approach to the use of relative entropy. Other more elaborate cases include the Vlasov-Navier-Stokes system [17], hydrodynamic limits for the Boltzmann equation [13].

    To prove Theorem 1.2, we make the following assumptions. First, we need conditions that give control of the hydrodynamical tensor G1(x) and potential V(x), i.e.,

    kG1L(Rnx)<,k(G1V(x))L(Rnx)<,1k3. (A1)

    We also assume that the initial condition ρ(0,x) to equation (6) satisfies

    aeV(x)ρ(0,x)AeV(x)for someA>a>0andρ(0,x)/eV(x)W3,(Rnx).} (A2)

    Finally, the use of the maximum principle for the parabolic equation (6) in Rnx requires certain admissibility conditions at infinity. We can choose for instance the following condition for a given T>0,

    sup0tTlim supx|kρ(t,x)eV(x)|CkforCk>0,0k3, (A3)

    where || is the Hilbert-Schmidt norm of the tensor. In Section 4 we prove

    Theorem 1.2. Let fϵ(0,x,v) be initial data to the F-P equation (2) such that fϵ(0,x,v)0, satisfying the energy bound

    supϵ>0fϵ(0,x,v)(1+V(x)+|v|2+logfϵ(0,x,v))dvdx<C<. (8)

    Moreover, we assume that eV(x)L1(Rnx) and that the hydrodynamic tensor G1(x) and potential V(x) satisfy condition (A1). Let ρ(0,x)D(Rnx) be initial data to the limit equation (6), satisfying

    ρ(0,x)dx=fϵ(0,x,v)dvdx=1,

    as well as condition (A2). We finally make the assumption that the initial data are prepared so that

    H(fϵ(0,,)|ρ(0,)M(v))0asϵ0.

    Then, for any T>0, if ρ(t,x)C([0,T],D(Rnx)) is a solution to the limit equation that satisfies (A3), we have

    sup0tTH(fϵ(t,,)|ρ(t,)M(v))0asϵ0.

    The rest of the paper is organized as follows. In the next section, we give a formal derivation of the macroscopic limit and present the main steps in the proof of the two theorems mentioned above. We also give an exact description of the type of solutions we assume for problem (2) in each theorem. Sections 3 & 4 are devoted to the proof of each theorem with all the a priori estimates.

    We begin by writing the collision operator in form

    v(G(x)(vfϵ+vfϵ))=v(M(v)G(x)v(fϵM(v))).

    This form is indicative of why the collision part of Lϵ is responsible for the dissipation of energies. Let us now introduce the hydrodynamical variables for the density ρϵ, the flux vector Jϵ, and the kinetic pressure tensor Pϵ of the particle system, i.e.,

    ρϵ(t,x):=fϵdv,Jϵ(t,x):=vfϵdv,Pϵ(t,x):=vvfϵdv. (9)

    In the study of the limit ϵ0, we want to derive an equation for the hydrodynamic variable ρ(t,x) which is formally the limit of ρϵ.

    First, integrating (2) in velocity space, we obtain

    tρϵ+1ϵxJϵ=0. (10)

    We want to derive an expression for the evolution of Jϵ and study the order of magnitude in ϵ of the terms involved in it. In the derivation of the equation for the first moment, we multiply the F-P equation (2) by v and integrate in velocity. The resulting equation is

    ϵ2tJϵ(t,x)+ϵ(xPϵ(t,x)+V(x)ρϵ(t,x))=G(x)Jϵ(t,x). (11)

    As we show in our proof, the main contributions in (11) come from the rhs term and the second and third terms in the lhs. Indeed, rewriting the pressure tensor we have

    vivjfϵdv=vi(M)vjfϵMdv=δijfϵdv+Mvi(fϵM)vjdv,

    which implies

    Pϵ(t,x)=ρϵI+Mv(fϵM)vdv. (12)

    With the help of (12), equation (11) now gives

    Jϵ=ϵG1(x)(xρϵ+V(x)ρϵ)ϵ2G1(x)tJϵϵG1(x)xMv(fϵM)vdv. (13)

    The last term in (13) contains the part Mv(fϵM)vdv which appears in the expression for Pϵ(t,x). This term will be shown to be of order ϵ if one uses the appropriate a priori estimate e.g. in L2(μ). This implies that in the limit ϵ0, we should be able to establish that Pϵ(t,x)ρ(t,x)I. The term ϵ2G1(x)tJϵ will be shown to be of order ϵ2, as long as we give an appropriate interpretation to a solution Jϵ(t,x) of (13). Hence, we will justify rigorously the following expansion for Jϵ,

    Jϵ(t,x)=ϵG1(x)(xρϵ+V(x)ρϵ)+ϵ2. (14)

    Finally, as we let ϵ0, the system of equations (10) & (14) converges to

    tρ+xJ=0J=G1(x)(xρ+V(x)ρ),

    where J is the limit of Jϵ/ϵ. At the same time, since fϵ approaches local Gibbs states, it follows that fϵρ(t,x)M(v). All this is enough to suggest that the limit equation for ρ solves the Smoluchowski equation (6).

    It is now time to give a brief step by step outline of the proof of Theorems 1.1 & 1.2. We begin with the first result, in which we show weak convergence to the solution of the limiting problem.

    In the first step of the proof, we decompose fϵ(t,x,v) into a local equilibrium state M(v)ρϵ(t,x), and a deviation M(v)˜gϵ(t,x,v). With the help of the a priori energy estimate we can extract convergent subsequences for ρϵ(t,x), ˜gϵ, and 1ϵG1/2(x)v˜gϵ(t,x,v). Then, we can show that ρϵ is compact in C([0,T],wL2(Rnx)), for any T>0. Next, we write an evolution equation for ˜gϵ(t,x,v) (an equation in the distributional sense) and pass to the limit ϵ0. To achieve this, since we are dealing with a weak formulation, we have to find the order in ϵ of each integral term in this equation and ignore all the lower order terms in ϵ. The last step is to use the limit equation for ˜gϵ(t,x,v) and the limit equation for ρϵ(t,x) to derive the Smoluchowski equation.

    In terms of the type of solutions we work with, we shall assume that the operator Lϵ generates a continuous semigroup in L2Meq, so we write fϵ(t,x,v)=etLϵfϵ(0,x,v). Using the maximum principle and energy dissipation (see Section 3), it is easy to show that solutions to tfϵ+Lϵfϵ=0 remain bounded in L2MeqL. We define

    Definition 2.1. A mild-weak solution fϵ of (2) lies in the space

    fϵC(R+;D(Rnx×Rnv))Lloc(R+;L2MeqL) (15)

    and satisfies

    fϵ(T,,)φ(,)dvdxfϵ(0,,)φ(,)dvdx1ϵT0(vxφV(x)vφ)fϵdvdxds+1ϵ2T0vφG(x)(vfϵ+vfϵ)dvdxds=0,

    for any test function φ(x,v)C1c(Rnx×Rnv) and T>0.

    For the second result, we use the relative entropy of fϵ with respect to local equilibrium states. The relative entropy functional H(f|g) between two probability densities f,g is a measure of distance between them. Indeed, by the celebrated Csiszár-Kullback-Pinsker inequality ([4,25,28]) we have

    fgL12H(f|g).

    Thus, by finding limϵ0H(fϵ|ρM) we can control the square of the L1 distance between fϵ and ρM in the limit ϵ0. Here we show that the dissipation of relative entropy H(fϵ|ρM) contains a non negative part and remainder terms. It is important to show that these remainder terms vanish as ϵ0. Once we show that in the limit the relative entropy is strictly dissipative, it will be enough to consider initial data ''prepared'' in a way such that H(fϵ(0,,)|ρ(0,)M)0 as ϵ0 and it follows that H(fϵ(t,,)|ρ(t,)M)0 with t[0,T], for any T>0.

    We work with weak solutions of equation (2). Such solutions have been shown to exist in [27] for coefficients that have a Sobolev type of regularity and satisfy certain growth assumptions (see Proposition 1 below).

    Definition 2.2. A weak solution fϵ of (2) belongs to the space

    X:={fϵ|fϵL([0,T];L1L)&G1/2vfϵ(L2([0,T],L2))n}, (16)

    for all times T>0 (with fϵ(0,,)L1L). To be more precise, a weak solution fϵ satisfies

    fϵ(T,,)φ(T,,)dvdxfϵ(0,,)φ(0,,)dvdxT0fϵtφdvdxds1ϵT0(vxφV(x)vφ)fϵdvdxds+1ϵ2T0vφG(x)(vfϵ+vfϵ)dvdxds=0,

    for any test function φ(t,x,v)C1((0,T);C1c(Rnx×Rnv))C([0,T];C1c(Rnx×Rnv)).

    Notice that the definition of a mild-weak solution (given earlier) is similar to the one for weak solutions presented above. Main difference is that in the case of weak solutions, the weak formulation requires that test functions are also functions of time t. The existence of a unique weak solution, for coefficients that are not smooth, is given in the following proposition borrowed from [27].

    Proposition 1. (see [27]) Assume that the potential V(x) and diffusion G1/2(x) satisfy the following assumptions:

    ()G(x)v+V(x)(W1,1loc(Rnx×Rnv))n()tr(G)L(Rnx)
    ()G(x)v+V(x)1+|x|+|v|(L(Rnx×Rnv))n
    ()G1/2(x)(W1,2loc(Rnx))n×n()G1/2(x)1+|x|(L(Rnx))n×n.

    Then, given initial data fϵ(0,,)L1L, there exists a unique weak solution fϵ of (2) that belongs to X.

    In this section we collect all the convergence results needed for the proof of Theorem 1.1. We begin with the decomposition of fϵ. We write

    fϵ=M(v)(ρϵ+˜gϵ),

    where the hydrodynamic variable ρϵ has already been defined in (9) and ˜gϵ is a deviation from the local equilibrium state ρϵM(v) that satisfies

    ˜gϵM(v)dv=0.

    We also note that integrating (2) in velocity we obtain the hydrodynamic equation for ρϵ

    tρϵ+1ϵxM(v)v˜gϵdv=0. (17)

    We prove the following.

    Lemma 3.1. Assume a mild-weak solution fϵ of (2) with an L2Meq bound on the initial data, i.e., fϵ(0,,)L2Meq<. Then, there exists a sequence ϵi0 such that

    ρϵiρweaklyinL2(dx)t0,˜gϵi˜gweaklyinL2(M(v)dvdx)t0,1ϵiG1/2v˜gϵiJweaklyinL2(M(v)dvdxdt).

    Proof. In order to study the limit ϵ0, we begin with the a priori estimate in L2Meq(Rnx×Rnv). This is an energy estimate for hϵ(t,x,v) in L2(dμ), with hϵ(t,x,v):=fϵ(t,x,v)/Meq. It is achieved by multiplying (2) with hϵ and integrating in dμ to get

    12h2ϵ(t,x,v)dμ+1ϵ2t0|G1/2(x)vhϵ(s,x,v)|2dμds=12h2ϵ(0,x,v)dμ. (18)

    To simplify the analysis, we consider the basic assumption infV(x)>. Then, a priori estimate (18) gives the following two bounds,

    ρ2ϵdx<,˜g2ϵM(v)dvdx<t0. (19)

    For the first bound in (19) we used a simple Jensen inequality on the L2(dμ) estimate for hϵ. We also have (as a result of equation (18)) the energy bound,

    1ϵ2T0|G1/2v˜gϵ|2M(v)dvdxds<for anyT>0. (20)

    Based on equations (19) & (20), and after picking a sequence ϵi0, we can extract a subsequence which without loss of generality we still call ϵi so that all the convergences in the statement of the lemma hold.

    It is important to comment that we want something stronger than just ρϵ being weakly compact in L2(dx)t0. We actually want a uniform (in time) type of convergence, so that we don't have a problem when we later pass to the limit in integrals of time. For this reason, we prove that ρϵ is compact in C([0,T],wL2(dx)) in the lemma that follows.

    Lemma 3.2. Under the assumptions of Theorem 1.1, ρϵ is compact in C([0,T],wL2(dx)), i.e.,

    ρϵρinC([0,T],wL2(dx)).

    Proof Consider the functional H(t)=ϕ(x)ρϵ(t,x)dx, 0<t<T, for a fixed T>0 and ϕCc(Rnx). H(t) can be proven to be pointwise finite for any 0<t<T, using the Cauchy-Schwartz inequality and always assuming finite initial energy.

    Now, if we consider t1,t2>0 such that 0t1t2T, we have

    H(t2)H(t1)=ϕ(x)ρϵ(t,x)|t=t2t=t1dx(Use weak form of (17))=1ϵt2t1xϕ(x)v˜gϵM(v)dvdxds(t2t1|G1/2xϕ(x)|2M(v)dvdxds)12(t2t1|G1/2v˜gϵ|2ϵ2M(v)dvdxds)12(t2t1)12(|G1/2xϕ(x)|2dx)12(t2t1|G1/2v˜gϵ|2ϵ2M(v)dvdxds)12C(t2t1)12.

    The Arzelá-Ascoli theorem states that pointwise boundedness and equicontinuity suffice to show that the family ϕ(x)ρϵ(t,x)dx is compact in C([0,T]) for a given function ϕCc(Rnx). Notice that condition G1(x)(L1loc(Rnx))n×n is important so that the first integral is finite.

    Next, we use a standard density argument to show that ϕ(x)ρϵ(t,x)dx is compact in C([0,T]) for ϕCc(Rnx). Since Cc(Rnx) is now a separable space, separability will allow us to make use of Cantor's diagonal argument and extract a subsequence ρϵj so that

    ϕ(x)ρϵj(t,x)dxϕ(x)ρ(t,x)dxasj,

    for any ϕ in a countable subset of Cc(Rnx), and uniformly on [0,T]. This last convergence can be extended to any ϕCc(Rnx) again by use of a density argument.

    We close by approximating any function ϕL2(Rnx) by a sequence ϕmCc(Rnx), so that ϕmϕ a.e. and ϕmϕL20. This way, we show

    ρϵ(t,x)(ϕ(x)ϕm(x))dx0asm0,

    uniformly in ϵ>0 and [0,T]. This yields that ρL2(Rnx) and that ρϵ is compact in C([0,T],wL2(Rnx)). (see [15]).

    Now that weak compactness of ρϵ has been established uniformly in [0,T], we can proceed with the derivation of an equation for the deviation ˜gϵ, i.e.,

    ϵt˜gϵxMv˜gϵdv+v(x(ρϵ+˜gϵ)+V(x)(ρϵ+˜gϵ))V(x)v˜gϵ=1ϵ1Mv(MG(x)v˜gϵ). (21)

    A mild solution of (21) will be in C(R+,D(Rnx×Rnv)). The weak formulation is given by the expression

    ϵM(v)φ(˜gϵ(t2)˜gϵ(t1))dvdx+t2t1M(v)xφ(M(v)v˜gϵdv)dvdxds+t2t1M(v)v(xφρϵ+φV(x)ρϵ)dvdxds+t2t1M(v)(xφv˜gϵ+φV(x)v˜gϵ)dvdxdst2t1M(v)φV(x)v˜gϵdvdxds=1ϵt2t1M(v)vφGv˜gϵdvdxds, (22)

    where φ(x,v)Cc(Rnx×Rnv). In the lemma that follows, we show what happens when we let ϵ0 in (22).

    Lemma 3.3. Under the assumptions of Theorem 1.1, in the limit ϵ0 the limiting functions ρ and J (from Lemma 3.1) satisfy

    t2t1M(v)v(xφρ+φV(x)ρ)dvdxds=t2t1M(v)vφG1/2JdvdxdsφCc(Rnx×Rnv). (23)

    Proof We use the notation Ij (1j6) for the integral terms that appear in the weak formulation (22) in their order of appearance. The study of the order of magnitude for each of them reveals that in the limit ϵ0 only terms I3 & I6 do not vanish. In all the estimates that follow we use (19) & (20), so that we have

    I1=ϵM(v)φ(˜gϵ(t2)˜gϵ(t1))dvdxϵ(φ2M(v)dvdx)1/2((|˜gϵ(t2)|2+|˜gϵ(t1)|2)M(v)dvdx)1/2Cϵ=O(ϵ).
    I2=t2t1M(v)M(v)xφ(x,v,s)v˜gϵ(x,v,s)dvdvdxdsϵt2t1M(v)M(v)|G1/2xφ||G1/2v˜gϵ|ϵdvdvdxdsϵ(t2t1M(v)|G1/2xφ|2dvdxds)1/2×(1ϵ2t2t1M(v)|G1/2v˜gϵ|2dvdxds)1/2Cϵ=O(ϵ).
    I3=t2t1M(v)v(xφρϵ+φV(x)ρϵ)dvdxds(t2t1M(v)|v|2(|xφ|2+|φV(x)|2)dvdxds)1/2(t2t1ρ2ϵdxds)1/2C(M(v)|v|2(|xφ|2+|φV(x)|2)dx)1/2(t2t1ρ2ϵdxds)1/2=O(1).
    I4=t2t1M(v)(xφv˜gϵ+φV(x)v˜gϵ)dvdxdsϵ(t2t1M(v)(|G1/2xφ|2+|φG1/2V(x)|2)dvdxds)1/2×(t2t11ϵ2|G1/2v˜gϵ|2M(v)dvdxds)1/2Cϵ=O(ϵ).
    I5=t2t1M(v)φV(x)v˜gϵdvdxdsϵ(t2t1M(v)|φG1/2V(x)|2dvdxds)1/2×(t2t11ϵ2|G1/2v˜gϵ|2M(v)dvdxds)1/2Cϵ=O(ϵ).
    I6=1ϵt2t1M(v)vφGv˜gϵdvdxds(t2t1|G1/2vφ|2M(v)dvdxds)1/2×(t2t11ϵ2|G1/2v˜gϵ|2M(v)dvdxds)1/2=O(1).

    Now that we have established all the above bounds, we take ϵ0 and use the convergence results in Lemmas 3.1 & 3.2 to derive (23).

    Proof of Theorem 1.1. We write the hydrodynamic equation (17) for ρϵ in its weak form, and take the limit ϵ0 to obtain

    ϕ()ρ(t,)|t=t2t=t1dx=t2t1M(v)xϕG1/2JdvdxdsϕCc(Rnx). (24)

    In order to give the limiting equation for ρ(t,x) we should combine (23) & (24). The two equations can be coupled for the choice of test function φ(x,v)=xϕG1v, where ϕCc(Rnx). The only problem is that this function is not smooth or compactly supported in Rnv, so we have to modify it slightly (for non smooth G1(x) regularization in x is also needed).

    We begin by taking the cut-off function χδ1(v)=χ(δ1v), where χ(v)Cc(Rnv) is a function with values 0χ(v)1 such that χ(v)=1for|v|1 and χ(v)=0for|v|2. We also consider the standard mollification function,

    ηδ2(v)=1δn2η(vδ2)forηCc(Rnv)such thatη(v)dv=1.

    We now take the function φδ1,δ2(x,v)=(χδ1(v)xϕG1v)ηδ2. A standard result for the mollified function is that φδ1,δ2 converges to φ a.e. in Rnv (as δ1,δ20). Obviously xφδ1,δ2 converges to xφ a.e. in Rnv, since the cut-off and mollification acts only in the v variable.

    By the substitution of φ with φδ1,δ2 in (23), we have

    t2t1M(v)v(xφδ1,δ2ρ+φδ1,δ2V(x)ρ)dvdxds=t2t1M(v)vφδ1,δ2G1/2Jdvdxds.

    We also have

    vφδ1,δ2(x,v)=v((xϕG1vχδ1(v))ηδ2)=v(xϕG1vχδ1(v))ηδ2=(xϕG1χδ1(v)+xϕG1vvχδ1(v))ηδ2,

    where we use the fact that v(fηδ)=vfηδ.

    A typical estimate for vχδ1(v) is |vχδ1(v)|Cδ1. This can be easily seen by the definition of χδ1 and the fact that |vχ|C for some C>0, since χCc(Rnv). This estimate, together with the computation of vφδ1,δ2(x,v) above and the dominated convergence theorem imply that in the limit δ1,δ20, we actually have that (23) holds with φ(x,v)=xϕ(x)G1v. This choice of test function allows the coupling of (24) and (23) that yields

    ϕ()ρ(t,)|t=t2t=t1dx=t2t1(x(G1xϕ)+xϕG1V(x))ρdxds,

    which is the weak form of (6). This completes the proof of Theorem 1.1.

    To prove Theorem 1.2, we begin with the a priori estimates that will be used later to show that the remainder term rϵ is of order O(ϵ). The exact formula for rϵ is given in the study of the evolution of H(fϵ|ρM) in Section 4.2.

    The following proposition contains all the estimates needed for a solution fϵ.

    Proposition 2. Assume fϵ is a solution of (2), with initial data satisfying (8). Let also 0<T<. Then, the following hold:

    (ⅰ) fϵ(1+V(x)+|v|2+lnfϵ) is bounded in L((0,T),L1(Rnx×Rnv)),

    (ⅱ) 1ϵG1/2(x)(vfϵ+2vfϵ) is in L2((0,T),Rnx×Rnv),

    (ⅲ) |v|2fϵ is bounded in L((0,T),L1(Rnx×Rnv)).

    Proof. (ⅰ) We introduce the free energy associated with the F-P equation (2),

    E(fϵ):=fϵ(lnfϵ+|v|22+V(x))dvdx.

    The free energy is dissipated since

    ddtE(fϵ)=1ϵ2|dϵ|2dvdx,

    where

    dϵ=G1/2(x)(vfϵ+2vfϵ)=2MG1/2(x)vfϵM.

    The dissipation of energy implies

    E(fϵ(T,,))+1ϵ2T0|dϵ|2dvdxds=E(fϵ(0,,)). (25)

    (ⅱ) It also follows from (25) that dϵ is of order O(ϵ) in L2, i.e.,

    T0|dϵ|2dvdxdtCϵ2,T>0.

    (ⅲ) We now give the bound for |v|2fϵdvdx in L(0,T). This bound (uniform in time) is a straightforward consequence of the elementary Frenchel-Young inequality

    abh(a)+h(b),

    where h, h are a Young's convex pair (h is explicitly computed by the Legendre transform of the convex function h). Here, we use h(z)=zlogz and h(z)=ez1, i.e.,

    14|v|2fϵdvdxfϵlogfϵMeqdvdx+e|v|241Meqdvdx.

    This implies that

    |v|2fϵ(t,v,x)dvdxCfor someC>0,t[0,T],

    since e^{-V(x)}\in L^{1}(\mathbb{R}^{n}_{x}) and the entropy integral is bounded by the a priori estimate.

    We now prove the following result for the evolution of the relative entropy.

    Lemma 4.1. Assume a sufficiently regular solution f_{\epsilon} of equation (2), with initial data f_{\epsilon}(0,\cdot, \cdot) satisfying (8). It is shown that

    H(f_{\epsilon}(t,\cdot,\cdot)|\rho(t,\cdot) \mathcal{M}) \leq H(f_{\epsilon}(0,\cdot,\cdot)|\rho(0,\cdot) \mathcal{M}) +\int_{0}^{t}r_{\epsilon}(s) \, ds , (26)

    for a remainder term r_{\epsilon} given explicitly by

    r_{\epsilon}(t)=- \int \left( \epsilon \partial_{t}J_{\epsilon} + \nabla_{x}\cdot \left( \int \mathcal{M} \nabla_{v} \left( \frac{f_{\epsilon}}{\mathcal{M}}\right) \otimes v \, dv \right)\right) \cdot G^{-1} \left( \frac{\nabla \rho}{\rho} +\nabla V(x)\right)\, dx . (27)

    Proof. We start with the computation of the evolution of the H(\rho_{\epsilon}|\rho) relative entropy. This computation becomes partly obsolete later when we perform a similar computation for H(f_{\epsilon}|\rho \mathcal{M}). Nevertheless, we begin with computing \frac{d}{dt}H(\rho_{\epsilon}|\rho), especially since it contains parts important in the computation of \frac{d}{dt}H(f_{\epsilon}|\rho \mathcal{M}). Hence,

    \begin{split} &\frac{d}{dt}H(\rho_{\epsilon}|\rho)=\frac{d}{dt}\int \rho_{\epsilon}\log{\frac{\rho_{\epsilon}}{\rho}} \, dx= \frac{d}{dt}\int \rho_{\epsilon} \log{\rho_{\epsilon}} \, dx -\frac{d}{dt} \int \rho_{\epsilon} \log{\rho} \, dx\\ &= \int \partial_{t}\rho_{\epsilon} (\log{\rho_{\epsilon}}+1) \, dx -\int \partial_{t}\rho_{\epsilon} \log{\rho} \, dx -\int \frac{\rho_{\epsilon}}{\rho} \partial_{t}\rho \, dx (\text{Use (10),(6)})\\ &=\frac{1}{\epsilon}\int J_{\epsilon} \cdot \frac{\nabla \rho_{\epsilon}} {\rho_{\epsilon}} \, dx -\frac{1}{\epsilon}\int J_{\epsilon} \cdot \frac{\nabla \rho}{\rho} \, dx -\int \frac{\rho_{\epsilon}}{\rho} \nabla \cdot \left( G^{-1}(\nabla \rho +\nabla V(x) \rho) \right) \, dx \\ &= \frac{1}{\epsilon}\int J_{\epsilon} \cdot \left( \frac{\nabla \rho_{\epsilon}} {\rho_{\epsilon}}-\frac{\nabla \rho}{\rho}\right)\, dx+ \int \rho \nabla \left( \frac{\rho_{\epsilon}}{\rho}\right) \cdot G^{-1} \left( \frac{\nabla \rho}{\rho} + \nabla V(x) \right) \, dx\\ &= \frac{1}{\epsilon}\int J_{\epsilon} \cdot \left( \frac{\nabla \rho_{\epsilon}} {\rho_{\epsilon}}-\frac{\nabla \rho}{\rho}\right)\, dx + \int \left( \frac{\nabla \rho_{\epsilon}}{\rho_{\epsilon}}-\frac{\nabla \rho}{\rho} \right)\cdot G^{-1} \left( \frac{\nabla \rho}{\rho} + \nabla V(x) \right) \rho_{\epsilon}\, dx \\ &= \int \left( \frac{\nabla \rho_{\epsilon}}{\rho_{\epsilon}}-\frac{\nabla \rho}{\rho} \right) \cdot \left( \frac{1}{\epsilon}\frac{J_{\epsilon}}{\rho_{\epsilon}}+ G^{-1} \left( \frac{\nabla \rho}{\rho} + \nabla V(x) \right)\!\! \right) \rho_{\epsilon}\, dx \\ &= \! - \! \int \!\! G \left( \frac{1}{\epsilon}\frac{J_{\epsilon}}{\rho_{\epsilon}}+ G^{-1} \left( \frac{\nabla \rho}{\rho} + \nabla V(x) \right)\!\! \right) \! \cdot \! \left(\frac{1}{\epsilon} \frac{J_{\epsilon}}{\rho_{\epsilon}}+ G^{-1} \left( \frac{\nabla \rho}{\rho} + \nabla V(x) \right)\!\! \right) \rho_{\epsilon} \, dx \\ &\quad +r'_{\epsilon}=- \int \Big| \frac{1}{\epsilon} G^{1/2} \frac{J_{\epsilon}}{\rho_{\epsilon}} +G^{-1/2}\left( \frac{\nabla \rho}{\rho} +\nabla V(x)\right)\Big|^{2} \rho_{\epsilon} \, dx +r'_{\epsilon}. \end{split}

    In the second to last equality, we made use of

    \begin{split} \frac{\nabla \rho_{\epsilon}}{\rho_{\epsilon}}-\frac{\nabla \rho}{\rho} =-\frac{1}{\epsilon} G \frac{J_{\epsilon}}{\rho_{\epsilon}}- \frac{\nabla \rho}{\rho}-&\nabla V(x) -\epsilon \frac{\partial_{t}J_{\epsilon}}{\rho_{\epsilon}} \\ -&\frac{1}{\rho_{\epsilon}} \nabla_{x} \cdot \int \mathcal{M} \nabla_{v} \left( \frac{f_{\epsilon}}{\mathcal{M}} \right) \otimes v \, dv , \end{split} (28)

    which is derived directly from (13). The remainder term r'_{\epsilon} equals

    r'_{\epsilon}= \! - \! \int \!\! \left( \epsilon \partial_{t}J_{\epsilon} +\nabla_{x}\! \cdot \!\! \int \mathcal{M} \nabla_{v}\left( \frac{f_{\epsilon}}{\mathcal{M}}\right) \otimes v \, dv \right) \! \cdot \! \left( \frac{1}{\epsilon}\frac{J_{\epsilon}}{\rho_{\epsilon}}+ G^{-1} \left( \frac{\nabla \rho}{\rho} + \nabla V(x) \right)\!\! \right) dx .

    Remark 1. Notice that r'_{\epsilon} is a remainder term that should vanish as \epsilon \to 0. We do not bother with showing that r'_{\epsilon} \to 0 in rigorous manner, as we mainly work with the relative entropy H(f_{\epsilon}|\rho \mathcal{M}). Yet, as we remark at the end of Section 4, the computation of \frac{d}{dt}H(\rho_{\epsilon}|\rho) alone can be used to establish the convergence of f_{\epsilon} that we prove in Theorem 1.2.

    At this point, we compute the evolution of H(f_{\epsilon}| \rho \mathcal{M}) in similar manner. To make things easier we can introduce the global equilibrium state \mathcal{M}_{eq}(x,v) in the computation that follows

    \begin{split} H(f_{\epsilon}|\rho \mathcal{M})&=\iint f_{\epsilon} \log{f_{\epsilon}} \, dv \, dx -\iint f_{\epsilon} \log{(\rho \mathcal{M})} \, dv \, dx \\ &= \iint f_{\epsilon} \log{\frac{f_{\epsilon}}{\mathcal{M}_{eq}}} \, dv \, dx + \iint f_{\epsilon} \log{\frac{\mathcal{M}_{eq}}{\rho \mathcal{M}}} \, dv \, dx \\ &= \iint f_{\epsilon} \log{\frac{f_{\epsilon}}{\mathcal{M}_{eq}}} \, dv \, dx +\int \rho_{\epsilon} \log{\frac{e^{-V(x)}}{\rho}} \, dx \\&=H(f_{\epsilon}|\mathcal{M}_{eq})-\int \rho_{\epsilon} \log{\rho} \, dx -\int \rho_{\epsilon}V(x) \, dx . \end{split} (29)

    The reason we introduced H(f_{\epsilon}|\mathcal{M}_{eq}) is that the term \frac{d}{dt}H(f_{\epsilon}|\mathcal{M}_{eq}) can be easily bounded by an integral involving only hydrodynamical variables. Indeed, the time derivative of H(f_{\epsilon}|\mathcal{M}_{eq}) is

    \begin{split} \frac{d}{dt} \iint f_{\epsilon}&\log{\frac{f_{\epsilon}}{\mathcal{M}_{eq}}} \, dv \, dx =-\frac{1}{\epsilon^{2}} \iint f_{\epsilon} \Big|G^{1/2}\nabla_{v}\log{\frac{f_{\epsilon}} {\mathcal{M}}} \Big|^{2} \, dv \, dx \\ &= -\frac{1}{\epsilon^{2}} \iint f_{\epsilon} \Big|G^{1/2}\left( \frac{\nabla_{v}f_{\epsilon}}{f_{\epsilon}} +v \right) \Big|^{2} \, dv \, dx \leq -\frac{1}{\epsilon^{2}}\int \frac{|G^{1/2}J_{\epsilon}|^{2}}{\rho_{\epsilon}} \, dx . \end{split} (30)

    The last inequality in (30) is in fact due to Hölder,

    \begin{align*} \int \frac{|G^{1/2}J_{\epsilon}|^{2}}{\rho_{\epsilon}} \, dx&= \int \frac{\left( \int G^{1/2} \left(v+\frac{\nabla_{v}f_{\epsilon}}{f_{\epsilon}} \right)f_{\epsilon} \, dv \right)^{2}}{\rho_{\epsilon}} \, dx \\& \leq \iint \Big|G^{1/2}\left( \frac{\nabla_{v} f_{\epsilon}}{f_{\epsilon}}+v \right) \Big|^{2} f_{\epsilon} \, dv \, dx . \end{align*}

    Combining equations (29) & (30), we obtain

    \begin{align*} \frac{d}{dt}H(f_{\epsilon}|\rho \mathcal{M})&=\frac{d}{dt}H(f_{\epsilon}|\mathcal{M}_{eq})-\frac{d}{dt} \int \rho_{\epsilon} \log{\rho} \, dx -\frac{d}{dt} \int \rho_{\epsilon} V(x) \, dx \\ &\leq - \frac{1}{\epsilon^{2}}\int \frac{|G^{1/2}J_{\epsilon}|^{2}}{\rho_{\epsilon}} \, dx - \frac{d}{dt} \int \rho_{\epsilon} \log{\rho} \, dx - \frac{1}{\epsilon} \int J_{\epsilon} \cdot \nabla V(x) \, dx .\end{align*}

    The computation of \frac{d}{dt} \int \rho_{\epsilon}\log{\rho} \, dx has been performed as a part of the computation of \frac{d}{dt}H(\rho_{\epsilon}|\rho) above. We thus have

    \begin{split} \frac{d}{dt}&H(f_{\epsilon}|\rho \mathcal{M}) \leq -\frac{1}{\epsilon^{2}}\int \frac{|G^{1/2}J_{\epsilon}|^{2}}{\rho_{\epsilon}} \, dx -\frac{1}{\epsilon}\int J_{\epsilon} \cdot \nabla V(x) \, dx -\frac{1}{\epsilon}\int J_{\epsilon} \cdot \frac{\nabla \rho}{\rho} \, dx\nonumber \\ &+ \int \left( \frac{\nabla \rho_{\epsilon}}{\rho_{\epsilon}} -\frac{\nabla \rho}{\rho}\right)\cdot G^{-1} \left( \frac{\nabla \rho}{\rho} +\nabla V(x) \right) \rho_{\epsilon} \, dx (\text{Use (28)}) \\ &= -\int \frac{1}{\epsilon} J_{\epsilon} \cdot \left(\frac{1}{\epsilon} G \frac{J_{\epsilon}} {\rho_{\epsilon}} +\frac{\nabla \rho}{\rho} +\nabla V(x) \right) \, dx \\ &-\int \left( \frac{1}{\epsilon} G \frac{J_{\epsilon}}{\rho_{\epsilon}} + \frac{\nabla \rho}{\rho} +\nabla V(x) \right) \cdot G^{-1} \left( \frac{\nabla \rho}{\rho}+\nabla V(x) \right) \rho_{\epsilon} \, dx + r_{\epsilon} \\ &= -\int \! \left( \frac{1}{\epsilon} G \frac{J_{\epsilon}}{\rho_{\epsilon}} +\frac{\nabla \rho}{\rho} +\nabla V(x)\right) \! \cdot \! \left( \frac{1}{\epsilon}\frac{J_{\epsilon}}{\rho_{\epsilon}}+G^{-1} \left( \frac{\nabla \rho}{\rho}+\nabla V(x) \right)\! \right) \rho_{\epsilon} \, dx + r_{\epsilon} \\ &= -\int \Big| \frac{1}{\epsilon} G^{1/2}\frac{J_{\epsilon}}{\rho_{\epsilon}} +G^{-1/2} \left( \frac{\nabla \rho}{\rho}+ \nabla V(x) \right)\Big|^{2} \rho_{\epsilon} \, dx + r_{\epsilon}, \end{split} (31)

    with a remainder term r_{\epsilon} given by (27). Finally, we integrate (31) in time and (26) follows.

    We already gave the formal computation for \frac{d}{dt}H(f_{\epsilon}|\rho \mathcal{M}). Our goal is to prove that \int_{0}^{t} r_{\epsilon}(s)\, ds \to 0. The remainder term r_{\epsilon} that we computed in Lemma 4.1 consists of two parts r_{1,\epsilon} and r_{2,\epsilon}, which integrated in time are

    \int_{0}^{T}r_{1,\epsilon} \, dt =-\epsilon \int_{0}^{T}\!\!\!\! \iint \partial_{t}f_{\epsilon} \, v \cdot G^{-1}\left( \frac{\nabla \rho}{\rho} +\nabla V(x)\right) \, dv \, dx \, dt ,
    \int_{0}^{T} r_{2,\epsilon} \, dt =\int_{0}^{T}\!\!\!\! \iint \left(\mathcal{M}\, v \otimes \nabla_{v} \left( \frac{f_{\epsilon}}{\mathcal{M}}\right)\right) : \nabla \left( G^{-1}\left( \frac{\nabla \rho}{\rho} +\nabla V(x)\right)\right) \, dv \, dx \, dt .

    Our task is to show that both integrals vanish as \epsilon \to 0.

    In the process of controlling the two terms, we introduce a new notation for expressions involving the hydrodynamic variable \rho. Thus, we denote with D the tensor D:=\nabla(G^{-1} (\nabla \log{\rho}+\nabla V(x))), and with E,F the vectors E:=G^{-1}(\nabla \log{\rho}+\nabla V(x)) and F:=G^{-1}\nabla \partial_{t} \log{\rho}. The easiest term to control is

    \begin{split} \Big|\int_{0}^{T}& r_{2,\epsilon}\, dt \Big| = \Big| \int_{0}^{T}\!\!\!\! \iint \mathcal{M} v \otimes \nabla_{v}\left( \frac{f_{\epsilon}}{\mathcal{M}}\right) : D \, dv \, dx \, dt \Big| \\ &\leq \int_{0}^{T} \!\!\!\! \iint | \sqrt{f_{\epsilon}} \left( v \otimes G^{-1/2}(x)d_{\epsilon}\right) : D | \, dv \, dx \, dt \\ &\leq C \epsilon \| D\|_{\infty} \! \left( \int_{0}^{T}\!\!\!\! \iint \frac{|G^{-1/2}(x)d_{\epsilon}|^{2}}{\epsilon^{2}} \, dv \, dx \, dt\right)^{1/2}\!\!\! \left( \int_{0}^{T}\!\!\!\! \iint |v|^{2}f_{\epsilon} \, dv \, dx \, dt \right)^{1/2}\!\!\!\!\!. \end{split} (32)

    Finally, for the first term we have

    \begin{split} \Big| \int_{0}^{T}&r_{1,\epsilon} \, dt \Big| = \Big| -\epsilon \iint \left(f_{\epsilon}(T,v,x)-f_{\epsilon}(0,v,x)\right)v \cdot E \, dv \, dx \\ &+ \epsilon \int_{0}^{T}\!\!\!\! \iint f_{\epsilon} \, v \cdot F \, dv \, dx \, dt \Big| \\ &\leq \epsilon \| E \|_{\infty} \! \left( \left( \iint \! f_{\epsilon}(T,x,v)|v|^{2}\, dv \, dx \right)^{1/2}\!\!\! + \left( \iint \! f_{\epsilon}(0,x,v)|v|^{2}\, dv \, dx\right)^{1/2} \right)\\ &\quad + \epsilon T^{1/2} \| F \|_{\infty} \! \left( \int_{0}^{T}\!\!\!\! \iint f_{\epsilon}(s,x,v)|v|^{2}\, dv \, dx \, ds \right)^{1/2}\!\!. \end{split} (33)

    We now show why the terms D, E, F are in L^{\infty}([0,T],\mathbb{R}^{n}_{x}).

    Stability estimates for the terms D, E, F.

    The control of terms D,E,F is achieved by controlling h_{0}=\frac{\rho}{e^{-V(x)}} and its derivatives. Notice that the L^{\infty} bound on \log \frac{\rho}{e^{-V(x)}} implies a bound of the type a e^{-V(x)} \leq \rho \leq A e^{-V(x)} for a,A>0. Such control of h_{0} is a direct consequence of the maximum principle for the Smoluchowski equation (6). Indeed, if a<h_{0}(0,x)<A it follows by the maximum principle that a<h_{0}(t,x)<A for all t \in [0,T], under the condition that \nabla \cdot (G^{-1}(x)\nabla V(x))<\infty and given that \sup_{0 \leq t \leq T} \limsup_{x \to \infty}| h_{0}(t,x)|\leq C_{0}. The control of derivatives also follows from a parabolic maximum principle as we prove in

    Lemma 4.2. Let \rho(t,x) be a solution to the Smoluchowski equation (6) with initial data \rho(0,x). Assume also that conditions (A1)-(A3) are satisfied. It follows that D, E, F are in L^{\infty}([0,T],\mathbb{R}^{n}_{x}). More precisely, \|D(t,.)\|_{L^{\infty}(\mathbb{R}^{n}_{x})} \leq C \|D(0,.)\|_{L^{\infty}(\mathbb{R}^{n}_{x})} for 0 \leq t \leq T, with similar estimates for E and F.

    Proof. First, we define h_{k}:=\nabla^{k}\frac{\rho}{e^{-V(x)}} and we want to prove that \| h_{k} \|_{L^{\infty}(\mathbb{R}^{n}_{x})} remains bounded on the interval [0,T] for 0\leq k \leq 3. It is enough to show that \| h_{k}(t,.)\|_{L^{\infty}}\leq C \| h_{k}(0,.)\|_{L^{\infty}} for 0 \leq t \leq T, with the constant C depending on T and other constants from the bounds in (A1).

    The time evolution of h_{0} is given by

    \begin{align*} \partial_{t} \left(\frac{\rho}{e^{-V(x)}}\right) &=\frac{\nabla \cdot \left(e^{-V(x)}G^{-1}(x) \nabla \frac{\rho}{e^{-V(x)}}\right)}{e^{-V(x)}} = \\&\nabla \cdot \left(G^{-1}(x)\nabla \frac{\rho}{e^{-V(x)}}\right) - \nabla V(x) \cdot G^{-1}(x)\nabla \frac{\rho}{e^{-V(x)}}. \end{align*}

    In the notation we introduced for h_{0}, this is written as

    \partial_{t}h_{0}=\nabla \cdot (G^{-1}(x) \nabla h_{0}) -\nabla V(x)\cdot G^{-1}(x)\nabla h_{0} . (34)

    Differentiating equation (34) m times, taking the inner product (for tensors) with h_{m} and integrating by parts we obtain an L^{2} estimate. As a matter of fact, we can get an L^{p} theory (for any p>1) and as a result a maximum principle for |h_{m}| given that we have the appropriate control of the coefficients.

    For instance, for h_{0} the L^{p} estimate is

    \begin{align*} \frac{d}{dt}\int h^{p}_{0}\, dx &+p(p-1)\int h^{p-2}_{0}\, \nabla h_{0} \cdot G^{-1}(x)\nabla h_{0}\, dx \\ &-\int h^{p}_{0}\, \nabla \cdot (G^{-1}(x)\nabla V(x))\, dx=0, p > 1, \end{align*}

    which under the assumption \|\nabla \cdot (G^{-1}\nabla V(x)) \|_{L^{\infty}}<\infty yields \|h_{0}(t,.) \|_{L^{\infty}}\leq C \|h_{0}(0,.) \|_{L^{\infty}} for 0 \leq t \leq T. It should be noted that for a divergence free or identity hydrodynamic mobility, the first condition translates to |\nabla^{2}V(x)|<C for the potential V(x).

    With a bit more work we obtain (see [27])

    \partial_{t}\left( \frac{|h_{1}|^{2}}{2}\right)-\nabla \cdot \left( G^{-1}(x)\nabla \left( \frac{|h_{1}|^{2}}{2}\right)\right) +\nabla V(x)\cdot G^{-1}(x)\nabla \left(\frac{|h_{1}|^{2}}{2} \right)\leq C|h_{1}|^{2}, (35)

    where the constant C now depends on \|\nabla G^{-1}\|_{L^{\infty}} and \|\nabla (G^{-1}\nabla V(x))\|_{L^{\infty}}. Using the maximum principle in (35), we have \|h_{1}(t,.) \|_{L^{\infty}}\leq C \|h_{1}(0,.) \|_{L^{\infty}} for 0 \leq t \leq T. The maximum principle for |h_{2}| and |h_{3}| is implemented in similar fashion leading to estimates just like (35).

    Remark 2. The careful reader has already noticed that Lemma 4.2 is the only part in the proof of Theorem 1.2 where we made use of conditions (A1)-(A3). These conditions appear to be optimal, at least when the diffusion limit is considered in unbounded space. For instance, the control of coefficients in (A1) is necessary for showing the propagation of Lipschitz regularity by deriving (35).

    Proof of Theorem 1.2. In Lemma 4.1 we have shown that the evolution of the relative entropy is controlled by the term \int_{0}^{T}r_{\epsilon}\, dt=\int_{0}^{T}r_{1,\epsilon}\, dt+\int_{0}^{T}r_{2,\epsilon}\, dt which is bounded with the help of equations 32 & 33. Using Proposition 2 and Lemma 4.2, it follows that \int_{0}^{T}r_{\epsilon}\, dt \to 0 as \epsilon \to 0, proving Theorem 1.2.

    The computations involving the relative entropy in this Section have been so far performed at a formal level, i.e., by assuming smooth solutions with derivatives vanishing polynomially fast. It is not a hard task to give a rigorous derivation of the results by performing a standard regularization argument which amounts to regularizing all the involved functions e.g. by convoluting with a mollifier, perform all the computations with the regularized ones, and finally pass to the limit. In fact, the whole procedure we present here follows closely the steps of the regularization argument in [27].

    The regularization procedure will be presented here for the simpler case H(\rho_{\epsilon}|\rho), since there are less computation involved and the reader can get a better grasp of the full argument. We begin with the assumption of smooth coefficients G(x), V(x) and we approximate a solution f_{\epsilon} by a mollified one f_{\epsilon,\delta}=f_{\epsilon}\star \eta_{\delta}\in C^{\infty}(\mathbb{R}^{n}_{x}\times\mathbb{R}^{n}_{v}). The mollifier is \eta_{\delta}=\frac{1}{\delta^{2n}}\eta\left(\frac{x}{\delta},\frac{v}{\delta}\right), with \eta \in C^{\infty}_{c}(\mathbb{R}^{n}_{x}\times \mathbb{R}^{n}_{v}) and \iint \eta(x,v)\, dv dx=1.

    The equation for the regularized f_{\epsilon,\delta} is (see [27])

    \partial_{t}f_{\epsilon,\delta}+L_{\epsilon}f_{\epsilon,\delta}=U^{1}_{\epsilon,\delta} +\nabla_{v}\cdot (G^{1/2}R^{1}_{\epsilon,\delta}) , (36)

    where the expressions U^{1}_{\epsilon,\delta}, R^{1}_{\epsilon,\delta} involve the following commutators

    \begin{align*} U^{1}_{\epsilon,\delta}=&-\frac{1}{\epsilon}[\eta_{\delta},v \cdot \nabla_{x} -\nabla V(x)\cdot \nabla_{v}](f_{\epsilon})+\frac{1}{\epsilon^{2}} [\eta_{\delta},(Gv)\cdot \nabla_{v}](f_{\epsilon})\\&+ \frac{1}{\epsilon^{2}}[\eta_{\delta},\nabla_{v}\cdot (Gv)](f_{\epsilon})+\frac{1}{\epsilon^{2}} [\eta_{\delta},G^{1/2}\nabla_{v}](G^{1/2}\nabla_{v}f_{\epsilon}), \\ R^{1}_{\epsilon,\delta}=&\frac{1}{\epsilon^{2}}[\eta_{\delta},G^{1/2}\nabla_{v}](f_{\epsilon}).\end{align*}

    The notation we follow for commutators is

    [\eta_{\delta},c](f)=\eta_{\delta}\star (cf)-c(\eta_{\delta} \star f) \;\; \text{and} \;\; [\eta_{\delta},c_{1}](c_{2}f)=\eta_{\delta}\star (c_{1}\cdot c_{2}f)-c_{1}\cdot(\eta_{\delta} \star c_{2}f),

    where c is a differential operator (or vector), and c_{1},c_{2} are general differential vectors. This implies that the equation for \rho_{\epsilon,\delta}:=\int f_{\epsilon,\delta}\, dv is

    \partial_{t}\rho_{\epsilon,\delta}+\frac{1}{\epsilon}\nabla_{x} \cdot J_{\epsilon,\delta}= \int U^{1}_{\epsilon,\delta}\, dv , (37)

    where J_{\epsilon,\delta}:=\int v f_{\epsilon,\delta}\, dv.

    The regularized limiting equation for \rho (with \rho_{\delta}=\rho \star \eta_{\delta}) is

    \partial_{t}\rho_{\delta}=\nabla_{x} \cdot (G^{-1}(\nabla_{x} \rho_{\delta} +\nabla V(x)\rho_{\delta}))+U^{2}_{\delta}+\nabla_{x} \cdot (G^{-1/2}R^{2}_{\delta}) , (38)

    with

    \begin{align*} U^{2}_{\delta}=&[\eta_{\delta},\nabla_{x}\cdot (G^{-1}\nabla V(x))](\rho)+ [\eta_{\delta},G^{-1}\nabla V(x) \cdot \nabla_{x}](\rho) \\ &+[\eta_{\delta},\nabla_{x}\cdot G^{-1/2}](G^{-1/2}\nabla_{x}\rho) +[\eta_{\delta},G^{-1/2}\nabla_{x}](G^{-1/2}\nabla_{x}\rho), \\ R^{2}_{\delta}=&[\eta_{\delta},G^{-1/2}\nabla_{x}](\rho) .\end{align*}

    It has be shown (see Section 5.3 in [27]) that

    \begin{align*} U^{1}_{\epsilon,\delta},\, \, U^{2}_{\delta}&\xrightarrow{\delta \to 0} 0 \quad L^{\infty}+L^{2}([0,T],L^{1}_{loc}) \\ R^{1}_{\epsilon,\delta},\, \, R^{2}_{\delta}&\xrightarrow{\delta \to 0} 0 \quad L^{\infty}([0,T],L^{2}_{loc}),\end{align*}

    for fixed \epsilon >0, as long as conditions in Proposition 1 are satisfied.

    Next, multiplying (36) by v and integrating in velocity while using the definition of \rho_{\epsilon,\delta} we get

    \begin{split} \epsilon^{2} \partial_{t}J_{\epsilon,\delta} &+ \epsilon(\nabla_{x}\rho_{\epsilon,\delta} + \nabla V(x)\rho_{\epsilon,\delta}) +\epsilon \nabla_{x} \cdot \int \mathcal{M} \nabla_{v}\left( \frac{f_{\epsilon,\delta}}{\mathcal{M}}\right) \otimes v \, dv \\ &+G J_{\epsilon,\delta}=\epsilon^{2}\int v \, U^{1}_{\epsilon,\delta}\, dv - \epsilon^{2}\int G^{1/2}R^{1}_{\epsilon,\delta} \, dv .\end{split} (39)

    Since we want to take advantage of the fact that commutators vanish (as \delta \to 0) on compact sets, we have to introduce a smooth cut-off function \phi_{R}(x)=\phi\left( \frac{x}{R}\right), where \phi is a smooth function on \mathbb{R}^{n}_{x}, s.t. 0\leq \phi \leq 1, with \phi(x)=1 for |x|\leq 1 and \phi(x)=0 for |x|\geq 2. It follows that \nabla \phi_{R}(\cdot)=\frac{1}{R}\nabla \phi \left( \frac{\cdot}{R}\right). The idea is to include the function \phi_{R} in every integral and send R \to \infty, after sending \delta \to 0. That way we can make integral terms that involve commutators vanish.

    For this reason, we introduce a relative entropy integral with a cut-off H_{R}(\rho_{\epsilon,\delta}|\rho_{\delta})=\int\rho_{\epsilon,\delta} \log\frac{\rho_{\epsilon,\delta}}{\rho_{\delta}} \phi_{R}\, dx. Differentiating the entropy H_{R}(\rho_{\epsilon,\delta}|\rho_{\delta}) and using (37)-(39) we obtain

    \begin{align*} \frac{d}{dt}H_{R}(\rho_{\epsilon,\delta}|\rho_{\delta}) &=\int \partial_{t}\rho_{\epsilon,\delta}(\log \rho_{\epsilon,\delta}+1)\phi_{R} \, dx - \int \partial_{t}\rho_{\epsilon,\delta} \log \rho_{\delta} \phi_{R} \, dx \\&- \int \frac{\rho_{\epsilon,\delta}}{\rho_{\delta}} \partial_{t}\rho_{\delta}\phi_{R} \, dx= \ldots = I_{1}+I_{2} +I_{3} .\end{align*}

    The expressions I_{1}, I_{2}, I_{3} that involve commutators are

    \begin{align*} I_{1}(\epsilon,\delta,R)=\iint U^{1}_{\epsilon,\delta}(\log \rho_{\epsilon,\delta}+1)&\phi_{R}\, dv dx- \iint U^{1}_{\epsilon,\delta}\log \rho_{\delta}\phi_{R}\, dv dx \\&-\iint (U^{2}_{\delta}+\nabla_{x}\cdot (G^{-1/2}R^{2}_{\delta}))\frac{\rho_{\epsilon,\delta}}{\rho_{\delta}}\phi_{R}\, dv dx ,\end{align*}
    \begin{align*} I_{2}(\epsilon,\delta,R)=\frac{1}{\epsilon}\int J_{\epsilon,\delta} \cdot \nabla \phi_{R}(\log &\rho_{\epsilon,\delta}+1)\, dx -\frac{1}{\epsilon}\int J_{\epsilon,\delta}\cdot \nabla \phi_{R}\log \rho_{\delta}\, dx \\ &+\int \rho_{\epsilon,\delta}\nabla \phi_{R}\cdot G^{-1} \left( \frac{\nabla_{x}\rho_{\delta}}{\rho_{\delta}}+\nabla V(x)\right)\, dx ,\end{align*}

    and

    I_{3}(\epsilon,\delta,R)=-\int \Big| \frac{1}{\epsilon}G^{1/2} \frac{J_{\epsilon,\delta}}{\rho_{\epsilon,\delta}}+G^{-1/2} \left( \frac{\nabla \rho_{\delta}}{\rho_{\delta}}+\nabla V(x)\right)\Big|^{2}\rho_{\epsilon,\delta}\phi_{R}\, dx+r'_{\epsilon,\delta,R} ,

    with the remainder term being

    \begin{align*} r'_{\epsilon,\delta,R}&=\!\!-\!\!\int \!\!\! \left( \epsilon \partial_{t}J_{\epsilon,\delta} +\!\!\nabla_{x}\!\cdot \!\int \mathcal{M}\nabla_{v}\left( \frac{f_{\epsilon,\delta}}{\mathcal{M}}\right) \otimes v \, dv \!-\!\epsilon \!\! \int v U^{1}_{\epsilon,\delta}\, dv +\!\epsilon \!\!\int G^{1/2}R^{1}_{\epsilon,\delta}\, dv \right) \\& \cdot \left( \frac{1}{\epsilon}\frac{J_{\epsilon,\delta}}{\rho_{\epsilon,\delta}}+ G^{-1}\left( \frac{\nabla \rho_{\delta}}{\rho_{\delta}}+\nabla V(x)\right)\right)\phi_{R} \, dx .\end{align*}

    The trick is to take both \delta \to 0 and R \to \infty while letting \epsilon \to 0. Since we have the freedom of choice of how \delta,R should behave for a fixed \epsilon, we will consider them as functions of \epsilon which we will describe in detail, i.e., \delta(\epsilon) and R(\epsilon). Indeed, for a given \epsilon>0, consider \delta(\epsilon) s.t. |U^{1}_{\epsilon,\delta}|,|R^{1}_{\epsilon,\delta}| < \epsilon, for all \delta < \delta(\epsilon). This way, we have |U^{1}_{\epsilon,\delta}|,|R^{1}_{\epsilon,\delta}| \to 0 while we let both \epsilon, \delta(\epsilon)\to 0.

    If we consider R fixed and take \delta \to 0, it is easy to see by the convergence properties of commutators (L^{\infty} in time) that \int_{0}^{T}I_{1}\, dt \to 0. The exception is the last term of I_{1} that is treated separately. Same thing holds for the part of the remainder term that involves commutators as we let \delta \to 0.

    A bound for the first term in I_{2} is

    \Big| \frac{1}{\epsilon}\! \int_{0}^{T}\!\!\!\!\int \!\! J_{\epsilon,\delta} \cdot \nabla \phi_{R}(\log \rho_{\epsilon,\delta}+1)\, dx dt \Big| \leq \frac{1}{R} \| \nabla \phi\|_{L^{\infty}} \!\! \int_{0}^{T}\!\!\!\!\int_{|x| > R}\!\!\! \frac{|J_{\epsilon,\delta}|} {\epsilon} |(\log \rho_{\epsilon,\delta}+1)| \, dx dt.

    The exact same treatment holds for the second term in I_{2}. It is obvious that these two integrals will vanish in the limit R \to \infty (partly due to the stability results similar to Lemma 4.2). It will not matter how fast R tends to infinity, so we can choose e.g. R(\epsilon)=\frac{1}{\epsilon}. For the third term in I_{2}, we have

    \begin{align*} \Big| \int_{0}^{T}\!\!\!\!\int \rho_{\epsilon,\delta}\nabla \phi_{R}\cdot G^{-1} \left( \frac{\nabla_{x}\rho_{\delta}}{\rho_{\delta}}+\nabla V(x)\right)\,&dx dt \Big| \leq C \Big \| \frac{G^{-1}(x)}{1+|x|}\Big\|_{L^{\infty}} \| \nabla \phi \|_{L^{\infty}} \\&\cdot \int_{0}^{T}\!\!\!\!\int_{|x| > R} |\rho_{\epsilon,\delta}| \Big| \frac{\nabla_{x}\rho_{\delta}}{\rho_{\delta}}+\nabla V(x) \Big|\, dx dt,\end{align*}

    which vanishes as R \to \infty given the growth condition (see Proposition 1) in G^{-1}(x).

    The last term in I_{1} which equals

    \iint G^{-1/2}R^{2}_{\delta}\cdot \nabla_{x}\left(\frac{\rho_{\epsilon,\delta}}{\rho_{\delta}}\right)\phi_{R}\, dv dx +\iint G^{-1/2}R^{2}_{\delta}\cdot \nabla \phi_{R} \, \frac{\rho_{\epsilon,\delta}}{\rho_{\delta}} \, dv dx ,

    contains two terms. The first one is treated like the terms in I_{1}, and the second like these in I_{2} with a growth condition for G^{-1/2} (Proposition 1).

    In the last step, we send \epsilon \to 0 (while \delta(\epsilon)\to 0 and R(\epsilon)\to \infty) and combine this with the fact that \lim \limits_{\epsilon \to 0}\int_{0}^{T}I_{1}(\epsilon,\delta(\epsilon),R(\epsilon))\, dt=\lim \limits_{\epsilon \to 0}\int_{0}^{T}I_{2}(\epsilon,\delta(\epsilon),R(\epsilon))\, dt=0 to derive

    \lim \limits_{\epsilon \to 0}H_{R(\epsilon)}(\rho_{\epsilon,\delta(\epsilon)}(T,\cdot)|\rho_{\delta(\epsilon)}(T)) \leq \lim \limits_{\epsilon \to 0}H_{R(\epsilon)}(\rho_{\epsilon,\delta(\epsilon)}(0,\cdot)|\rho_{\delta(\epsilon)}(0)) + \!\! \int_{0}^{T}\!\!\!\lim \limits_{\epsilon \to 0} r'_{\epsilon,\delta,R} \, dt.

    We finish with the estimates of previous subsection that prove that the remainder term vanishes as \epsilon \to 0. This yields the desired estimate

    \lim \limits_{\epsilon \to 0}H(\rho_{\epsilon}(T,\cdot)|\rho(T,\cdot)) \leq \lim \limits_{\epsilon \to 0}H(\rho_{\epsilon}(0,\cdot)|\rho(0,\cdot)) .

    Finally, we can remove the assumption on the smoothness of coefficients by regularizing them in x and pass to the limit.

    Remark 3. The regularization procedure above was carried out for the H(\rho_{\epsilon}|\rho) relative entropy which instantly implies f_{\epsilon}\to \rho \mathcal{M} in L^{1}. Indeed, by showing L^{1} convergence of \rho_{\epsilon}(t,x) to the limiting distribution \rho(t,x) it follows that f_{\epsilon} converges to \rho \mathcal{M} (in L^{1}) using the following simple argument. We decompose f_{\epsilon}-\rho \mathcal{M} as in

    f_{\epsilon}-\rho \mathcal{M}=f_{\epsilon}- \rho_{\epsilon}\mathcal{M}+(\rho_{\epsilon}-\rho)\mathcal{M} .

    It is trivial to show that the second term (\rho_{\epsilon}-\rho)\mathcal{M} of the decomposition \to 0 in L^{1} by assumption. For the first term f_{\epsilon}-\rho_{\epsilon}\mathcal{M}, we have

    \begin{align*} \| f_{\epsilon}-\rho_{\epsilon} \mathcal{M}\|_{L^{1}}&\leq \! \sqrt{2} \left( \iint f_{\epsilon}\log{\frac{f_{\epsilon}}{\rho_{\epsilon} \mathcal{M}}}\, dv \, dx \! \right)^{1/2} \!\!\! \leq \! \sqrt{2} \epsilon \left( \iint \frac{|G^{-1/2}d_{\epsilon}|^{2}}{\epsilon^{2}} \, dv \, dx \!\right)^{1/2} \\&\leq \sqrt{2} \epsilon C \to 0 \quad \text{as} \quad \epsilon \to 0. \end{align*}

    The inequalities used in the first line are the Csiszár-Kullback-Pinsker and log-Sobolev in that order. Finally, the a priori energy bound (used in second line) concludes the argument.

    The author is indebted to C. David Levermore and P-E Jabin for the discussions that led to the birth of this work. Special thanks to Athanasios Tzavaras for the initial motivation that led to the consideration of this problem. Manoussos Grillakis and Julia Dobrosotskaya helped by proofreading this article.

    [1] The nonaccretive radiative transfer equations: Existence of solutions and Rosseland approximation. J. Funct. Anal. (1988) 77: 434-460.
    [2] (1994) Dynamics of Polymeric Liquids: Kinetic Theory, vol 2.John Wiley & Sons.
    [3] Friction and mobility of many spheres in Stokes flow. J. Chem. Phys. (1994) 100: 3780-3790.
    [4] Information-type measures of difference of probability distributions and indirect observations. Stud. Sci. Math. Hung. (1967) 2: 299-318.
    [5] Diffusion limit for nonhomogeneous and non-micro-reversible processes. Indiana Univ. Math. J. (2000) 49: 1175-1198.
    [6] Kinetic models for polymers with inertial effects. Netw. Heterog. Media (2009) 4: 625-647.
    [7] (1996) Introduction to Polymer Physics.Oxford University Press.
    [8] (1986) The Theory of Polymer Dynamics. New York: Oxford University Press.
    [9] Non linear diffusions as limit of kinetic equations with relaxation collision kernels. Arch. Ration. Mech. Anal. (2007) 186: 133-158.
    [10] Über die von der molekularkinetischen Theorie der Wärme geforderte Bewegung von in ruhenden Flüssigkeiten suspendierten Teilchen. Ann. Phys. (1905) 322: 549-560.
    [11] Some remarks on the Smoluchowski-Kramers approximation. J. Stat. Phys. (2004) 117: 617-634.
    [12] Diffusion limit of The Vlassov-Poisson-Fokker-Planck system. Commun. Math. Sci. (2010) 8: 463-479.
    [13] F. Golse, C. D. Levermore and L. Saint-Raymond, La Méthode de L'entropie Relative Pour les Limites Hydrodynamiques de Modéles Cinétiques Séminaire Equations aux Derivées Partielles, Exp. No. XIX, Ecole Polytechnique, 2000.
    [14] Limite fluide des équations de Boltzmann des semi-conducteurs pour une statistique de Fermi-Dirac. Asympot. Anal. (1992) 6: 135-160.
    [15] Hydrodynamic limit for the Vlasov-Poisson-Fokker-Planck system: Analysis of the two-dimensional case. Math. Models Methods Appl. Sci. (2005) 15: 737-752.
    [16] Hydrodynamic limit for the Vlasov-Navier-Stokes equation. Part Ⅰ: Light particles regime. Indiana Univ. Math. J. (2004) 53: 1495-1515.
    [17] Hydrodynamic limit for the Vlasov-Navier-Stokes equation. Part Ⅱ: Fine particles regime. Indiana Univ. Math. J. (2004) 53: 1517-1536.
    [18] P. -E. Jabin, private communication.
    [19] Identification of the dilute regime in particle sedimentation. Comm. Math. Phys. (2004) 250: 415-432.
    [20] P. -E. Jabin and B. Perthame, Notes on mathematical problems on the dynamics of dispersed particles interacting through a fluid, in Modeling in Applied Sciences, a Kinetic Theory Approach (eds. N. Bellomo and M. Pulvirenti), Birkhäuser, (2000), 111-147.
    [21] (1990) Polymers in Solution: Their Modelling and Structure.Oxford University Press.
    [22] Calculation of the resistance and mobility functions for two unequal rigid spheres in low-Reynolds-number flow. J. Fluid Mech. (1984) 139: 261-290.
    [23] (1991) Microhydrodynamics: Principles and Selected Applications. Boston: Butterworth-Heinemann.
    [24] J. G. Kirkwood, John Gamble Kirkwood Collected Works: Macromolecules, vol 3, Documents on modern physics, Gordon and Breach, 1967.
    [25] A lower bound for discrimination information in terms of variation. IEEE Trans. Inform. Theory (1967) 13: 126-127.
    [26] Renormalized solutions of some transport equations with partially W^{1, 1} velocities and applications. Ann. Mat. Pura Appl. (2004) 183: 97-130.
    [27] Existence and uniqueness of solutions to Fokker-Planck type equations with irregular coefficients. Comm. Partial Differential Equations (2008) 33: 1272-1317.
    [28] (1964) Information and Information Stability of Random Variables and Processes. San Francisco: Holden-Day.
    [29] Diffusion approximation of the linear semiconductor Boltzmann equation: analysis of boundary layers. Asympot. Anal. (1991) 4: 293-317.
    [30] Charge transport in semiconductors with degeneracy effects. Math. Methods Appl. Sci. (1991) 14: 301-318.
    [31] M. Reichert, Hydrodynamic Interactions in Colloidal and Biological Systems, Ph. D thesis, University Konstanz, 2006.
    [32] H. Risken, The Fokker-Planck Equation. Methods of Solution and Applications, in Springer Series in Synergetics, 18 2nd edition, Berlin, 1989. doi: 10.1007/978-3-642-61544-3
    [33] Variational treatment of hydrodynamic iteractions in polymers. J. Chem. Phys. (1969) 50: 4831-4837.
    [34] S. Varadhan, Entropy methods in hydrodynamic scaling, Proceedings of the International Congress of Mathematicians, Birkhäuser, Basel, 1 (1995), 196-208
    [35] Zur kinetischen Theorie der Brownschen Molekularbewegung und der Suspensionen. Ann. Phys. (1906) 326: 756-780.
    [36] Transport properties of polymer chains in dilute solutions: Hydrodynamic interactions. J. Chem. Phys. (1970) 53: 436-443.
    [37] Relative entropy and hydrodynamics of Ginzburg-Landau models. Lett. Math. Phys. (1991) 22: 63-80.
  • This article has been cited by:

    1. Francesca Anceschi, Yuzhe Zhu, On a spatially inhomogeneous nonlinear Fokker–Planck equation : Cauchy problem and diffusion asymptotics, 2024, 17, 1948-206X, 379, 10.2140/apde.2024.17.379
  • Reader Comments
  • © 2017 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(5447) PDF downloads(168) Cited by(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog