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
[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
∂tf+v⋅∇xf+1m∇v⋅(Ff)=1m2∇v⋅(G(x)∇vf), | (1) |
with the particle mass represented by
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
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
Before we proceed with the details of the limiting approximation, we should note the difficulties in computing the exact formula for friction
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
μOSij={18πη|Rij|(I+ˆRij⊗ˆRij),i≠j16πηaI,i=j, |
where
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
μRPYij={18πη|Rij|[(1+2a23|Rij|2)I+(1−2a2|Rij|2)ˆRij⊗ˆRij],|Rij|>2a16πηa[(1−9|Rij|32a)I+3|Rij|32aˆRij⊗ˆRij],|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
We should note that the exact computation of the eigenvalues of the RPY mobility for
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
∂tfϵ+Lϵfϵ=0,fϵ(0,x,v)=f0,ϵ(x,v),withLϵ=1ϵ(v⋅∇xfϵ−∇V(x)⋅∇vfϵ)−1ϵ2∇v⋅(G(x)(∇vfϵ+vfϵ)). | (2) |
Our main objective is to study the (zero mass) limit
M(v)=e−|v|22/(2π)n2 | (3) |
and
We give two results of convergence which are discussed in Section 1.2. First, we show in Theorem 1.1 that
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
F(x,v)∫σ(x,v,ω)dμ(ω)=∫σ(x,v,ω)F(x,ω)dμ(ω)a.e. |
The collision kernel
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
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
Meq(x,v)=e−V(x)M(v)/Z,withZ=(2π)n2∫e−V(x)dx. | (4) |
We also consider
In the first theorem, we establish weak convergence of the hydrodynamic variable
‖fϵ(0,x,v)‖L2Meq<C∀ϵ>0,for someC>0. | (5) |
We prove in Section 3 the following theorem.
Theorem 1.1. Let
ρϵ⇀ρinC([0,T],w−L2(dx)), |
where
∂tρ=∇x⋅(G−1(∇xρ+∇V(x)ρ))inC([0,T],D′(Rnx)). | (6) |
In the second theorem, we use the relative entropy functional to prove an
H(f|g)=∬flogfgdvdx, | (7) |
and in the present work it will be used to control the distance of a solution
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
‖∇kG−1‖L∞(Rnx)<∞,‖∇k(G−1∇V(x))‖L∞(Rnx)<∞,1≤k≤3. | (A1) |
We also assume that the initial condition
ae−V(x)≤ρ(0,x)≤Ae−V(x)for someA>a>0andρ(0,x)/e−V(x)∈W3,∞(Rnx).} | (A2) |
Finally, the use of the maximum principle for the parabolic equation (6) in
sup0≤t≤Tlim supx→∞|∇kρ(t,x)e−V(x)|≤CkforCk>0,0≤k≤3, | (A3) |
where
Theorem 1.2. Let
supϵ>0∬fϵ(0,x,v)(1+V(x)+|v|2+logfϵ(0,x,v))dvdx<C<∞. | (8) |
Moreover, we assume that
∫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))→0asϵ→0. |
Then, for any
sup0≤t≤TH(fϵ(t,⋅,⋅)|ρ(t,⋅)M(v))→0asϵ→0. |
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
ρϵ(t,x):=∫fϵdv,Jϵ(t,x):=∫vfϵdv,Pϵ(t,x):=∫v⊗vfϵdv. | (9) |
In the study of the limit
First, integrating (2) in velocity space, we obtain
∂tρϵ+1ϵ∇x⋅Jϵ=0. | (10) |
We want to derive an expression for the evolution of
ϵ2∂tJϵ(t,x)+ϵ(∇x⋅Pϵ(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+∫M∂vi(fϵM)vjdv, |
which implies
Pϵ(t,x)=ρϵI+∫M∇v(fϵM)⊗vdv. | (12) |
With the help of (12), equation (11) now gives
Jϵ=−ϵG−1(x)(∇xρϵ+∇V(x)ρϵ)−ϵ2G−1(x)∂tJϵ−ϵG−1(x)∇x⋅∫M∇v(fϵM)⊗vdv. | (13) |
The last term in (13) contains the part
Jϵ(t,x)=−ϵG−1(x)(∇xρϵ+∇V(x)ρϵ)+ϵ2…. | (14) |
Finally, as we let
∂tρ+∇x⋅J=0J=−G−1(x)(∇xρ+∇V(x)ρ), |
where
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
In terms of the type of solutions we work with, we shall assume that the operator
Definition 2.1. A mild-weak solution
fϵ∈C(R+;D′(Rnx×Rnv))∩L∞loc(R+;L2Meq∩L∞) | (15) |
and satisfies
∬fϵ(T,⋅,⋅)φ(⋅,⋅)dvdx−∬fϵ(0,⋅,⋅)φ(⋅,⋅)dvdx−1ϵ∫T0∬(v⋅∇xφ−∇V(x)⋅∇vφ)fϵdvdxds+1ϵ2∫T0∬∇vφ⋅G(x)(∇vfϵ+vfϵ)dvdxds=0, |
for any test function
For the second result, we use the relative entropy of
‖f−g‖L1≤√2H(f|g). |
Thus, by finding
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
X:={fϵ|fϵ∈L∞([0,T];L1∩L∞)&G1/2∇vfϵ∈(L2([0,T],L2))n}, | (16) |
for all times
∬fϵ(T,⋅,⋅)φ(T,⋅,⋅)dvdx−∬fϵ(0,⋅,⋅)φ(0,⋅,⋅)dvdx−∫T0∬fϵ∂tφdvdxds−1ϵ∫T0∬(v⋅∇xφ−∇V(x)⋅∇vφ)fϵdvdxds+1ϵ2∫T0∬∇vφ⋅G(x)(∇vfϵ+vfϵ)dvdxds=0, |
for any test function
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
Proposition 1. (see [27]) Assume that the potential
(ⅰ)G(x)v+∇V(x)∈(W1,1loc(Rnx×Rnv))n(ⅱ)tr(G)∈L∞(Rnx) |
(ⅲ)G(x)v+∇V(x)1+|x|+|v|∈(L∞(Rnx×Rnv))n |
(ⅳ)G1/2(x)∈(W1,2loc(Rnx))n×n(ⅴ)G1/2(x)1+|x|∈(L∞(Rnx))n×n. |
Then, given initial data
In this section we collect all the convergence results needed for the proof of Theorem 1.1. We begin with the decomposition of
fϵ=M(v)(ρϵ+˜gϵ), |
where the hydrodynamic variable
∫˜gϵM(v)dv=0. |
We also note that integrating (2) in velocity we obtain the hydrodynamic equation for
∂tρϵ+1ϵ∇x⋅∫M(v)∇v˜gϵdv=0. | (17) |
We prove the following.
Lemma 3.1. Assume a mild-weak solution
ρϵi⇀ρweaklyinL2(dx)∀t≥0,˜gϵi⇀˜gweaklyinL2(M(v)dvdx)∀t≥0,1ϵiG1/2∇v˜gϵi⇀JweaklyinL2(M(v)dvdxdt). |
Proof. In order to study the limit
12∫h2ϵ(t,x,v)dμ+1ϵ2∫t0∫|G1/2(x)∇vhϵ(s,x,v)|2dμds=12∫h2ϵ(0,x,v)dμ. | (18) |
To simplify the analysis, we consider the basic assumption
∫ρ2ϵdx<∞,∬˜g2ϵM(v)dvdx<∞∀t≥0. | (19) |
For the first bound in (19) we used a simple Jensen inequality on the
1ϵ2∫T0∬|G1/2∇v˜gϵ|2M(v)dvdxds<∞for anyT>0. | (20) |
Based on equations (19) & (20), and after picking a sequence
It is important to comment that we want something stronger than just
Lemma 3.2. Under the assumptions of Theorem 1.1,
ρϵ⇀ρinC([0,T],w−L2(dx)). |
Proof Consider the functional
Now, if we consider
H(t2)−H(t1)=∫ϕ(x)ρϵ(t,x)|t=t2t=t1dx(Use weak form of (17))=1ϵ∫t2t1∬∇xϕ(x)⋅∇v˜gϵM(v)dvdxds≤(∫t2t1∬|G−1/2∇xϕ(x)|2M(v)dvdxds)12(∫t2t1∬|G1/2∇v˜gϵ|2ϵ2M(v)dvdxds)12≤(t2−t1)12(∫|G−1/2∇xϕ(x)|2dx)12(∫t2t1∬|G1/2∇v˜gϵ|2ϵ2M(v)dvdxds)12≤C(t2−t1)12. |
The Arzelá-Ascoli theorem states that pointwise boundedness and equicontinuity suffice to show that the family
Next, we use a standard density argument to show that
∫ϕ(x)ρϵj(t,x)dx→∫ϕ(x)ρ(t,x)dxasj→∞, |
for any
We close by approximating any function
∫ρϵ(t,x)(ϕ(x)−ϕm(x))dx→0asm→0, |
uniformly in
Now that weak compactness of
ϵ∂t˜gϵ−∇x⋅∫M∇v˜gϵdv+v⋅(∇x(ρϵ+˜gϵ)+∇V(x)(ρϵ+˜gϵ))−∇V(x)⋅∇v˜gϵ=1ϵ1M∇v⋅(MG(x)∇v˜gϵ). | (21) |
A mild solution of (21) will be in
ϵ∬M(v)φ(˜gϵ(t2)−˜gϵ(t1))dvdx+∫t2t1∬M(v)∇xφ⋅(∫M(v′)∇v′˜gϵdv′)dvdxds+∫t2t1∬M(v)v⋅(−∇xφρϵ+φ∇V(x)ρϵ)dvdxds+∫t2t1∬M(v)(−∇xφ⋅∇v˜gϵ+φ∇V(x)⋅∇v˜gϵ)dvdxds−∫t2t1∬M(v)φ∇V(x)⋅∇v˜gϵdvdxds=−1ϵ∫t2t1∬M(v)∇vφ⋅G∇v˜gϵdvdxds, | (22) |
where
Lemma 3.3. Under the assumptions of Theorem 1.1, in the limit
∫t2t1∬M(v)v⋅(−∇xφρ+φ∇V(x)ρ)dvdxds=−∫t2t1∬M(v)∇vφ⋅G1/2Jdvdxds∀φ∈C∞c(Rnx×Rnv). | (23) |
Proof We use the notation
I1=ϵ∬M(v)φ(˜gϵ(t2)−˜gϵ(t1))dvdx≤ϵ(∬φ2M(v)dvdx)1/2(∬(|˜gϵ(t2)|2+|˜gϵ(t1)|2)M(v)dvdx)1/2≤Cϵ=O(ϵ). |
I2=∫t2t1∭M(v)M(v′)∇xφ(x,v,s)⋅∇v′˜gϵ(x,v′,s)dv′dvdxds≤ϵ∫t2t1∭M(v)M(v′)|G−1/2∇xφ||G1/2∇v′˜gϵ|ϵdv′dvdxds≤ϵ(∫t2t1∬M(v)|G−1/2∇xφ|2dvdxds)1/2×(1ϵ2∫t2t1∬M(v′)|G1/2∇v′˜gϵ|2dv′dxds)1/2≤Cϵ=O(ϵ). |
I3=∫t2t1∬M(v)v⋅(−∇xφρϵ+φ∇V(x)ρϵ)dvdxds≤(∫t2t1∬M(v)|v|2(|∇xφ|2+|φ∇V(x)|2)dvdxds)1/2(∫t2t1∫ρ2ϵdxds)1/2≤C(∬M(v)|v|2(|∇xφ|2+|φ∇V(x)|2)dx)1/2(∫t2t1∫ρ2ϵdxds)1/2=O(1). |
I4=∫t2t1∬M(v)(−∇xφ⋅∇v˜gϵ+φ∇V(x)⋅∇v˜gϵ)dvdxds≤ϵ(∫t2t1∬M(v)(|G−1/2∇xφ|2+|φG−1/2∇V(x)|2)dvdxds)1/2×(∫t2t1∬1ϵ2|G1/2∇v˜gϵ|2M(v)dvdxds)1/2≤Cϵ=O(ϵ). |
I5=−∫t2t1∬M(v)φ∇V(x)⋅∇v˜gϵdvdxds≤ϵ(∫t2t1∬M(v)|φG−1/2∇V(x)|2dvdxds)1/2×(∫t2t1∬1ϵ2|G1/2∇v˜gϵ|2M(v)dvdxds)1/2≤Cϵ=O(ϵ). |
I6=1ϵ∫t2t1∬M(v)∇vφ⋅G∇v˜gϵdvdxds≤(∫t2t1∬|G1/2∇vφ|2M(v)dvdxds)1/2×(∫t2t1∬1ϵ2|G1/2∇v˜gϵ|2M(v)dvdxds)1/2=O(1). |
Now that we have established all the above bounds, we take
Proof of Theorem 1.1. We write the hydrodynamic equation (17) for
∫ϕ(⋅)ρ(t,⋅)|t=t2t=t1dx=∫t2t1∬M(v)∇xϕ⋅G−1/2Jdvdxds∀ϕ∈C∞c(Rnx). | (24) |
In order to give the limiting equation for
We begin by taking the cut-off function
ηδ2(v)=1δn2η(vδ2)forη∈C∞c(Rnv)such that∫η(v)dv=1. |
We now take the function
By the substitution of
∫t2t1∬M(v)v⋅(−∇xφδ1,δ2ρ+φδ1,δ2∇V(x)ρ)dvdxds=−∫t2t1∬M(v)∇vφδ1,δ2⋅G1/2Jdvdxds. |
We also have
∇vφδ1,δ2(x,v)=∇v((∇xϕ⋅G−1vχδ1(v))⋆ηδ2)=∇v(∇xϕ⋅G−1vχδ1(v))⋆ηδ2=(∇xϕ⋅G−1χδ1(v)+∇xϕ⋅G−1v∇vχδ1(v))⋆ηδ2, |
where we use the fact that
A typical estimate for
∫ϕ(⋅)ρ(t,⋅)|t=t2t=t1dx=∫t2t1∫(∇x⋅(G−1∇xϕ)+∇xϕ⋅G−1∇V(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
The following proposition contains all the estimates needed for a solution
Proposition 2. Assume
(ⅰ)
(ⅱ)
(ⅲ)
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)(v√fϵ+2∇v√fϵ)=2√MG1/2(x)∇v√fϵM. |
The dissipation of energy implies
E(fϵ(T,⋅,⋅))+1ϵ2∫T0∬|dϵ|2dvdxds=E(fϵ(0,⋅,⋅)). | (25) |
(ⅱ) It also follows from (25) that
∫T0∬|dϵ|2dvdxdt≤Cϵ2,T>0. |
(ⅲ) We now give the bound for
ab≤h(a)+h∗(b), |
where
14∬|v|2fϵdvdx≤∬fϵlogfϵMeqdvdx+∬e|v|24−1Meqdvdx. |
This implies that
∬|v|2fϵ(t,v,x)dvdx≤Cfor someC>0,∀t∈[0,T], |
since
We now prove the following result for the evolution of the relative entropy.
Lemma 4.1. Assume a sufficiently regular solution
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}(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
\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}= \! - \! \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
At this point, we compute the evolution of
\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
\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
\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
We already gave the formal computation for
\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
In the process of controlling the two terms, we introduce a new notation for expressions involving the hydrodynamic variable
\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
Stability estimates for the terms
The control of terms
Lemma 4.2. Let
Proof. First, we define
The time evolution of
\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
\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)
For instance, for
\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
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
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
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
The equation for the regularized
\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
\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
\partial_{t}\rho_{\epsilon,\delta}+\frac{1}{\epsilon}\nabla_{x} \cdot J_{\epsilon,\delta}= \int U^{1}_{\epsilon,\delta}\, dv , | (37) |
where
The regularized limiting equation for
\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
Next, multiplying (36) by
\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
For this reason, we introduce a relative entropy integral with a cut-off
\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
\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
If we consider
A bound for the first term in
\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
\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
The last term in
\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
In the last step, we send
\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
\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
Remark 3. The regularization procedure above was carried out for the
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
\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. ![]() |
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 |