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

Stability estimates for the Vlasov-Poisson system in pp-kinetic Wasserstein distances

Mikaela Iacobelli ETH Zürich, Department of Mathematics, Rämistrasse 101, 8092 Zürich, Switzerland [email protected]  and  Jonathan Junné TU Delft, Delft Institute of Applied Mathematics, Mekelweg 4, 2628 CD Delft, Netherlands [email protected]
Abstract.

We extend Loeper’s L2L^{2}-estimate [13, Theorem 2.9] relating the force fields to the densities for the Vlasov-Poisson system to LpL^{p}, with 1<p<+1<p<+\infty, based on the Helmholtz-Weyl decomposition. This allows us to generalize both the classical Loeper’s 22-Wasserstein stability estimate [13, Theorem 1.2] and the recent stability estimate by the first author relying on the newly introduced kinetic Wasserstein distance [10, Theorem 3.1] to kinetic Wasserstein distances of order 1<p<+1<p<+\infty.

2020 Mathematics Subject Classification:
35Q83, 82C40, 82D10, 35B35

1. Introduction

1.1. General overview

Monge-Kantorovich distances, also known as Wasserstein distances, are used extensively in kinetic theory, in particularly in the context of stability, convergence to equilibrium and mean-field limits. A first celebrated result for the 11-Monge-Kantorovich distance is due to Dobrushin [2, Theorem 1], who proved the well-posedness for Vlasov equations with C1,1C^{1,1} potentials. An explanation of Dobrushin’s stability estimate and its consequences on the mean-field limit for the Vlasov equation can be found in [5, Chapter 1] and [11, Chapter 3], and we refer to [7, Section 3] for a survey on well-posedness for the Vlasov-Poisson system.

Regarding the 22-Wasserstein distance, Loeper proved [13, Theorem 1.2] a uniqueness criterion for solutions with bounded density based on a 22-Wasserstein distance stability estimate using both a link between the H˙1\dot{H}^{-1}-seminorm and the 22-Wasserstein distance, and the fact that the Coulomb kernel is generated by a potential solving the Poisson equation. In addition to the Vlasov-Poisson system, this criterion gives a new proof of uniqueness à la Yudovich for 2D2D Euler. Beyond bounded density, Loeper’s uniqueness criterion has been extended for some suitable Orlicz spaces using the 1-Monge-Kantorovich distance by Miot [16, Theorem 1.1] and Miot, Holding [9, Theorem 1.1].

On the Torus, Loeper’s criterion was improved by Han-Kwan, Iacobelli [8, Theorem 3.1] for the Vlasov-Poisson system, and more recently for the Vlasov-Poisson system with massless electrons by Griffin-Pickering, Iacobelli [6, Theorem 4.1].

The aim of this work is twofold. The first goal is to generalize Loeper’s 22-Wasserstein distance stability estimate to pp-Wasserstein distances for 1<p<+1<p<+\infty. The second goal is to extend the recent stability estimate [10, Theorem 3.1] by the first author relying on the newly introduced kinetic Wasserstein distance [10, Theorem 3.1] to kinetic Wasserstein distances of order 1<p<+1<p<+\infty.

1.2. Definitions and main results

We first recall the classical Wasserstein distance (see [21, Chapter 6]) on the product space 𝒳×d\mathcal{X}\times\mathbb{R}^{d}, with 𝒳\mathcal{X} denoting in the sequel either the dd-dimensional torus 𝕋d\mathbb{T}^{d} or the Euclidean space d\mathbb{R}^{d}:

Definition 1.1.

Let μ,ν\mu,\nu be two probability measures on 𝒳×d\mathcal{X}\times\mathbb{R}^{d}. The Wasserstein distance of order pp, with p1p\geq 1, between μ\mu and ν\nu is defined as

Wp(μ,ν):=(infπΠ(μ,ν)(𝒳×d)2|xy|p+|vw|pdπ(x,v,y,w))1/p,W_{p}(\mu,\nu):=\left(\inf_{\pi\in\Pi(\mu,\nu)}\int_{(\mathcal{X}\times\mathbb{R}^{d})^{2}}\left\lvert x-y\right\rvert^{p}+\left\lvert v-w\right\rvert^{p}\>d\pi(x,v,y,w)\right)^{1/p},

where Π(μ,ν)\Pi(\mu,\nu) is the set of couplings; that is, the set of probability measures with marginals μ\mu and ν\nu. A coupling is said to be optimal if it minimizes the Wasserstein distance.

We consider two solutions f1,f2f_{1},f_{2} of the Vlasov-Poisson system on 𝒳\mathcal{X}, with either gravitational or electrostatic interaction encoded by σ=±1\sigma=\pm 1, namely,

tf+vxfUvf=0,σΔU:=ρf1,ρf:=df𝑑v\partial_{t}f+v\cdot\partial_{x}f-\nabla U\cdot\nabla_{v}f=0,\quad\sigma\Delta U:=\rho_{f}-1,\quad\rho_{f}:=\int_{\mathbb{R}^{d}}f\>dv (1.1)

on the torus, and

tf+vxfUvf=0,σΔU:=ρf,ρf:=df𝑑v\partial_{t}f+v\cdot\partial_{x}f-\nabla U\cdot\nabla_{v}f=0,\quad\sigma\Delta U:=\rho_{f},\quad\rho_{f}:=\int_{\mathbb{R}^{d}}f\>dv (1.2)

on the whole space, with initial profiles f1(0),f2(0)f_{1}(0),f_{2}(0), and respective flows Z1:=(X1,V1)Z_{1}:=(X_{1},V_{1}) and Z2:=(X2,V2)Z_{2}:=(X_{2},V_{2}) satisfying the system of characteristics

X˙=V,V˙=U,X(0,x,v)=x,V(0,x,v)=v.\dot{X}=V,\quad\dot{V}=-\nabla U,\quad X(0,x,v)=x,\quad V(0,x,v)=v.

The flows yield solutions f1(t)=Z1(t,,)#f1(0)f_{1}(t)=Z_{1}(t,\cdot,\cdot)_{\#}f_{1}(0) and f2(t)=Z2(t,,)#f2(0)f_{2}(t)=Z_{2}(t,\cdot,\cdot)_{\#}f_{2}(0) as pushforwards of the initial data.

(1) A new LpL^{p}-estimate for the difference of force fields. Loeper estimates the L2L^{2}-norm [13, Theorem 2.9] of the difference of force fields with the Wasserstein distance between the densities. We extend the L2L^{2}-estimate to LpL^{p} for 1<p<+1<p<+\infty using the Helmholtz-Weyl decomposition of Lp(𝒳)=Gp(𝒳)Hp(𝒳)L^{p}(\mathcal{X})=G_{p}(\mathcal{X})\oplus H_{p}(\mathcal{X}) into its hydrodynamic space that we recall (see [3, Chapter III]):

Definition 1.2.

The hydrodynamic spaces are the closed subspaces of Lp(𝒳)L^{p}(\mathcal{X}) defined as

Gp(𝒳):={uLp(𝒳);u=w for some wWloc1,p(𝒳)}G_{p}(\mathcal{X}):=\left\{u\in L^{p}(\mathcal{X});\quad u=\nabla w\text{ for some }w\in W^{1,p}_{\operatorname{loc}}(\mathcal{X})\right\}

and

Hp(𝒳):={uCc(𝒳);divu=0 on 𝒳}¯.H_{p}(\mathcal{X}):=\overline{\left\{u\in C_{c}^{\infty}(\mathcal{X});\quad\operatorname{div}u=0\text{ on }\mathcal{X}\right\}}.
Remark 1.3.

Note that this decomposition breaks down for p=1p=1 or p=+p=+\infty and does not hold for general domains in LpL^{p}. It is equivalent to the solvability of an Neumann problem (see [3, Lemma III.1.2]), while an orthogonal decomposition in L2L^{2} is always possible, whatever the domain is.

In our setting; that is either on the torus or on the Euclidean space, the Helmholtz-Weyl decomposition holds [3, Theorem III.1.1 & Theorem III.1.2];

Theorem 1.4 (Helmholtz-Weyl decomposition).

The Helmholtz-Weyl decomposition holds for Lp(𝒳)L^{p}(\mathcal{X}), for any 1<p<+1<p<+\infty; that is,

Lp(𝒳)=Gp(𝒳)Hp(𝒳).L^{p}(\mathcal{X})=G_{p}(\mathcal{X})\oplus H_{p}(\mathcal{X}).

Moreover, when p=2p=2, this decomposition is orthogonal.

The validity of the Helmholtz-Weyl decomposition implies the existence of an Helmholtz-Weyl bounded linear projection operator (see [3, Remark III.1.1])

Pp:Lp(𝒳)Hp(𝒳)\displaystyle P_{p}:L^{p}(\mathcal{X})\to H_{p}(\mathcal{X})

with range Hp(𝒳)H_{p}(\mathcal{X}) and with Gp(𝒳)G_{p}(\mathcal{X}) as null space. More precisely, there is a constant CPp>0C_{P_{p}}>0 that only depends on pp and 𝒳\mathcal{X} such that for all uLp(𝒳)u\in L^{p}(\mathcal{X}), it holds

Pp(u)Lp(𝒳)CPpuLp(𝒳).\displaystyle\left\lVert P_{p}(u)\right\rVert_{L^{p}(\mathcal{X})}\leq C_{P_{p}}\left\lVert u\right\rVert_{L^{p}(\mathcal{X})}. (1.3)

Using optimal transport techniques, Loeper manages to link the strong dual homogeneous Sobolev norm and Wasserstein distances between densities, and we recall those notions:

Definition 1.5.

Let 1<p<+1<p<+\infty. The homogeneous Sobolev space is the space

W˙1,p(𝒳):={[g];gWloc1,p(𝒳),gLp(𝒳)},\dot{W}^{1,p}(\mathcal{X}):=\left\{[g];\quad g\in W^{1,p}_{\operatorname{loc}}(\mathcal{X}),\quad\nabla g\in L^{p}(\mathcal{X})\right\},

where []:={+c;c}[\cdot]:=\{\cdot+c;\>c\in\mathbb{R}\} denotes the equivalence class of functions up to a constant, together with the norm

[g]W˙1,p(𝒳):=gLp(𝒳).\left\lVert[g]\right\rVert_{\dot{W}^{1,p}(\mathcal{X})}:=\left\lVert\nabla g\right\rVert_{L^{p}(\mathcal{X})}.

This is a Banach space for which the equivalence classes of test functions

𝒟˙(𝒳):={[ϕ];ϕCc(𝒳)}\dot{\mathcal{D}}(\mathcal{X}):=\{[\phi];\quad\phi\in C^{\infty}_{c}(\mathcal{X})\}

are dense in it (see [17, Theorem 2.1]).

Definition 1.6.

We define the dual homogeneous Sobolev space W˙1,p(𝒳)\dot{W}^{-1,p}(\mathcal{X}) to be the topological dual of W˙1,p(𝒳)\dot{W}^{1,p^{\prime}}(\mathcal{X}) equipped with the strong dual homogeneous Sobolev norm. For a function hh with h=0\int h=0, by density,

hW˙1,p(𝒳)\displaystyle\left\lVert h\right\rVert_{\dot{W}^{-1,p}(\mathcal{X})} :=sup{𝒳h[g]dx;gW˙1,p(𝒳),[g]W˙1,p(𝒳)1}\displaystyle:=\sup\left\{\int_{\mathcal{X}}h\,[g]\>dx;\quad g\in\dot{W}^{1,p^{\prime}}(\mathcal{X}),\quad\left\lVert[g]\right\rVert_{\dot{W}^{1,p^{\prime}}(\mathcal{X})}\leq 1\right\}
=sup{𝒳h[ϕ]dx;ϕ𝒟˙(𝒳),[ϕ]W˙1,p(𝒳)1}.\displaystyle=\sup\left\{\int_{\mathcal{X}}h\,[\phi]\>dx;\quad\phi\in\dot{\mathcal{D}}(\mathcal{X}),\quad\left\lVert[\phi]\right\rVert_{\dot{W}^{1,p^{\prime}}(\mathcal{X})\leq 1}\right\}.

First, we extend this connection for densities to LpL^{p}. Using the machinery of Helmholtz-Weyl decomposition, we generalize [13, Lemma 2.10] into the following:

Lemma 1.7.

Let ρ1,ρ2L(𝒳)\rho_{1},\rho_{2}\in L^{\infty}(\mathcal{X}) be two probability measures, and let UiU_{i} satisfy σΔUi=ρi\sigma\Delta U_{i}=\rho_{i} for 𝒳=d\mathcal{X}=\mathbb{R}^{d}, or σΔUi=ρi1\sigma\Delta U_{i}=\rho_{i}-1 for 𝒳=𝕋d\mathcal{X}=\mathbb{T}^{d}, with i=1,2,σ=±1i=1,2,\sigma=\pm 1, in the distributional sense. Let 1<p<+1<p<+\infty. Then there is a constant CHW>0C_{\mathrm{HW}}>0 that only depends on pp and 𝒳\mathcal{X} such that

U1U2Lp(𝒳)CHWρ1ρ2W˙1,p(𝒳).\left\lVert\nabla U_{1}-\nabla U_{2}\right\rVert_{L^{p}(\mathcal{X})}\leq C_{\mathrm{HW}}\left\lVert\rho_{1}-\rho_{2}\right\rVert_{\dot{W}^{-1,p}(\mathcal{X})}. (1.4)

Second, we adapt Loeper’s argument of the L2L^{2}-estimate [13, Theorem 2.9](see also [20, Proposition 1.1] in bounded convex domains) relating negative homogeneous Sobolev norms to Wasserstein distances with this new link on force fields to get the new LpL^{p}-estimate allowing us to generalize stability estimates;

Proposition 1.8.

Let ρ1,ρ2L(𝒳)\rho_{1},\rho_{2}\in L^{\infty}(\mathcal{X}) be two probability measures, and let UiU_{i} satisfy σΔUi=ρi\sigma\Delta U_{i}=\rho_{i} for 𝒳=d\mathcal{X}=\mathbb{R}^{d}, or σΔUi=ρi1\sigma\Delta U_{i}=\rho_{i}-1 for 𝒳=𝕋d\mathcal{X}=\mathbb{T}^{d}, with i=1,2,i=1,2, σ=±1\sigma=\pm 1, in the distributional sense. Let 1<p<+1<p<+\infty. Then there is a constant CHW>0C_{\mathrm{HW}}>0 that only depends on pp and 𝒳\mathcal{X} such that

U1U2Lp(𝒳)CHWmax{ρ1L(𝒳),ρ2L(𝒳)}1/pWp(ρ1,ρ2).\left\lVert\nabla U_{1}-\nabla U_{2}\right\rVert_{L^{p}(\mathcal{X})}\leq C_{\mathrm{HW}}\max\left\{\left\lVert\rho_{1}\right\rVert_{L^{\infty}(\mathcal{X})},\left\lVert\rho_{2}\right\rVert_{L^{\infty}(\mathcal{X})}\right\}^{1/p^{\prime}}W_{p}(\rho_{1},\rho_{2}). (1.5)

(2) Loeper’s stability estimate in WpW_{p}. Loeper noted [13, Lemma 3.6] that both the Wasserstein distance of order two of the solutions and of the associated densities are bounded by a flow quantity QQ given by

Q(t):=d×d|X1(t,x,v)X2(t,x,v)|2+|V1(t,x,v)V2(t,x,v)|2df0(x,v),Q(t):=\int_{\mathbb{R}^{d}\times\mathbb{R}^{d}}\left\lvert X_{1}(t,x,v)-X_{2}(t,x,v)\right\rvert^{2}+\left\lvert V_{1}(t,x,v)-V_{2}(t,x,v)\right\rvert^{2}\>df^{0}(x,v),

and the bounds read as

W22(f1(t),f2(t))Q(t),W22(ρf1(t),ρf2(t))Q(t).W_{2}^{2}(f_{1}(t),f_{2}(t))\leq Q(t),\quad W_{2}^{2}(\rho_{f_{1}}(t),\rho_{f_{2}}(t))\leq Q(t). (1.6)

Loeper uses the quantity Q(t)Q(t) together with the L2L^{2}-estimate on the force fields to prove the stability estimate [13, Theorem 1.2] leading to the uniqueness of weak solutions. By modifying the quantity Q(t)Q(t) to

Qp(t)\displaystyle Q_{p}(t) :=(𝒳×d)2|X1(t,x,v)X2(t,y,w)|p+|V1(t,x,v)V2(t,y,w)|pdπ0(x,v,y,w)\displaystyle:=\int_{(\mathcal{X}\times\mathbb{R}^{d})^{2}}\left\lvert X_{1}(t,x,v)-X_{2}(t,y,w)\right\rvert^{p}+\left\lvert V_{1}(t,x,v)-V_{2}(t,y,w)\right\rvert^{p}\>d\pi_{0}(x,v,y,w)
=(𝒳×d)2|xy|p+|vw|pdπt(x,v,y,w),\displaystyle=\int_{(\mathcal{X}\times\mathbb{R}^{d})^{2}}\left\lvert x-y\right\rvert^{p}+\left\lvert v-w\right\rvert^{p}\>d\pi_{t}(x,v,y,w),

where πtΠ(f1(t),f2(t))\pi_{t}\in\Pi(f_{1}(t),f_{2}(t)) and π0\pi_{0} is an optimal WpW_{p} coupling (see [6, Section 4] for a construction of πt\pi_{t}), we are able to generalise Loeper’s stability estimate [13, Theorem 1.2], and [8, Theorem 3.1] both on the torus 𝕋d\mathbb{T}^{d} and on the whole space d\mathbb{R}^{d}, to any Wasserstein distance of order pp, with 1<p<+1<p<+\infty;

Theorem 1.9.

Let f1,f2f_{1},f_{2} be two weak solutions to the Vlasov-Poisson system on 𝒳\mathcal{X} with respective densities

ρf1:=df1𝑑v,ρf2:=df2𝑑v.\rho_{f_{1}}:=\int_{\mathbb{R}^{d}}f_{1}\>dv,\quad\rho_{f_{2}}:=\int_{\mathbb{R}^{d}}f_{2}\>dv.

Let 1<p<+1<p<+\infty, and set

A(t):=ρf2(t)L(𝒳)+ρf1(t)L(𝒳)1/pmax{ρf1(t)L(𝒳),ρf2(t)L(𝒳)}1/p,A(t):=\left\lVert\rho_{f_{2}}(t)\right\rVert_{L^{\infty}(\mathcal{X})}+\left\lVert\rho_{f_{1}}(t)\right\rVert_{L^{\infty}(\mathcal{X})}^{1/p}\max\left\{\left\lVert\rho_{f_{1}}(t)\right\rVert_{L^{\infty}(\mathcal{X})},\left\lVert\rho_{f_{2}}(t)\right\rVert_{L^{\infty}(\mathcal{X})}\right\}^{1/p^{\prime}}, (1.7)

which is assumed to be in L1([0,T))L^{1}([0,T)) for some T>0T>0. Then there is a constant CL>0C_{\mathrm{L}}>0 that only depends on pp and dd such that if Wpp(f1(0),f2(0))W_{p}^{p}(f_{1}(0),f_{2}(0)) is sufficiently small so that Wpp(f1(0),f2(0))(4d/e)pW_{p}^{p}(f_{1}(0),f_{2}(0))\leq(4\sqrt{d}/e)^{p} and

|log(Wpp(f1(0),f2(0))(4d)p)|pexp(CL0TA(s)𝑑s)\left\lvert\log\left(\frac{W_{p}^{p}(f_{1}(0),f_{2}(0))}{\left(4\sqrt{d}\right)^{p}}\right)\right\rvert\geq p\exp\left(-C_{\mathrm{L}}\int_{0}^{T}A(s)\>ds\right) (1.8)

then

Wpp(f1(t),f2(t))(4d)pexp{log(Wpp(f1(0),f2(0))(4d)p)exp(CL0tA(s)𝑑s)}.W_{p}^{p}(f_{1}(t),f_{2}(t))\leq\left(4\sqrt{d}\right)^{p}\exp\left\{\log\left(\frac{W_{p}^{p}(f_{1}(0),f_{2}(0))}{\left(4\sqrt{d}\right)^{p}}\right)\exp\left(C_{\mathrm{L}}\int_{0}^{t}A(s)\>ds\right)\right\}. (1.9)

(3) An improved WpW_{p} stability estimate via kinetic Wasserstein distance. Due to the anisotropy between position and momentum variables, we use an adapted Wasserstein distance designed for kinetic problems taking this into account as introduced in [10, Section 4]:

Definition 1.10.

Let μ,ν\mu,\nu be two probability measures on 𝒳×d\mathcal{X}\times\mathbb{R}^{d}. The kinetic Wasserstein distance of order pp, with p1p\geq 1, between μ\mu and ν\nu is defined as

Wλ,p(μ,ν):=(infπΠ(μ,ν)Dp(π,λ))1/p,W_{\lambda,p}(\mu,\nu):=\left(\inf_{\pi\in\Pi(\mu,\nu)}D_{p}(\pi,\lambda)\right)^{1/p},

where Dp(π,λ)D_{p}(\pi,\lambda) is the unique number ss such that

sλ(s)(𝒳×d)2|xy|p𝑑π(x,y,v,w)=(𝒳×d)2|vw|p𝑑π(x,y,v,w),s-\lambda(s)\int_{(\mathcal{X}\times\mathbb{R}^{d})^{2}}\left\lvert x-y\right\rvert^{p}\>d\pi(x,y,v,w)=\int_{(\mathcal{X}\times\mathbb{R}^{d})^{2}}\left\lvert v-w\right\rvert^{p}\>d\pi(x,y,v,w),

with λ:++\lambda:\mathbb{R}^{+}\to\mathbb{R}^{+} a decreasing function.

We consider the quantity Dp(t)D_{p}(t) for πt\pi_{t} and λ(t)=|logDp(t)|p/2\lambda(t)=\left\lvert\log D_{p}(t)\right\rvert^{p/2} (see [10, Lemma 3.7] for the proof of existence) given by

Dp(t)\displaystyle D_{p}(t) :=1p(𝒳×d)2λ(t)|X1(t,x,v)X2(t,y,w)|p+|V1(t,x,v)V2(t,y,w)|pdπ0(x,v,y,w)\displaystyle:=\frac{1}{p}\int_{(\mathcal{X}\times\mathbb{R}^{d})^{2}}\lambda(t)\left\lvert X_{1}(t,x,v)-X_{2}(t,y,w)\right\rvert^{p}+\left\lvert V_{1}(t,x,v)-V_{2}(t,y,w)\right\rvert^{p}\>d\pi_{0}(x,v,y,w)
=1p(𝒳×d)2λ(t)|xy|p+|vw|pdπt(x,v,y,w).\displaystyle=\frac{1}{p}\int_{(\mathcal{X}\times\mathbb{R}^{d})^{2}}\lambda(t)\left\lvert x-y\right\rvert^{p}+\left\lvert v-w\right\rvert^{p}\>d\pi_{t}(x,v,y,w).

This quantity also compares to the usual Wasserstein distance WpW_{p} as does Qp(t)Q_{p}(t), and this allows us to generalize the recent Iacobelli’s stability estimate [10, Theorem 3.1] to the following:

Theorem 1.11.

Let f1,f2f_{1},f_{2} be two weak solutions to the Vlasov-Poisson system on 𝒳\mathcal{X} (1.2) with respective densities

ρf1:=df1𝑑v,ρf2:=df2𝑑v.\rho_{f_{1}}:=\int_{\mathbb{R}^{d}}f_{1}\>dv,\quad\rho_{f_{2}}:=\int_{\mathbb{R}^{d}}f_{2}\>dv.

Let 1<p<+1<p<+\infty, and set

A(t):=ρf2(t)L(𝒳)+ρf1(t)L(𝒳)1/pmax{ρf1(t)L(𝒳),ρf2(t)L(𝒳)}1/p,A(t):=\left\lVert\rho_{f_{2}}(t)\right\rVert_{L^{\infty}(\mathcal{X})}+\left\lVert\rho_{f_{1}}(t)\right\rVert_{L^{\infty}(\mathcal{X})}^{1/p}\max\left\{\left\lVert\rho_{f_{1}}(t)\right\rVert_{L^{\infty}(\mathcal{X})},\left\lVert\rho_{f_{2}}(t)\right\rVert_{L^{\infty}(\mathcal{X})}\right\}^{1/p^{\prime}}, (1.10)

which is assumed to be in L1([0,T))L^{1}([0,T)) for some T>0T>0. Then there is a universal constant c0>0c_{0}>0 and a constant CKW>0C_{\mathrm{KW}}>0 that depends only on pp and dd such that if Wpp(f1(0),f2(0))W_{p}^{p}(f_{1}(0),f_{2}(0)) is sufficiently small so that Wpp(f1(0),f2(0))pc0W_{p}^{p}(f_{1}(0),f_{2}(0))\leq pc_{0} and

|log{Wpp(f1(0),f2(0))|log(1pWpp(f1(0),f2(0)))|}|CKW0TA(s)𝑑s+1,\sqrt{\left\lvert\log\Bigg{\{}W_{p}^{p}(f_{1}(0),f_{2}(0))\left\lvert\log\left(\frac{1}{p}W_{p}^{p}(f_{1}(0),f_{2}(0))\right)\right\rvert\Bigg{\}}\right\rvert}\geq C_{\mathrm{KW}}\int_{0}^{T}A(s)\>ds+1, (1.11)

then

Wpp(f1(t),f2(t))pexp{(|log{Wpp(f1(0),f2(0))|log(1pWpp(f1(0),f2(0)))|}|CKW0tA(s)𝑑s)2}.W_{p}^{p}(f_{1}(t),f_{2}(t))\\ \leq p\exp\Bigg{\{}-\left(\sqrt{\left\lvert\log\Bigg{\{}W_{p}^{p}(f_{1}(0),f_{2}(0))\left\lvert\log\left(\frac{1}{p}W_{p}^{p}(f_{1}(0),f_{2}(0))\right)\right\rvert\Bigg{\}}\right\rvert}-C_{\mathrm{KW}}\int_{0}^{t}A(s)\>ds\right)^{2}\Bigg{\}}. (1.12)

The improvement of this stability estimate (1.12) of Theorem 1.11 via pp-kinetic Wasserstein distance compared to Loeper’s stability estimate in WpW_{p} (1.9) of Theorem 1.9 lies in the order of magnitude of the time interval in which the two solutions are close to each other in Wasserstein distance. Indeed, if Wpp(f1(0),f2(0))=δ1W_{p}^{p}(f_{1}(0),f_{2}(0))=\delta\ll 1, then Loeper’s stability estimate yields Wpp(f1(t),f2(t))1W_{p}^{p}(f_{1}(t),f_{2}(t))\lesssim 1 for t[0,log(|logδ|)]t\in[0,\log(\left\lvert\log\delta\right\rvert)] while the kinetic stability estimates yields a better control of the time interval; Wpp(f1(t),f2(t))1W_{p}^{p}(f_{1}(t),f_{2}(t))\lesssim 1 for t[0,|logδ|]t\in[0,\sqrt{\left\lvert\log\delta\right\rvert}].

2. A new LpL^{p}-estimate via the Helmholtz-Weyl decomposition for 1<p<+1<p<+\infty

2.1. Proof of the LpL^{p}-estimate

Proof of Lemma 1.7.

Let [ϕ]𝒟˙(𝒳)[\phi]\in\dot{\mathcal{D}}(\mathcal{X}) be a quotient test function. Note that ρ1ρ2=0\int\rho_{1}-\rho_{2}=0 as both ρ1\rho_{1} and ρ2\rho_{2} are probability measures, an integration by parts yields

𝒳[ϕ](ρ1ρ2)𝑑x=𝒳ϕ(ρ1ρ2)𝑑x=σ𝒳ϕ(U1U2)𝑑x.\int_{\mathcal{X}}\left[\phi\right]\left(\rho_{1}-\rho_{2}\right)\>dx=\int_{\mathcal{X}}\phi\left(\rho_{1}-\rho_{2}\right)\>dx=-\sigma\int_{\mathcal{X}}\nabla\phi\cdot\left(\nabla U_{1}-\nabla U_{2}\right)\>dx.

First, we consider the torus case 𝒳=𝕋d\mathcal{X}=\mathbb{T}^{d}: We use the Helmholtz-Weyl decomposition given by Theorem 1.4 to write any d\mathbb{R}^{d}-valued test function ΦCc(𝕋d)\Phi\in C_{c}^{\infty}(\mathbb{T}^{d}) as Φ=ϕ+g\Phi=\nabla\phi+g, where ϕGp(𝕋d)\nabla\phi\in G_{p^{\prime}}(\mathbb{T}^{d}) and gHp(𝕋d)g\in H_{p^{\prime}}(\mathbb{T}^{d}) with 1/p+1/p=11/p+1/p^{\prime}=1. By definition, there is a divergence-free sequence of test functions (gk)k(g_{k})_{k\in\mathbb{N}} whose LpL^{p^{\prime}}-limit is gg. By continuity of the force fields (see [8, Lemma 3.2]), U1U2L(𝕋d)\nabla U_{1}-\nabla U_{2}\in L^{\infty}(\mathbb{T}^{d}), and in particular U1U2Lp(𝕋d)\nabla U_{1}-\nabla U_{2}\in L^{p}(\mathbb{T}^{d}). An integration by parts yields

U1U2Lp(𝕋d)\displaystyle\left\lVert\nabla U_{1}-\nabla U_{2}\right\rVert_{L^{p}(\mathbb{T}^{d})} =supΦLp(𝕋d)1{𝕋dΦ(U1U2)𝑑x}\displaystyle=\sup_{\left\lVert\Phi\right\rVert_{L^{p^{\prime}}(\mathbb{T}^{d})}\leq 1}\left\{\int_{\mathbb{T}^{d}}\Phi\cdot\left(\nabla U_{1}-\nabla U_{2}\right)\>dx\right\}
=supΦLp(𝕋d)1{𝕋dϕ(U1U2)𝑑xlimk𝕋ddivgk(U1U2)𝑑x}\displaystyle=\sup_{\left\lVert\Phi\right\rVert_{L^{p^{\prime}}(\mathbb{T}^{d})}\leq 1}\left\{\int_{\mathbb{T}^{d}}\nabla\phi\cdot\left(\nabla U_{1}-\nabla U_{2}\right)\>dx-\lim_{k\to\infty}\int_{\mathbb{T}^{d}}\operatorname{div}g_{k}\left(U_{1}-U_{2}\right)\>dx\right\}
=supϕ;ΦLp(𝕋d)1{𝕋dϕ(U1U2)𝑑x}.\displaystyle=\sup_{\phi;\>\left\lVert\Phi\right\rVert_{L^{p^{\prime}}(\mathbb{T}^{d})}\leq 1}\left\{\int_{\mathbb{T}^{d}}\nabla\phi\cdot\left(\nabla U_{1}-\nabla U_{2}\right)\>dx\right\}.

Since the projection operator Pp:ΦgP_{p^{\prime}}:\Phi\,\mapsto\,g is bounded from Lp(𝕋d)L^{p^{\prime}}(\mathbb{T}^{d}) to Hp(𝕋d)H_{p^{\prime}}(\mathbb{T}^{d}), we have that

ϕLp(𝕋d)=gΦLp(𝕋d)(1+CPp)ΦLp(𝕋d),\left\lVert\nabla\phi\right\rVert_{L^{p^{\prime}}(\mathbb{T}^{d})}=\left\lVert g-\Phi\right\rVert_{L^{p^{\prime}}(\mathbb{T}^{d})}\leq\left(1+C_{P_{p^{\prime}}}\right)\left\lVert\Phi\right\rVert_{L^{p^{\prime}}(\mathbb{T}^{d})},

where CPpC_{P_{p^{\prime}}} is the constant from (1.3), and we set CHW:=1+CPpC_{\mathrm{HW}}:=1+C_{P_{p^{\prime}}}. Consider the larger set

{ϕLp(𝕋d)/CHW1}{ϕ;ϕLp(𝕋d)/CHWΦLp(𝕋d)1}\left\{\left\lVert\nabla\phi\right\rVert_{L^{p^{\prime}}(\mathbb{T}^{d})}/C_{\mathrm{HW}}\leq 1\right\}\supset\left\{\phi;\>\left\lVert\nabla\phi\right\rVert_{L^{p^{\prime}}(\mathbb{T}^{d})}/C_{\mathrm{HW}}\leq\left\lVert\Phi\right\rVert_{L^{p^{\prime}}(\mathbb{T}^{d})}\leq 1\right\}

that does not depend on Φ\Phi anymore. By replacing the supremum over this set, we obtain

U1U2Lp(𝕋d)\displaystyle\left\lVert\nabla U_{1}-\nabla U_{2}\right\rVert_{L^{p}(\mathbb{T}^{d})} CHWsupϕLp(𝕋d)1{𝕋dϕ(U1U2)𝑑x}\displaystyle\leq C_{\mathrm{HW}}\sup_{\left\lVert\nabla\phi\right\rVert_{L^{p^{\prime}}(\mathbb{T}^{d})}\leq 1}\left\{\int_{\mathbb{T}^{d}}\nabla\phi\cdot\left(\nabla U_{1}-\nabla U_{2}\right)\>dx\right\}
=CHWsup[ϕ]Lp(𝕋d)1{𝕋d[ϕ](ρ1ρ2)𝑑x}.\displaystyle=C_{\mathrm{HW}}\sup_{\left\lVert\nabla[\phi]\right\rVert_{L^{p^{\prime}}(\mathbb{T}^{d})}\leq 1}\left\{\int_{\mathbb{T}^{d}}[\phi]\left(\rho_{1}-\rho_{2}\right)\>dx\right\}.

We conclude by density of quotient test functions 𝒟˙(𝕋d)\dot{\mathcal{D}}(\mathbb{T}^{d}) in W˙1,p(𝕋d)\dot{W}^{1,p}(\mathbb{T}^{d}) and by the definition of the strong dual homogeneous Sobolev norm.

Second, we consider the whole space case 𝒳=d\mathcal{X}=\mathbb{R}^{d}: Let φCc(d)\varphi\in C_{c}^{\infty}(\mathbb{R}^{d}) be a test function and set 1/p+1/p=1.1/p+1/p^{\prime}=1. We have that U1U2=σGd(ρ1ρ2)U_{1}-U_{2}=\sigma G_{d}*(\rho_{1}-\rho_{2}) a.e., where GdG_{d} is the fundamental solution of the Laplace equation. Then, by symmetry of the convolution,

xjU1xjU2Lp(d)\displaystyle\left\lVert\partial_{x_{j}}U_{1}-\partial_{x_{j}}U_{2}\right\rVert_{L^{p}(\mathbb{R}^{d})} =supφLp(d)1{dφxjGd(ρ1ρ2)dx}\displaystyle=\sup_{\left\lVert\varphi\right\rVert_{L^{p^{\prime}}(\mathbb{R}^{d})}\leq 1}\left\{\int_{\mathbb{R}^{d}}\varphi\,\partial_{x_{j}}G_{d}*\left(\rho_{1}-\rho_{2}\right)\>dx\right\}
=supφLp(d)1{d(ρ1ρ2)xjGdφdx}.\displaystyle=\sup_{\left\lVert\varphi\right\rVert_{L^{p^{\prime}}(\mathbb{R}^{d})}\leq 1}\left\{\int_{\mathbb{R}^{d}}\left(\rho_{1}-\rho_{2}\right)\partial_{x_{j}}G_{d}*\varphi\>dx\right\}.

We denote ϕ=xjGdφ\phi=\partial_{x_{j}}G_{d}*\varphi, and Calderon-Zygmund’s inequality [3, Theorem II.11.4] yields

ϕLp(d)CHWφLp(d)\left\lVert\nabla\phi\right\rVert_{L^{p^{\prime}}(\mathbb{R}^{d})}\leq C_{\mathrm{HW}}\left\lVert\varphi\right\rVert_{L^{p^{\prime}}(\mathbb{R}^{d})}

for some constant CHW>0C_{\mathrm{HW}}>0 which only depends on dd and pp, so that the supremum can be replaced by the larger set

{ϕW˙1,p(d);ϕLp(d)/CHW1}{φ;xjGdφLp(d)/CHWφLp(d)1}\left\{\phi\in\dot{W}^{1,p^{\prime}}(\mathbb{R}^{d});\>\left\lVert\nabla\phi\right\rVert_{L^{p^{\prime}}(\mathbb{R}^{d})}/C_{\mathrm{HW}}\leq 1\right\}\supset\left\{\varphi;\>\left\lVert\partial_{x_{j}}G_{d}*\varphi\right\rVert_{L^{p^{\prime}}(\mathbb{R}^{d})}/C_{\mathrm{HW}}\leq\left\lVert\varphi\right\rVert_{L^{p^{\prime}}(\mathbb{R}^{d})}\leq 1\right\}

independent of φ\varphi. We obtain

xjU1xjU2Lp(d)\displaystyle\left\lVert\partial_{x_{j}}U_{1}-\partial_{x_{j}}U_{2}\right\rVert_{L^{p^{\prime}}(\mathbb{R}^{d})} CHWsupϕLp(d)1{dϕ(ρ1ρ2)𝑑x}\displaystyle\leq C_{\mathrm{HW}}\sup_{\left\lVert\nabla\phi\right\rVert_{L^{p^{\prime}}(\mathbb{R}^{d})}\leq 1}\left\{\int_{\mathbb{R}^{d}}\phi\left(\rho_{1}-\rho_{2}\right)\>dx\right\}
=CHWsup[ϕ]Lp(d)1{d[ϕ](ρ1ρ2)𝑑x},\displaystyle=C_{\mathrm{HW}}\sup_{\left\lVert\nabla\left[\phi\right]\right\rVert_{L^{p^{\prime}}(\mathbb{R}^{d})}\leq 1}\left\{\int_{\mathbb{R}^{d}}\left[\phi\right]\left(\rho_{1}-\rho_{2}\right)\>dx\right\},

and we conclude by definition of the strong dual homogeneous Sobolev norm. ∎

Before proving our new LpL^{p}-estimate, we first state the existence of an optimal transport map adapted to our context;

Theorem 2.1 (Gangbo-McCann [4, Theorem 1.2]).

Let ρ1,ρ2\rho_{1},\rho_{2} be two probability measures on 𝒳\mathcal{X} that are absolutely continuous with respect to the Lebesgue measure. Then

Wp(ρ1,ρ2)=(infT#ρ1=ρ2{𝒳|xT(x)|p𝑑ρ1(x)})1/p,\displaystyle W_{p}(\rho_{1},\rho_{2})=\left(\inf_{T_{\#}\rho_{1}=\rho_{2}}\bigg{\{}\int_{\mathcal{X}}\left\lvert x-T(x)\right\rvert^{p}\>d\rho_{1}(x)\bigg{\}}\right)^{1/p},

where the infimum runs over all measurable mappings T:𝒳𝒳T:\mathcal{X}\to\mathcal{X} that push forward ρ1\rho_{1} onto ρ2\rho_{2}. Moreover, the infimum is reached by a ρ1(dx)\rho_{1}(dx)-almost surely unique mapping TT, and there is a ||p\left\lvert\cdot\right\rvert^{p}-convex function ψ\psi such that T=Id𝒳(||p)1ψT=\operatorname{Id}_{\mathcal{X}}-\left(\nabla\left\lvert\cdot\right\rvert^{p}\right)^{-1}\circ\nabla\psi, where we denote (h)\left(\nabla h^{*}\right) by (h)1\left(\nabla h\right)^{-1} for a function hh with hh^{*} its Legendre transform.

Proof of Proposition 1.8.

Let us denote by

ρθ:=[(θ1)T+(2θ)Id𝒳]#ρ1\rho_{\theta}:=\left[\left(\theta-1\right)T+\left(2-\theta\right)\operatorname{Id}_{\mathcal{X}}\right]_{\#}\rho_{1}

the interpolant measure between ρ1\rho_{1} and ρ2\rho_{2}, where TT is the optimal transport map of Theorem 2.1. Let ϕCc(𝒳)\phi\in C_{c}^{\infty}(\mathcal{X}) be a test function. By the properties of pushforwards of measures, it follows immediately that

𝒳ϕ(x)𝑑ρθ(x)=𝒳ϕ((θ1)T(x)+(2θ)x)𝑑ρ1(x).\int_{\mathcal{X}}\phi(x)\>d\rho_{\theta}(x)=\int_{\mathcal{X}}\phi\left((\theta-1)T(x)+(2-\theta)x\right)\>d\rho_{1}(x).

Lebesgue’s dominated convergence theorem yields

ddθ𝒳ϕ(x)𝑑ρθ(x)=𝒳ϕ((θ1)T(x)+(2θ)x)(T(x)x)𝑑ρ1(x).\frac{d}{d\theta}\int_{\mathcal{X}}\phi(x)\>d\rho_{\theta}(x)=\int_{\mathcal{X}}\nabla\phi\left((\theta-1)T(x)+(2-\theta)x\right)\cdot\left(T(x)-x\right)\>d\rho_{1}(x).

Now, by using Hölder inequality with respect to the measure ρ1\rho_{1}, we get

ddθ𝒳ϕ(x)ρθ(x)𝑑x\displaystyle\frac{d}{d\theta}\int_{\mathcal{X}}\phi(x)\rho_{\theta}(x)\>dx (𝒳|ϕ((θ1)T(x)+(2θ)x)|p𝑑ρ1(x))1/p(𝒳|xT(x)|p𝑑ρ1(x))1/p\displaystyle\leq\left(\int_{\mathcal{X}}\left\lvert\nabla\phi\left((\theta-1)T(x)+(2-\theta)x\right)\right\rvert^{p^{\prime}}\>d\rho_{1}(x)\right)^{1/p^{\prime}}\left(\int_{\mathcal{X}}\left\lvert x-T(x)\right\rvert^{p}\>d\rho_{1}(x)\right)^{1/p}
=(𝒳|ϕ(x)|p𝑑ρθ(x))1/p(𝒳|xT(x)|p𝑑ρ1(x))1/p.\displaystyle=\left(\int_{\mathcal{X}}\left\lvert\nabla\phi(x)\right\rvert^{p^{\prime}}\>d\rho_{\theta}(x)\right)^{1/p^{\prime}}\left(\int_{\mathcal{X}}\left\lvert x-T(x)\right\rvert^{p}\>d\rho_{1}(x)\right)^{1/p}.

The second term in the product is exactly Wp(ρ1,ρ2)W_{p}(\rho_{1},\rho_{2}) by Theorem 2.1. For the first one, thanks to [18, Remark 8], the LL^{\infty}-norm of the interpolant is controlled by the one of the two measures;

ρθL(𝒳)max{ρ1L(𝒳),ρ2L(𝒳)}.\left\lVert\rho_{\theta}\right\rVert_{L^{\infty}(\mathcal{X})}\leq\max\left\{\left\lVert\rho_{1}\right\rVert_{L^{\infty}(\mathcal{X})},\left\lVert\rho_{2}\right\rVert_{L^{\infty}(\mathcal{X})}\right\}.

Therefore,

ddθ𝒳ϕ(x)ρθ(x)dxmax{ρ1L(𝒳),ρ2L(𝒳)}1/p(𝒳|ϕ(x)|pdx)1/pWp(ρ1,ρ2).\frac{d}{d\theta}\int_{\mathcal{X}}\phi(x)\rho_{\theta}(x)\>dx\leq\max\left\{\left\lVert\rho_{1}\right\rVert_{L^{\infty}(\mathcal{X})},\left\lVert\rho_{2}\right\rVert_{L^{\infty}(\mathcal{X})}\right\}^{1/p^{\prime}}\left(\int_{\mathcal{X}}\left\lvert\nabla\phi(x)\right\rvert^{p^{\prime}}\>dx\right)^{1/p^{\prime}}W_{p}(\rho_{1},\rho_{2}).

Combining the above estimate with the fact that ρ2ρ1=0\int\rho_{2}-\rho_{1}=0 and Fubini’s theorem yields

𝒳[ϕ](x)(ρ2(x)ρ1(x))𝑑x\displaystyle\int_{\mathcal{X}}[\phi](x)\left(\rho_{2}(x)-\rho_{1}(x)\right)\>dx =𝒳ϕ(x)(ρ2(x)ρ1(x))𝑑x\displaystyle=\int_{\mathcal{X}}\phi(x)\left(\rho_{2}(x)-\rho_{1}(x)\right)\>dx
=12(ddθ𝒳ϕ(x)ρθ(x)𝑑x)𝑑θ\displaystyle=\int_{1}^{2}\left(\frac{d}{d\theta}\int_{\mathcal{X}}\phi(x)\rho_{\theta}(x)\>dx\right)d\theta
max{ρ1L(𝒳),ρ2L(𝒳)}1/p(𝒳|[ϕ](x)|pdx)1/pWp(ρ1,ρ2).\displaystyle\leq\max\left\{\left\lVert\rho_{1}\right\rVert_{L^{\infty}(\mathcal{X})},\left\lVert\rho_{2}\right\rVert_{L^{\infty}(\mathcal{X})}\right\}^{1/p^{\prime}}\left(\int_{\mathcal{X}}\left\lvert\nabla[\phi](x)\right\rvert^{p^{\prime}}\>dx\right)^{1/p^{\prime}}W_{p}(\rho_{1},\rho_{2}).

By restricting to quotient test functions [ϕ][\phi] such that [ϕ]Lp(𝒳)1\left\lVert\nabla[\phi]\right\rVert_{L^{p^{\prime}}(\mathcal{X})}\leq 1, we get the strong dual homogeneous norm so that

ρ1ρ2W˙1,p(𝒳)max{ρ1L(𝒳),ρ2L(𝒳)}1/pWp(ρ1,ρ2),\left\lVert\rho_{1}-\rho_{2}\right\rVert_{\dot{W}^{-1,p}(\mathcal{X})}\leq\max\left\{\left\lVert\rho_{1}\right\rVert_{L^{\infty}(\mathcal{X})},\left\lVert\rho_{2}\right\rVert_{L^{\infty}(\mathcal{X})}\right\}^{1/p^{\prime}}W_{p}(\rho_{1},\rho_{2}),

and we conclude by Lemma 1.7. ∎

Remark 2.2.

Loeper uses extensively that the optimal transport map TT is convex to rely on the gas internal energy theory developed by McCann (see [15, Section 2]) to estimate the LL^{\infty}-norm of the interpolant. Here, we only have ||p|\cdot|^{p}-convexity instead, while still the LL^{\infty}-estimate on the interpolant is valid as showed, for instance, by Santambrogio [18, Remark 8]. Loeper gives also an alternative proof [12, Proposition 3.1] using the Benamou-Brenier formula [1, Proposition 1.1]. The interpolant measure satisfies the continuity equation

θρθ+divx(ρθvθ)=0\partial_{\theta}\rho_{\theta}+\operatorname{div}_{x}(\rho_{\theta}v_{\theta})=0

for a vector field vθv_{\theta} related to the Wasserstein distance through

𝒳|vθ(x)|2𝑑ρθ(x)=W22(ρ1,ρ2).\int_{\mathcal{X}}\left\lvert v_{\theta}(x)\right\rvert^{2}\>d\rho_{\theta}(x)=W_{2}^{2}(\rho_{1},\rho_{2}).

Differentiating both sides of Poisson’s equation gives

ΔθUθ=θρθ=divx(ρθvθ),\Delta\partial_{\theta}U_{\theta}=-\partial_{\theta}\rho_{\theta}=\operatorname{div}_{x}(\rho_{\theta}v_{\theta}),

and integrating by parts against θUθ\partial_{\theta}U_{\theta} itself as test function yields

𝒳|θUθ|2𝑑x=𝒳ρθvθθUθdx\int_{\mathcal{X}}\left\lvert\partial_{\theta}\nabla U_{\theta}\right\rvert^{2}\>dx=\int_{\mathcal{X}}\rho_{\theta}v_{\theta}\cdot\partial_{\theta}\nabla U_{\theta}\>dx

so that

θUθL2(𝒳)ρθL(𝒳)1/2W2(ρ1,ρ2)max{ρ1L(𝒳),ρ2L(𝒳)}1/2W2(ρ1,ρ2),\left\lVert\partial_{\theta}\nabla U_{\theta}\right\rVert_{L^{2}(\mathcal{X})}\leq\left\lVert\rho_{\theta}\right\rVert_{L^{\infty}(\mathcal{X})}^{1/2}W_{2}(\rho_{1},\rho_{2})\leq\max\left\{\left\lVert\rho_{1}\right\rVert_{L^{\infty}(\mathcal{X})},\left\lVert\rho_{2}\right\rVert_{L^{\infty}(\mathcal{X})}\right\}^{1/2}W_{2}(\rho_{1},\rho_{2}),

and the conclusion follows after integrating over θ[1,2]\theta\in[1,2].

Even though there is a WpW_{p} version of Benamou-Brenier formula [19, Theorem 5.28], there is no analog test function that allows to mimic this proof for LpL^{p}.

3. Stability estimates revisited for Wasserstein-like distances

3.1. Loeper’s estimate revisited

The proof of Loeper’s stability estimate in WpW_{p} on the torus 𝕋d\mathbb{T}^{d} is similar to [8, Theorem 3.1] using the new LpL^{p}-estimate (1.5) from Proposition 1.8. It relies on the modified quantity (see [6, Section 4])

Qp(t)\displaystyle Q_{p}(t) :=(𝕋d×d)2|X1(t,x,v)X2(t,y,w)|p+|V1(t,x,v)V2(t,y,w)|pdπ0(x,v,y,w)\displaystyle:=\int_{(\mathbb{T}^{d}\times\mathbb{R}^{d})^{2}}\left\lvert X_{1}(t,x,v)-X_{2}(t,y,w)\right\rvert^{p}+\left\lvert V_{1}(t,x,v)-V_{2}(t,y,w)\right\rvert^{p}\>d\pi_{0}(x,v,y,w)
=(𝕋d×d)2|xy|p+|vw|pdπt(x,v,y,w),\displaystyle=\int_{(\mathbb{T}^{d}\times\mathbb{R}^{d})^{2}}\left\lvert x-y\right\rvert^{p}+\left\lvert v-w\right\rvert^{p}\>d\pi_{t}(x,v,y,w),

where πtΠ(f1(t),f2(t))\pi_{t}\in\Pi(f_{1}(t),f_{2}(t)) and π0\pi_{0} is an optimal coupling that satisfies the marginal property;

(𝕋d×d)2ϕ(x,v,y,w)𝑑πt(x,v,y,w)=(𝕋d×d)2ϕ(Z1(t,x,v),Z2(t,y,w))𝑑π0(x,v,y,w),ϕCb((𝕋d×d)2).\int_{\left(\mathbb{T}^{d}\times\mathbb{R}^{d}\right)^{2}}\phi(x,v,y,w)\>d\pi_{t}(x,v,y,w)\\ =\int_{\left(\mathbb{T}^{d}\times\mathbb{R}^{d}\right)^{2}}\phi\left(Z_{1}(t,x,v),Z_{2}(t,y,w)\right)\>d\pi_{0}(x,v,y,w),\quad\forall\phi\in C_{b}((\mathbb{T}^{d}\times\mathbb{R}^{d})^{2}). (3.1)

The last ingredients are the following estimates analog to (1.6) relying on the definition of Wasserstein distance (see [13, Lemma 3.6]):

Wpp(f1(t),f2(t))Qp(t),Wpp(ρf1(t),ρf2(t))Qp(t).W_{p}^{p}(f_{1}(t),f_{2}(t))\leq Q_{p}(t),\quad W_{p}^{p}(\rho_{f_{1}}(t),\rho_{f_{2}}(t))\leq Q_{p}(t). (3.2)

3.2. Kinetic Wasserstein distance revisited

We prove the recent Iacobelli’s stability estimate in WpW_{p} both on the torus and on the whole space adapting the proof of [10, Theorem 3.1].

The proof relies on the modified quantity from the kinetic Wasserstein distance (see [10, Section 4])

Dp(t)\displaystyle D_{p}(t) :=1p(𝒳×d)2λ(t)|X1(t,x,v)X2(t,y,w)|p+|V1(t,x,v)V2(t,y,w)|pdπ0(x,v,y,w)\displaystyle:=\frac{1}{p}\int_{(\mathcal{X}\times\mathbb{R}^{d})^{2}}\lambda(t)\left\lvert X_{1}(t,x,v)-X_{2}(t,y,w)\right\rvert^{p}+\left\lvert V_{1}(t,x,v)-V_{2}(t,y,w)\right\rvert^{p}\>d\pi_{0}(x,v,y,w)
=1p(𝒳×d)2λ(t)|xy|p+|vw|pdπt(x,v,y,w),\displaystyle=\frac{1}{p}\int_{(\mathcal{X}\times\mathbb{R}^{d})^{2}}\lambda(t)\left\lvert x-y\right\rvert^{p}+\left\lvert v-w\right\rvert^{p}\>d\pi_{t}(x,v,y,w),

where λ(t)=|logDp(t)|p/2\lambda(t)=\left\lvert\log D_{p}(t)\right\rvert^{p/2}, whose specific choice of comes from optimisation considerations that will become apparent in the proof. One is able to bound

D˙p(t)λ˙(t)λDp(t)+Dp(t)(λ1/p(t)+λ1/p(t)log(pDp(t)λ(t)))\dot{D}_{p}(t)\lesssim\frac{\dot{\lambda}(t)}{\lambda}D_{p}(t)+D_{p}(t)\left(\lambda^{1/p}(t)+\lambda^{-1/p}(t)\log\left(\frac{pD_{p}(t)}{\lambda(t)}\right)\right)

which can be rewritten as

D˙p(t)Dp(t)[λ1/p(t)+λ1/p(t)log(Dp(t))]\dot{D}_{p}(t)\lesssim D_{p}(t)\Big{[}\lambda^{1/p}(t)+\lambda^{-1/p}(t)\log\left(D_{p}(t)\right)\Big{]}

recalling that λ\lambda is a decreasing function and assuming that |log(pDp(t)/λ(t))||logDp(t)|\left\lvert\log(pD_{p}(t)/\lambda(t))\right\rvert\lesssim\left\lvert\log D_{p}(t)\right\rvert in some regime. The term inside the square brackets is now optimised considering Dp(t)D_{p}(t) as a function of λ(t)\lambda(t).

We recall the Log-Lipschitz estimate on the force fields [8, Lemma 3.2], see also [6, Lemma 3.3] and [14, Lemma 8.1]:

Lemma 3.1.

Let UiU_{i} satisfy σΔUi=ρfi1,i=1,2,σ=±1\sigma\Delta U_{i}=\rho_{f_{i}}-1,\>i=1,2,\sigma=\pm 1, on 𝕋d\mathbb{T}^{d} in the distributional sense. Then there is a constant C>0C>0 such that for all x,y𝕋d,i=1,2x,y\in\mathbb{T}^{d},\>i=1,2, it holds

|Ui(x)Ui(y)|C|xy|log(4d|xy|)ρfi1L(𝕋d).\left\lvert\nabla U_{i}(x)-\nabla U_{i}(y)\right\rvert\leq C\left\lvert x-y\right\rvert\log\left(\frac{4\sqrt{d}}{\left\lvert x-y\right\rvert}\right)\left\lVert\rho_{f_{i}}-1\right\rVert_{L^{\infty}(\mathbb{T}^{d})}. (3.3)
Lemma 3.2.

Let UiU_{i} satisfy σΔUi=ρfi,i=1,2,σ=±1\sigma\Delta U_{i}=\rho_{f_{i}},\>i=1,2,\sigma=\pm 1, on d\mathbb{R}^{d} in the distributional sense. Then there is a constant Cd>0C_{d}>0 such that

UiL(d)Cd(ρfiL1(d)+ρfiL(d)),\left\lVert\nabla U_{i}\right\rVert_{L^{\infty}(\mathbb{R}^{d})}\leq C_{d}\left(\left\lVert\rho_{f_{i}}\right\rVert_{L^{1}(\mathbb{R}^{d})}+\left\lVert\rho_{f_{i}}\right\rVert_{L^{\infty}(\mathbb{R}^{d})}\right),

and for all x,ydx,y\in\mathbb{R}^{d} with |xy|<1/e\left\lvert x-y\right\rvert<1/e, i=1,2i=1,2, it holds

|Ui(x)Ui(y)|Cd|xy|log(1|xy|)(ρfiL1(d)+ρfiL(d)).\left\lvert\nabla U_{i}(x)-\nabla U_{i}(y)\right\rvert\leq C_{d}\left\lvert x-y\right\rvert\log\left(\frac{1}{\left\lvert x-y\right\rvert}\right)\left(\left\lVert\rho_{f_{i}}\right\rVert_{L^{1}(\mathbb{R}^{d})}+\left\lVert\rho_{f_{i}}\right\rVert_{L^{\infty}(\mathbb{R}^{d})}\right). (3.4)

In particular, for all x,ydx,y\in\mathbb{R}^{d},

|Ui(x)Ui(y)|Cd|xy|(1+log(|xy|))(ρfiL1(d)+ρfiL(d)),\left\lvert\nabla U_{i}(x)-\nabla U_{i}(y)\right\rvert\leq C_{d}\left\lvert x-y\right\rvert\left(1+\log^{-}(\left\lvert x-y\right\rvert)\right)\left(\left\lVert\rho_{f_{i}}\right\rVert_{L^{1}(\mathbb{R}^{d})}+\left\lVert\rho_{f_{i}}\right\rVert_{L^{\infty}(\mathbb{R}^{d})}\right), (3.5)

with log(s):=max{log(s),0}\log^{-}(s):=\max\{-\log(s),0\}.

Proof of Theorem 1.11.
D˙p(t)=\displaystyle\dot{D}_{p}(t)= 1p(𝒳×d)2λ˙(t)|X1X2|p𝑑π0\displaystyle\>\frac{1}{p}\int_{(\mathcal{X}\times\mathbb{R}^{d})^{2}}\dot{\lambda}(t)\left\lvert X_{1}-X_{2}\right\rvert^{p}\>d\pi_{0}
+(𝒳×d)2λ(t)|X1X2|p2(X1X2)(V1V2)𝑑π0\displaystyle+\int_{(\mathcal{X}\times\mathbb{R}^{d})^{2}}\lambda(t)\left\lvert X_{1}-X_{2}\right\rvert^{p-2}(X_{1}-X_{2})\cdot(V_{1}-V_{2})\>d\pi_{0}
+(𝒳×d)2|V1V2|p2(V1V2)(xU2(t,X2)xU1(t,X1))𝑑π0.\displaystyle+\int_{(\mathcal{X}\times\mathbb{R}^{d})^{2}}\left\lvert V_{1}-V_{2}\right\rvert^{p-2}(V_{1}-V_{2})\cdot(\nabla_{x}U_{2}(t,X_{2})-\nabla_{x}U_{1}(t,X_{1}))\>d\pi_{0}.

The last two terms are estimated using Hölder’s inequality with respect to the measure π0\pi_{0}, and we have

D˙p(t)1p(𝒳×d)2λ˙(t)|X1X2|p𝑑π0+λ1/p(t)(pDp(t))+(pDp(t))1/p((𝒳×d)2|xU2(t,X2)xU1(t,X1)|p𝑑π0)1/p.\dot{D}_{p}(t)\leq\frac{1}{p}\int_{(\mathcal{X}\times\mathbb{R}^{d})^{2}}\dot{\lambda}(t)\left\lvert X_{1}-X_{2}\right\rvert^{p}\>d\pi_{0}\\ +\lambda^{1/p}(t)\left(pD_{p}(t)\right)+\left(pD_{p}(t)\right)^{1/p^{\prime}}\left(\int_{(\mathcal{X}\times\mathbb{R}^{d})^{2}}\left\lvert\nabla_{x}U_{2}(t,X_{2})-\nabla_{x}U_{1}(t,X_{1})\right\rvert^{p}\>d\pi_{0}\right)^{1/p}. (3.6)

Recall the separation of the difference of force fields;

|xU2(t,X2)xU1(t,X1)||xU2(t,X2)xU2(t,X1)|+|xU1(t,X1)xU2(t,X2)|,\left\lvert\nabla_{x}U_{2}(t,X_{2})-\nabla_{x}U_{1}(t,X_{1})\right\rvert\leq\left\lvert\nabla_{x}U_{2}(t,X_{2})-\nabla_{x}U_{2}(t,X_{1})\right\rvert+\left\lvert\nabla_{x}U_{1}(t,X_{1})-\nabla_{x}U_{2}(t,X_{2})\right\rvert,

whence

((𝒳×d)2|xU2(t,X2)xU1(t,X1)|p𝑑π0)1/pT1(t)+T2(t),\left(\int_{(\mathcal{X}\times\mathbb{R}^{d})^{2}}\left\lvert\nabla_{x}U_{2}(t,X_{2})-\nabla_{x}U_{1}(t,X_{1})\right\rvert^{p}\>d\pi_{0}\right)^{1/p}\leq T_{1}(t)+T_{2}(t), (3.7)

where

T1(t):=((𝒳×d)2|xU2(t,X2)xU2(t,X1)|p𝑑π0)1/p,T2(t):=((𝒳×d)2|xU1(t,X1)xU2(t,X1)|p𝑑π0)1/p.T_{1}(t):=\left(\int_{(\mathcal{X}\times\mathbb{R}^{d})^{2}}\left\lvert\nabla_{x}U_{2}(t,X_{2})-\nabla_{x}U_{2}(t,X_{1})\right\rvert^{p}\>d\pi_{0}\right)^{1/p},\\ T_{2}(t):=\left(\int_{(\mathcal{X}\times\mathbb{R}^{d})^{2}}\left\lvert\nabla_{x}U_{1}(t,X_{1})-\nabla_{x}U_{2}(t,X_{1})\right\rvert^{p}\>d\pi_{0}\right)^{1/p}. (3.8)

First, consider the torus case 𝒳=𝕋d\mathcal{X}=\mathbb{T}^{d}: We estimate T1T_{1} (3.8) using the non-decreasing concave function on [0,+)[0,+\infty) given by

Φp(s):={slogp((4d)ps)if s(4d/e)p,(4pde)pif s>(4d/e)p,\Phi_{p}(s):=\begin{cases}s\log^{p}\left(\frac{(4\sqrt{d})^{p}}{s}\right)&\text{if }s\leq(4\sqrt{d}/e)^{p},\\ \left(\frac{4p\sqrt{d}}{e}\right)^{p}&\text{if }s>(4\sqrt{d}/e)^{p},\end{cases}

together with the Log-Lipschitz estimate (3.3) from Lemma 3.1 and (1.10) to get111Note that, since ρfi(t)0\rho_{f_{i}}(t)\geq 0 and ρfi(t)L(𝕋d)1\left\lVert\rho_{f_{i}}(t)\right\rVert_{L^{\infty}(\mathbb{T}^{d})}\geq 1, then ρfi(t)1L(𝕋d)ρfi(t)L(𝕋d)A(t)\left\lVert\rho_{f_{i}}(t)-1\right\rVert_{L^{\infty}(\mathbb{T}^{d})}\leq\left\lVert\rho_{f_{i}}(t)\right\rVert_{L^{\infty}(\mathbb{T}^{d})}\leq A(t) for i = 1, 2.

T1(t)CA(t)((𝕋d×d)2Φp(|X1X2|p)𝑑π0)1/pT_{1}(t)\leq CA(t)\left(\int_{(\mathbb{T}^{d}\times\mathbb{R}^{d})^{2}}\Phi_{p}\left(\left\lvert X_{1}-X_{2}\right\rvert^{p}\right)\>d\pi_{0}\right)^{1/p}

provided that |X1X2|p(4d/e)p\left\lvert X_{1}-X_{2}\right\rvert^{p}\leq(4\sqrt{d}/e)^{p}, but this is always the case since the distance between points in the torus cannot exceed d\sqrt{d}. Thus by Jensen’s inequality we have

T1(t)\displaystyle T_{1}(t) CA(t)[Φp((𝕋d×d)2|X1X2|p𝑑π0)]1/pCA(t)[Φp(pDp(t)λ(t))]1/p.\displaystyle\leq CA(t)\Bigg{[}\Phi_{p}\left(\int_{(\mathbb{T}^{d}\times\mathbb{R}^{d})^{2}}\left\lvert X_{1}-X_{2}\right\rvert^{p}\>d\pi_{0}\right)\Bigg{]}^{1/p}\leq CA(t)\Bigg{[}\Phi_{p}\left(\frac{pD_{p}(t)}{\lambda(t)}\right)\Bigg{]}^{1/p}.

Now, the considered regime becomes

pDp(t)λ(t)(4d/e)p,\frac{pD_{p}(t)}{\lambda(t)}\leq\left(4\sqrt{d}/e\right)^{p}, (3.9)

so that

T1(t)CA(t)(pDp(t)λ(t))1/p[|log(pDp(t)λ(t))|+plog(4d)].T_{1}(t)\leq CA(t)\left(\frac{pD_{p}(t)}{\lambda(t)}\right)^{1/p}\Bigg{[}\left\lvert\log\left(\frac{pD_{p}(t)}{\lambda(t)}\right)\right\rvert+p\log\left(4\sqrt{d}\right)\Bigg{]}. (3.10)

We replace λ(t)=|logDp(t)|p/2\lambda(t)=\left\lvert\log D_{p}(t)\right\rvert^{p/2}, and consider yet another regime, now dictated by

Dp(t)1e,D_{p}(t)\leq\frac{1}{e}, (3.11)

so that |logDp(t)|1\left\lvert\log D_{p}(t)\right\rvert\geq 1. Note that this regime is compatible with the regime (3.9) needed for the function Φp\Phi_{p} in the sense that if pDp(t)(4d/e)ppD_{p}(t)\leq(4\sqrt{d}/e)^{p} holds, then pDp(t)/|logDp(t)|p/2pDp(t)(4d/e)ppD_{p}(t)/\left\lvert\log D_{p}(t)\right\rvert^{p/2}\leq pD_{p}(t)\leq(4\sqrt{d}/e)^{p}, and since p/e(4d/e)pp/e\leq(4\sqrt{d}/e)^{p}, we can only consider the regime (3.11). We estimate the logarithms in (3.10) in this new regime (3.11) using an elementary inequality valid within this regime;

|log(pDp(t)|logDp(t)|p/2)|2(1+logp+p2)|logDp(t)|,\left\lvert\log\left(\frac{pD_{p}(t)}{\left\lvert\log D_{p}(t)\right\rvert^{p/2}}\right)\right\rvert\leq 2\left(1+\log p+\frac{p}{2}\right)\left\lvert\log D_{p}(t)\right\rvert, (3.12)

and set Cp:=2(1+logp+p/2)C_{p}:=2(1+\log p+p/2). Hence (3.10) becomes

T1(t)CA(t)(pDp(t)λ(t))1/p[Cp|logDp(t)|+plog(4d)].T_{1}(t)\leq CA(t)\left(\frac{pD_{p}(t)}{\lambda(t)}\right)^{1/p}\Big{[}C_{p}\left\lvert\log D_{p}(t)\right\rvert+p\log\left(4\sqrt{d}\right)\Big{]}. (3.13)

We move to the estimation of T2T_{2} (3.8). The LpL^{p}-estimate (1.5) from Proposition 1.8 yields

T2(t)CHWA(t)Wp(ρf1(t),ρf2(t))T_{2}(t)\leq C_{\mathrm{HW}}A(t)W_{p}(\rho_{f_{1}}(t),\rho_{f_{2}}(t)) (3.14)

recalling (1.10). Since (X1(t),X2(t))#π0(X_{1}(t),X_{2}(t))_{\#}\pi_{0} has marginals ρf1(t)\rho_{f_{1}}(t) and ρf2(t)\rho_{f_{2}}(t), we can estimate the Wasserstein distance between the densities by DpD_{p} (see [13, Lemma 3.6]). More precisely, by the definition of the Wasserstein distance, we have

Wpp(ρf1(t),ρf2(t))\displaystyle W_{p}^{p}(\rho_{f_{1}}(t),\rho_{f_{2}}(t)) =infγΠ(ρf1(t),ρf2(t))𝕋d×𝕋d|xy|p𝑑γ(x,y)\displaystyle=\inf_{\gamma\in\Pi(\rho_{f_{1}}(t),\rho_{f_{2}}(t))}\int_{\mathbb{T}^{d}\times\mathbb{T}^{d}}\left\lvert x-y\right\rvert^{p}\>d\gamma(x,y)
𝕋d×𝕋d|xy|pd[(X1(t),X2(t))#π0](x,y)\displaystyle\leq\int_{\mathbb{T}^{d}\times\mathbb{T}^{d}}\left\lvert x-y\right\rvert^{p}\>d\Big{[}\left(X_{1}(t),X_{2}(t)\right)_{\#}\pi_{0}\Big{]}(x,y)
=(𝕋d×d)2|X1(t,x,v)X2(t,y,w)|p𝑑π0(x,v,y,w)pDp(t)λ(t).\displaystyle=\int_{(\mathbb{T}^{d}\times\mathbb{R}^{d})^{2}}\left\lvert X_{1}(t,x,v)-X_{2}(t,y,w)\right\rvert^{p}\>d\pi_{0}(x,v,y,w)\leq\frac{pD_{p}(t)}{\lambda(t)}.

We replace λ(t)=|logDp(t)|p/2\lambda(t)=\left\lvert\log D_{p}(t)\right\rvert^{p/2} and the above estimate in (3.14) to get

T2(t)CHWA(t)(pDp(t)|logDp(t)|p/2)1/p.T_{2}(t)\leq C_{\mathrm{HW}}A(t)\left(\frac{pD_{p}(t)}{\left\lvert\log D_{p}(t)\right\rvert^{p/2}}\right)^{1/p}. (3.15)

Putting altogether estimates (3.6, 3.7, 3.13, 3.15) gives in the considered regime (3.11) that

D˙p(t)1p(𝕋d×d)2D˙p(t)Dp(t)|logDp(t)|p/21|X1X2|p𝑑π0+(pDp(t))[|logDp(t)|+CA(t)(Cp|logDp(t)|(t)+plog(4d)|logDp(t)|)+CHWA(t)|logDp(t)|].\dot{D}_{p}(t)\leq\frac{1}{p}\int_{(\mathbb{T}^{d}\times\mathbb{R}^{d})^{2}}\frac{-\dot{D}_{p}(t)}{D_{p}(t)}\left\lvert\log D_{p}(t)\right\rvert^{p/2-1}\left\lvert X_{1}-X_{2}\right\rvert^{p}\>d\pi_{0}\\ +\left(pD_{p}(t)\right)\Bigg{[}\sqrt{\left\lvert\log D_{p}(t)\right\rvert}+CA(t)\left(C_{p}\sqrt{\left\lvert\log D_{p}(t)\right\rvert}(t)+\frac{p\log\left(4\sqrt{d}\right)}{\sqrt{\left\lvert\log D_{p}(t)\right\rvert}}\right)+\frac{C_{\mathrm{HW}}A(t)}{\sqrt{\left\lvert\log D_{p}(t)\right\rvert}}\Bigg{]}.

If D˙p0\dot{D}_{p}\leq 0, then we do not do anything. Otherwise, then the first term in the above estimate is negative, and therefore

D˙p(t)(pDp(t))[|logDp(t)|+CA(t)(Cp|logDp(t)|(t)+plog(4d)|logDp(t)|)+CHWA(t)|logDp(t)|].\dot{D}_{p}(t)\leq\left(pD_{p}(t)\right)\Bigg{[}\sqrt{\left\lvert\log D_{p}(t)\right\rvert}+CA(t)\left(C_{p}\sqrt{\left\lvert\log D_{p}(t)\right\rvert}(t)+\frac{p\log\left(4\sqrt{d}\right)}{\sqrt{\left\lvert\log D_{p}(t)\right\rvert}}\right)+\frac{C_{\mathrm{HW}}A(t)}{\sqrt{\left\lvert\log D_{p}(t)\right\rvert}}\Bigg{]}.

Since the right-hand side is non-negative, independently of the sign of D˙p\dot{D}_{p}, this bound is always valid in the regime, and using that |logDp(t)|1\left\lvert\log D_{p}(t)\right\rvert\geq 1, together with A(t)1A(t)\geq 1, we get

D˙p(t)C~KWA(t)Dp(t)|logDp(t)|,\dot{D}_{p}(t)\leq\tilde{C}_{\mathrm{KW}}A(t)D_{p}(t)\sqrt{\left\lvert\log D_{p}(t)\right\rvert},

where C~KW:=p×[1+C×(Cp+plog(4d))+CHW]\tilde{C}_{\mathrm{KW}}:=p\,\times\left[1+C\times\left(C_{p}+p\log\left(4\sqrt{d}\right)\right)+C_{\mathrm{HW}}\right]. Therefore,

Dp(t)exp{(|logDp(0)|CKW0tA(s)𝑑s)2},D_{p}(t)\leq\exp\left\{-\left(\sqrt{\left\lvert\log D_{p}(0)\right\rvert}-C_{\mathrm{KW}}\int_{0}^{t}A(s)\>ds\right)^{2}\right\}, (3.16)

where CKW:=C~KW/2C_{\mathrm{KW}}:=\tilde{C}_{\mathrm{KW}}/2 depends only on pp and dd. This implies in particular that (3.11) holds if

|logDp(0)|CKW0TA(s)𝑑s+1.\sqrt{\left\lvert\log D_{p}(0)\right\rvert}\geq C_{\mathrm{KW}}\int_{0}^{T}A(s)\>ds+1. (3.17)

It remains to compare DpD_{p} to the Wasserstein distance between the solutions in the regime (3.11). By the definition of the Wasserstein distance, since (Z1(t),Z2(t))#π0(Z_{1}(t),Z_{2}(t))_{\#}\pi_{0} has marginals f1(t)f_{1}(t) and f2(t)f_{2}(t), we have that

Wpp(f1(t),f2(t))\displaystyle W_{p}^{p}(f_{1}(t),f_{2}(t)) =infγΠ(f1(t),f2(t))(𝕋d×d)2|xy|p𝑑γ(x,v,y,w)\displaystyle=\inf_{\gamma\in\Pi(f_{1}(t),f_{2}(t))}\int_{(\mathbb{T}^{d}\times\mathbb{R}^{d})^{2}}\left\lvert x-y\right\rvert^{p}\>d\gamma(x,v,y,w)
(𝕋d×d)2|xy|p+|vw|pd[(Z1(t),Z2(t))#π0](x,v,y,w)\displaystyle\leq\int_{(\mathbb{T}^{d}\times\mathbb{R}^{d})^{2}}\left\lvert x-y\right\rvert^{p}+\left\lvert v-w\right\rvert^{p}\>d\Big{[}\left(Z_{1}(t),Z_{2}(t)\right)_{\#}\pi_{0}\Big{]}(x,v,y,w)
=(𝕋d×d)2|X1X2|p+|V1V2|pdπ0pDp(t).\displaystyle=\int_{(\mathbb{T}^{d}\times\mathbb{R}^{d})^{2}}\left\lvert X_{1}-X_{2}\right\rvert^{p}+\left\lvert V_{1}-V_{2}\right\rvert^{p}\>d\pi_{0}\leq pD_{p}(t).

For the initial Wasserstein distance, since π0\pi_{0} is optimal, we get

Dp(0)1p|logDp(0)|(𝕋d×d)2|xy|p+|vw|pdπ0=1p|logDp(0)|Wpp(f1(0),f2(0)),D_{p}(0)\leq\frac{1}{p}\left\lvert\log D_{p}(0)\right\rvert\int_{(\mathbb{T}^{d}\times\mathbb{R}^{d})^{2}}\left\lvert x-y\right\rvert^{p}+\left\lvert v-w\right\rvert^{p}\>d\pi_{0}=\frac{1}{p}\left\lvert\log D_{p}(0)\right\rvert W_{p}^{p}(f_{1}(0),f_{2}(0)),

which we rewrite as

Dp(0)|logDp(0)|1pWpp(f1(0),f2(0)).\frac{D_{p}(0)}{\left\lvert\log D_{p}(0)\right\rvert}\leq\frac{1}{p}W_{p}^{p}(f_{1}(0),f_{2}(0)).

Note that, near the origin, the inverse of the function ss/|logs|s\mapsto s/\left\lvert\log s\right\rvert behaves like ττ|logτ|\tau\mapsto\tau\left\lvert\log\tau\right\rvert. In particular, there is a universal constant c0>0c_{0}>0 such that

s|logs|τfor some 0τc0spτ|logτ|.\frac{s}{\left\lvert\log s\right\rvert}\leq\tau\quad\text{for some $0\leq\tau\leq c_{0}$}\quad\implies\quad s\leq p\tau\left\lvert\log\tau\right\rvert.

Hence for sufficiently small initial distance such that Wpp(f1(0),f2(0))pc0W_{p}^{p}(f_{1}(0),f_{2}(0))\leq pc_{0}, then

Dp(0)Wpp(f1(0),f2(0))|log(1pWpp(f1(0),f2(0)))|.D_{p}(0)\leq W_{p}^{p}(f_{1}(0),f_{2}(0))\left\lvert\log\left(\frac{1}{p}W_{p}^{p}(f_{1}(0),f_{2}(0))\right)\right\rvert.

Combining these bounds with (3.16), and recalling (3.17), this implies

Wpp(f1(t),f2(t))pexp{(|log{Wpp(f1(0),f2(0))|log(1pWpp(f1(0),f2(0)))|}|CKW0tA(s)𝑑s)2}W_{p}^{p}(f_{1}(t),f_{2}(t))\\ \leq p\exp\Bigg{\{}-\left(\sqrt{\left\lvert\log\Bigg{\{}W_{p}^{p}(f_{1}(0),f_{2}(0))\left\lvert\log\left(\frac{1}{p}W_{p}^{p}(f_{1}(0),f_{2}(0))\right)\right\rvert\Bigg{\}}\right\rvert}-C_{\mathrm{KW}}\int_{0}^{t}A(s)\>ds\right)^{2}\Bigg{\}}

provided Wpp(f1(0),f2(0))pc0W_{p}^{p}(f_{1}(0),f_{2}(0))\leq pc_{0} and

|log{Wpp(f1(0),f2(0))|log(1pWpp(f1(0),f2(0)))|}|CKW0TA(s)𝑑s+1.\sqrt{\left\lvert\log\Bigg{\{}W_{p}^{p}(f_{1}(0),f_{2}(0))\left\lvert\log\left(\frac{1}{p}W_{p}^{p}(f_{1}(0),f_{2}(0))\right)\right\rvert\Bigg{\}}\right\rvert}\geq C_{\mathrm{KW}}\int_{0}^{T}A(s)\>ds+1.

We conclude the proof of the torus case by [10, Lemma 3.7 & Remark 3.8].

Second, we consider the whole space case 𝒳=d\mathcal{X}=\mathbb{R}^{d}: The only difference lies in the the separation of force fields. We have to estimate T1T_{1} and T2T_{2} defined in (3.8). We split T1T_{1} in two integrals;

T1(t)p=(|X1X2|<1/e𝑑π0+|X1X2|1/e𝑑π0)[|xU2(t,X2)xU2(t,X1)|p]:=I1(t)+I2(t).T_{1}(t)^{p}=\left(\int_{\left\lvert X_{1}-X_{2}\right\rvert<1/e}d\pi_{0}+\int_{\left\lvert X_{1}-X_{2}\right\rvert\geq 1/e}d\pi_{0}\right)\Big{[}\left\lvert\nabla_{x}U_{2}(t,X_{2})-\nabla_{x}U_{2}(t,X_{1})\right\rvert^{p}\Big{]}:=I_{1}(t)+I_{2}(t).

On one hand, for I1I_{1}, using the Log-Lipschitz estimate (3.4) from Lemma 3.2 and (1.10), we get

I1(t)\displaystyle I_{1}(t) CdpAp(t)|X1X2|<1/e|X1X2|plogp(1|X1X2|p)𝑑π0\displaystyle\leq C_{d}^{p}A^{p}(t)\int_{\left\lvert X_{1}-X_{2}\right\rvert<1/e}\left\lvert X_{1}-X_{2}\right\rvert^{p}\log^{p}\left(\frac{1}{\left\lvert X_{1}-X_{2}\right\rvert^{p}}\right)\>d\pi_{0}
CdpAp(t)|X1X2|<1/eΦp(|X1X2|p)𝑑π0.\displaystyle\leq C_{d}^{p}A^{p}(t)\int_{\left\lvert X_{1}-X_{2}\right\rvert<1/e}\Phi_{p}(\left\lvert X_{1}-X_{2}\right\rvert^{p})\>d\pi_{0}.

Applying Jensen’s inequality, we have

I1(t)CdpAp(t)Φp(|X1X2|<1/e|X1X2|p𝑑π0)CdpAp(t)Φp(pDp(t)λ(t)).I_{1}(t)\leq C_{d}^{p}A^{p}(t)\Phi_{p}\left(\int_{\left\lvert X_{1}-X_{2}\right\rvert<1/e}\left\lvert X_{1}-X_{2}\right\rvert^{p}\>d\pi_{0}\right)\leq C_{d}^{p}A^{p}(t)\Phi_{p}\left(\frac{pD_{p}(t)}{\lambda(t)}\right).

On the other hand, for I2I_{2}, the estimate (3.5) from Lemma 3.2 yields

I2(t)CdpAp(t)|X1X2|1/e|X1X2|p𝑑π0CdpAp(t)(pDp(t)λ(t)).I_{2}(t)\leq C_{d}^{p}A^{p}(t)\int_{\left\lvert X_{1}-X_{2}\right\rvert\geq 1/e}\left\lvert X_{1}-X_{2}\right\rvert^{p}\>d\pi_{0}\leq C_{d}^{p}A^{p}(t)\left(\frac{pD_{p}(t)}{\lambda(t)}\right).

Again, we impose the regime Dp(t)1/eD_{p}(t)\leq 1/e, with λ(t)=|logDp(t)|p/2\lambda(t)=\left\lvert\log D_{p}(t)\right\rvert^{p/2}, so that

T1(t)(I1(t)+I2(t))1/p21/pCdA(t)(pDp(t)λ(t))1/p|log(pDp(t)λ(t))+plog(4d)|T_{1}(t)\leq\left(I_{1}(t)+I_{2}(t)\right)^{1/p}\leq 2^{1/p}C_{d}A(t)\left(\frac{pD_{p}(t)}{\lambda(t)}\right)^{1/p}\left\lvert\log\left(\frac{pD_{p}(t)}{\lambda(t)}\right)+p\log\left(4\sqrt{d}\right)\right\rvert

becomes

T1(t)21/pCdA(t)(pDp(t)λ(t))1/p[Cp|logDp(t)|+plog(4d)]T_{1}(t)\leq 2^{1/p}C_{d}A(t)\left(\frac{pD_{p}(t)}{\lambda(t)}\right)^{1/p}\left[C_{p}\left\lvert\log D_{p}(t)\right\rvert+p\log\left(4\sqrt{d}\right)\right]

after using the elementary inequality (3.12) valid within the considered regime. The estimation of T2T_{2} (3.8) is again a direct consequence of the LpL^{p}-estimate (1.5) from Proposition 1.8;

T2(t)CHWA(t)Wp(ρf1(t),ρf2(t)).T_{2}(t)\leq C_{\mathrm{HW}}A(t)W_{p}(\rho_{f_{1}}(t),\rho_{f_{2}}(t)).

From now on, the proof is the same as in the torus case 𝒳=𝕋d\mathcal{X}=\mathbb{T}^{d} with

CKW:=p×[1+21/pCd×(Cp+plog(4d))+CHW].C_{\mathrm{KW}}:=p\,\times\left[1+2^{1/p}C_{d}\times\left(C_{p}+p\log\left(4\sqrt{d}\right)\right)+C_{\mathrm{HW}}\right].

Acknowledgement: The authors would like to thank the anonymous referees for their useful and detailed comments, which improved the presentation of the paper. The second author also acknowledges partial financial support from the Dutch Research Council (NWO): project number OCENW.M20.251.

References

  • [1] Jean Benamou and Yann Brenier “A computational fluid mechanics solution to the Monge-Kantorovich mass transfer problem” In Numerische Mathematik 84.3, 2000, pp. 375–393 DOI: 10.1007/s002110050002
  • [2] Roland L. Dobrushin “Vlasov Equations” In Funktsional Anal.i Prilozhen. 13, 1979, pp. 48–58
  • [3] Giovanni P. Galdi “An Introduction to the Mathematical Theory of the Navier-Stokes Equations”, Springer Monographs in Mathematics New York, NY: Springer New York, 2011 DOI: 10.1007/978-0-387-09620-9
  • [4] Wilfrid Gangbo and Robert J. McCann “The geometry of optimal transportation” In Acta Math. 177.2, 1996, pp. 113–161 DOI: 10.1007/BF02392620
  • [5] F Golse “On the dynamics of large particle systems in the mean field limit” In Macroscopic and Large Scale Phenomena: Coarse Graining, Mean Field Limits and Ergodicity 3, Lect. Notes Appl. Math. Mech. Springer, 2016, pp. 1–144
  • [6] Megan Griffin-Pickering and Mikaela Iacobelli “Global well-posedness for the Vlasov-Poisson system with massless electrons in the 3-dimensional torus” Preprint, 2020 arXiv:1810.06928 [math.AP]
  • [7] Megan Griffin-Pickering and Mikaela Iacobelli “Recent developments on the well-posedness theory for Vlasov-type equations” Preprint In Proceedings of the conference Particle Systems and Partial Differential Equations editions VI, VII and VIII., 2021
  • [8] Daniel Han-Kwan and Mikaela Iacobelli “Quasineutral limit for Vlasov-Poisson via Wasserstein stability estimates in higher dimension” In J. Differential Equations 263.1, 2017, pp. 1–25
  • [9] Thomas Holding and Evelyne Miot “Uniqueness and stability for the Vlasov-Poisson system with spatial density in Orlicz spaces” In Mathematical analysis in fluid mechanics—selected recent results 710, Contemp. Math. Amer. Math. Soc., Providence, RI, 2018, pp. 145–162 DOI: 10.1090/conm/710/14368
  • [10] Mikaela Iacobelli “A New Perspective on Wasserstein Distances for Kinetic Problems” In Arch Rational Mech Anal 244.1, 2022, pp. 27–50 DOI: 10.1007/s00205-021-01705-9
  • [11] Pierre-Emmanuel Jabin “A review of the mean field limits for Vlasov equations” In Kinet. Relat. Models 7.4, 2014, pp. 661–711 DOI: 10.3934/krm.2014.7.661
  • [12] Grégoire Loeper “Uniqueness of the solution to the Vlasov-Poisson system with bounded density” arXiv, 2005 arXiv: http://arxiv.org/abs/math/0504140
  • [13] Grégoire Loeper “Uniqueness of the solution to the Vlasov-Poisson system with bounded density” In J. Math. Pures Appl. (9) 86.1, 2006, pp. 68–79
  • [14] Andrew J. Majda and Andrea L. Bertozzi “Vorticity and Incompressible Flow”, Cambridge Texts in Applied Mathematics Cambridge University Press, 2001
  • [15] Robert J. McCann “A Convexity Principle for Interacting Gases” In Advances in Mathematics 128.1, 1997, pp. 153–179 DOI: 10.1006/aima.1997.1634
  • [16] Evelyne Miot “A uniqueness criterion for unbounded solutions to the Vlasov-Poisson system” In Comm. Math. Phys. 346.2, 2016, pp. 469–482 DOI: 10.1007/s00220-016-2707-7
  • [17] Christoph Ortner and Endre Suli “A Note on Linear Elliptic Systems on d\mathbb{R}^{d} arXiv, 2012 arXiv: http://arxiv.org/abs/1202.3970
  • [18] Filippo Santambrogio “Absolute continuity and summability of transport densities: simpler proofs and new estimates” In Calc. Var. 36.3, 2009, pp. 343–354 DOI: 10.1007/s00526-009-0231-8
  • [19] Filippo Santambrogio “Optimal Transport for Applied Mathematicians” 87, Progress in Nonlinear Differential Equations and Their Applications Springer International Publishing, 2015 DOI: 10.1007/978-3-319-20828-2
  • [20] Filippo Santambrogio “Sharp Wasserstein estimates for integral sampling and Lorentz summability of transport densities” In Journal of Functional Analysis 284.4, 2023, pp. 109783 DOI: https://doi.org/10.1016/j.jfa.2022.109783
  • [21] Cédric Villani “Optimal transport” Old and new 338, Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences] Springer-Verlag, Berlin, 2009, pp. xxii+973 DOI: 10.1007/978-3-540-71050-9