1: \documentclass[a4paper, 12pt,leqno]{article}
2: %\usepackage{amssymb,amsmath}
3: \usepackage{amsmath,amsfonts,here, epsf}
4: \usepackage[latin1]{inputenc}
5: \usepackage[T1]{fontenc}
6: \usepackage{ae,aecompl}
7: %\usepackage{pslatex}
8:
9:
10:
11: \begin{document}
12: \newtheorem{theo}{Theorem}[section]
13: \newtheorem{lemme}[theo]{Lemma}
14: \newtheorem{cor}[theo]{Corollary}
15: \newtheorem{defi}[theo]{Definition}
16: \newtheorem{problem}[theo]{Problem}
17: \newtheorem{prop}[theo]{Proposition}
18: \newtheorem{assu}[theo]{Assumption}
19: \newtheorem{nontheo}[theo]{Conjectured theorem}
20: \newcommand{\beq}{\begin{eqnarray}}
21: \newcommand{\enq}{\end{eqnarray}}
22: \newcommand{\be}{\begin{eqnarray*}}
23: \newcommand{\en}{\end{eqnarray*}}
24: \newcommand{\Td}{\mathbb T^d}
25: \newcommand{\T}{\mathbb T}
26: \newcommand{\R}{\mathbb R}
27: \newcommand{\N}{\mathbb N}
28: \newcommand{\Rd}{\mathbb R^d}
29: \newcommand{\Zd}{\mathbb Z^d}
30: \newcommand{\Linf}{L^{\infty}}
31: \newcommand{\dt}{\partial_t}
32: \newcommand{\Dt}{\frac{d}{dt}}
33: \newcommand{\demi}{\frac{1}{2}}
34: \newcommand{\ep}{^{\epsilon}}
35: \newcommand{\epu}{_{\epsilon}}
36: \newcommand{\Dtt}{\frac{d^2}{dt^2}}
37: \newcommand{\vf}{\varphi}
38: \newcommand{\bfi}{{\mathbf \Phi}}
39: \newcommand{\bpsi}{{\mathbf \Psi}}
40: \newcommand{\bm}{{\mathbf m}}
41: \newcommand{\NN}{\mathbb N}
42: \newcommand{\RR}{\mathbb R}
43: \newcommand{\dx}{\partial_x}
44: \newcommand{\vp}{v^{\perp}}
45: \newcommand{\E}{{\mathbf E}}
46: \newcommand{\Sz}{{\mathcal{S}}}
47: \newcommand{\ds}{\displaystyle}
48: \newcommand{\fe}{f_\epsilon}
49: %\let\cal=\mathcal
50: \bibliographystyle{plain}
51:
52:
53: \title{A geometric approximation to the Euler equations :
54: the Vlasov-Monge-Amp\`ere system}
55: \author{Yann Brenier, Gr\'egoire Loeper,
56: \\UMR 6621, Parc Valrose, 06108 Nice, France}
57: \maketitle
58:
59:
60: \begin{abstract}
61: This paper studies the Vlasov-Monge-Amp\`ere system ($VMA$),
62: a fully non-linear version of the Vlasov-Poisson system ($VP$)
63: where the (real) Monge-Amp\`ere equation
64: $\det \frac{\partial^2 \Psi}{\partial x_i \partial x_j}=\rho$
65: substitutes for the usual Poisson equation.
66: This system can be derived as a geometric approximation of the Euler
67: equations of incompressible fluid mechanics in the spirit of Arnold
68: and Ebin. Global
69: existence of weak solutions and local existence of smooth solutions
70: are obtained. Links between the $VMA$ system, the $VP$ system and the
71: Euler equations are established through rigorous asymptotic analysis.
72: \end{abstract}
73:
74:
75: \section{Introduction}
76:
77: The classical Vlasov-Poisson ($VP$) system describes the evolution of
78: an electronic cloud in a neutralizing uniform background
79: through the following equations
80: \beq
81: &&\frac{\partial f}{\partial t}+ \xi\cdot\nabla_x f
82: +\nabla_x\varphi\cdot\nabla_{\xi} f=0\label{1introprincipale}\\
83: && \epsilon^2\Delta\varphi=\rho-1\label{1intropoisson},
84: \enq
85: where $f(t,x,\xi)\ge 0$
86: denotes the electronic density at time $t\ge 0$,
87: point $x\in{\mathbb R}^d$, velocity $\xi\in{\mathbb R}^d$
88: (usually $d=3$), $\rho(t,x)\ge 0$ denotes the 'macroscopic' density
89: \beq\label{1defrho}
90: \rho(t,x)=\int_{{\mathbb R}^d}f(t,x,\xi)d\xi, \label{1rho}
91: \enq
92: and $\varphi(t,x)$ denotes the electric potential at time $t$ and point
93: $x$ generated, through the Poisson equation (\ref{1intropoisson}),
94: where $\epsilon$ is a coupling constant,
95: by the difference between the electronic density $\rho(t,x)$
96: and the neutralizing background density, which is supposed to be uniform
97: and normalized to unity.
98: Standard notations $\nabla=(\partial_1,...,\partial_d)$
99: and $\Delta=\partial_1^2+...+\partial_d^2$ have been used and~$\cdot$
100: stands for the inner product in ${\mathbb R}^d$.
101: The mathematical theory of the $VP$ system is now well understood.
102: In particular, existence of global
103: smooth solutions in three space dimensions has been proved in \cite{Pf}
104: (see also \cite{LP}, \cite{Sc}).
105: In the present paper, a fully nonlinear version of the $VP$ system is
106: addressed~:
107: \beq
108: &&\frac{\partial f}{\partial t}+ \xi\cdot\nabla_x f
109: +\nabla_x\varphi\cdot\nabla_{\xi} f=0\\
110: && \det (\mathbb I + \epsilon^2 D^2 \varphi )=\rho\label{1intromonge},
111: \enq
112: where the (real) Monge-Amp\`ere equation (\ref{1intromonge})
113: substitutes for the Poisson equation (\ref{1intropoisson}).
114: Here, $D^2\varphi(t,x)$ stands for the $d\times d$ symmetric matrix
115: made of all second order $x-$partial derivatives of $\varphi$,
116: $\mathbb I$ stands for the $d\times d$ identity matrix and $\det$
117: for the determinant of a square matrix.
118: The occurrence of the Monge-Amp\`ere equation in mathematical modeling
119: is not very common. Notice, however, that a very similar system
120: can be found in meteorology with Hoskins' semi-geostrophic equations
121: (cf. \cite{BB2}, \cite{CuG} and the included references). In a simplified
122: two dimensional setting, the semi-geostrophic equations read
123: \beq
124: &&\frac{\partial \rho}{\partial t}+\{\varphi,\rho\}=0
125: \\
126: && \det (\mathbb I + \epsilon^2 D^2 \varphi )=\rho\label{semigeo},
127: \enq
128: where $\{\cdot,\cdot\}$ denotes the usual Poisson bracket.
129: \\
130: Formally, as the coupling constant $\epsilon$ is small, the $VP$ and $VMA$
131: equations asymptotically approach each other up to order $O(\epsilon^4)$.
132: Indeed, linearizing the determinant about the identity matrix leads
133: to
134: \beq
135: \det (\mathbb I + \epsilon^2 D^2 \varphi )=1+\epsilon^2\Delta\varphi
136: +O(\epsilon^4).
137: \enq
138: The formal limit, as $\epsilon=0$, reads
139: \beq
140: \label{extended euler}
141: &&\frac{\partial f}{\partial t}+ \xi\cdot\nabla_x f
142: +\nabla_x\varphi\cdot\nabla_{\xi} f=0\\
143: && \rho=1\label{1introcontrainte},
144: \enq
145: where constraint (\ref{1introcontrainte}) substitutes for both the Poisson
146: and the Monge-Amp\`ere equations. The limit system
147: (\ref{extended euler},\ref{1introcontrainte}), that we call constrained
148: Vlasov system, can be seen as a
149: 'kinetic' extension of the Euler equations of classical incompressible
150: fluid mechanics,
151: \beq
152: &&\dt v +(v\cdot\nabla) v=-\nabla p\label{1euler}\\
153: &&\nabla\cdot v =0\label{1div=0},
154: \enq
155: where $v(t,x)\in {\mathbb R}^d$ and $p(t,x)\in {\mathbb R}$
156: respectively are the velocity and the pressure of the fluid at time $t$
157: and position $x$.
158: Indeed, any smooth solution $(v,p)$ provides a 'monokinetic' solution to
159: the constrained Vlasov system (\ref{extended euler},\ref{1introcontrainte}),
160: defined by
161: \[
162: f(t,x,\xi)=\delta(\xi-v(t,x)),\;\;\; \varphi=-p.
163: \]
164: Here a monokinetic solution means a delta-valued solution
165: in the $\xi$ variable.
166: In addition, the constrained Vlasov system
167: (\ref{extended euler},\ref{1introcontrainte})
168: turns out to be a natural
169: extension (or $\Gamma$ limit)
170: of the Euler equations from both geometrical and variational
171: reasons, as explained in section \ref{1section-euler}
172: \\
173: In a similar way, there is a monokinetic version of the $VP$ system,
174: the so-called (pressureless) Euler-Poisson ($EP$) system, which reads
175: \beq
176: &&\dt v +(v\cdot\nabla) v=\nabla \varphi\\
177: &&\dt \rho +\nabla\cdot(\rho v)=0\\
178: && \epsilon^2\Delta\varphi=\rho-1.
179: \enq
180: A rigorous asymptotic analysis of the $VMA$ system
181: as $\epsilon \rightarrow 0$ will
182: be provided (sections \ref{1asymptotic} and \ref{1section-Euler-Poisson}),
183: in the case when the initial electronic density
184: \beq
185: f(t=0,x,\xi)=f^0(x,\xi)\label{1introprinit}
186: \enq
187: is asymptotically monokinetic, namely approaching
188: $\delta(\xi-v_0(x)),$
189: for some smooth divergence free velocity field $v_0$, as
190: $\epsilon$ tends to zero.
191: Before this asymptotic analysis, we want to explain the
192: geometric origin
193: of the $VMA$ system.
194: It has been known,
195: since Arnold's celebrated work (cf. \cite{AK}),
196: that the Euler equations (formally) describe
197: geodesics curves along a suitable group of volume preserving maps,
198: lengths being measured in the $L^2$ sense.
199: We will show (section \ref{1section-euler})
200: that the $VMA$ system just describes
201: approximate geodesics obtained through
202: a very natural penalty method, where $\epsilon$ stands for
203: the penalty parameter.
204: For this geometric interpretation to be valid, the Monge-Amp\`ere
205: equation (\ref{1intromonge}) must be understood in the
206: following weak sense:
207: for each fixed $t$, $\varphi(t,\cdot)$ is the unique (up to an
208: additive constant) function such that
209: $\Psi(x)=x^2/2 + \epsilon^2\varphi(t,x)$ is convex in $x$ and
210: \beq
211: \forall g\in C^0(\mathbb R^d), \int_{ {\mathbb R}^d}
212: g(\nabla\Psi(x))\rho(t,x)dx\,=\,\int_{\Omega}g(y)dy,
213: \enq
214: where $\Omega$ is a fixed bounded open convex set where the neutralizing
215: background of the electrons is assumed to be located.
216: (This definition is made precise in section \ref{1section-projection}.)
217: Notice that, by construction, $\nabla\Psi$ must be valued in the closure
218: of $\Omega$ and, therefore, the potential $\varphi$ enjoys the following
219: property
220: $$
221: |x+\epsilon^2\nabla_x\varphi(t,x)|\le \sup_{y\in\Omega}|y|<+\infty.
222: $$
223: There is no similar bound for the electrostatic potential of the
224: classical $VP$ system. Thus, in some sense, the $VMA$ system can be seen
225: as a nonlinearly saturated version of the $VP$ system.
226: \\
227: Beyond the geometric derivation of the $VMA$ system,
228: our main analytic results are as follows:
229: \\
230: \begin{itemize}
231: \item The $VMA$ system admits global energy preserving weak solutions.
232: \item The $VMA$ system admits local strong solutions in periodic domains.
233: \item For well prepared, nearly monokinetic initial data, the solutions of the $VMA$ system
234: converge when $\epsilon$ goes to 0 to those of the Euler equations.
235: \item In this asymptotic, the $EP$ system is a higher
236: order approximation of the $VMA$ system.
237: \end{itemize}
238: \bigskip
239: The paper is organized as follows: in section 2,
240: we first recall the geometric nature of the Euler equations, then we explain
241: why the constrained Vlasov system
242: (\ref{extended euler},\ref{1introcontrainte}) is a natural
243: extension of the Euler equations from a variational point of view,
244: finally we introduce the concept of approximate geodesics
245: for volume preserving maps,
246: and derive the $VMA$ system.
247: Section 3 is devoted to the proof of existence of global
248: energy preserving weak solutions. In section 4,
249: we prove existence of local strong solutions, in the case of a periodic
250: domain. Finally, in section 5, we study the asymptotic behavior of the
251: $VMA$ system as $\epsilon$ goes to 0.
252:
253:
254: \section{The geometric origin of the Vlasov-Monge-Amp\`ere system}
255: \label{1section-euler}
256: \subsection{The Euler equations}
257: The motion of an incompressible fluid in a domain $\Omega\subset\Rd$
258: is classically described by the
259: Euler equations $(E)$:
260: \beq
261: &&\dt v +(v\cdot\nabla) v=-\nabla p\\
262: &&\nabla\cdot v =0,
263: \enq
264: with $t\in \R$, $x\in \Omega$, where $v=v(t,x)$
265: stands for the velocity field and $p=p(t,x)$ for the scalar pressure field.
266: These equations have a nice geometrical interpretation
267: going back to Arnold (see \cite{AK}).
268: Introducing $G(\Omega)$ the group of
269: all volume preserving diffeomorphisms of $\Omega$
270: with jacobian determinant equal to 1, and measuring lengths in the $L^2$ sense,
271: we may define
272: (at least formally) geodesic curves along $G(\Omega)$. It turns out
273: that the Euler equations just describe these curves.
274: For the same reasons, the Euler equations can be seen as the optimality equations
275: for the corresponding minimization problem: given two maps chosen in
276: $G(\Omega)$, find an $L^2-$shortest path between them along
277: $G(\Omega)$. It was shown by Shnirelman \cite{Shn1} (see also
278: \cite{AK} and \cite{Shn2}) that, in the case
279: when $\Omega$ is the unit cube in $\R^3$, there are many maps for which there
280: are no such shortest paths.
281: Beyond this negative result, \cite{Br5} established that
282: minimizing paths are more appropriately described by
283: doubly stochastic measures. These measures (also called polymorphisms)
284: generalize volume preserving maps in the following way:
285: a doubly stochastic measure $\mu(dx,dy)$ is a (Borel) probability measure on
286: $\Omega\times\Omega$ with two projections on each copy of $\Omega$ both equal
287: to the (normalized) Lebesgue measure. It is known -see \cite{Neretin},
288: for instance-
289: that any such $\mu$ can be weakly approximated by a sequence
290: $\mu_n(dx,dy)=\delta(x-g_n(y))dy$ where each $g_n$ is a volume preserving
291: map of $\Omega$.
292: In \cite{Br5} it was shown that, in the case considered by Shnirelman for which
293: there is no classical shortest path, minimizing paths along $G(\Omega)$
294: converge to paths of doubly stochastic measures $t\rightarrow \mu(t;dx,dy)$
295: governed by the following extension of the Euler equations
296: \beq
297: &&\dt \mu+\nabla_x\cdot(\mu v)=0,\label{varia1}\\
298: &&\dt (v\mu) +\nabla_x\cdot(\mu v\otimes v)+\mu\nabla_x p=0,\label{varia2}
299: \enq
300: where $v=v(t;x,y)$ and $p=p(t,x)$ can be respectively seen as the
301: velocity field and the pressure field attached to $\mu$.
302: (Notice that the velocity field $v$ generally depends on the extra variable $y$
303: and is not a classical but rather a multivalued
304: velocity field.)
305: These equations are just a reformulation of
306: the constrained Vlasov system (\ref{extended euler},\ref{1introcontrainte}).
307: Indeed, it can be checked, under appropriate regularity assumptions,
308: that the kinetic measure $f$ defined by
309: \beq\label{defmuf}
310: f(t;dx,d\xi)=\int_{y\in\Omega} \delta(\xi-v(t;x,y))\mu(t;dx,dy)
311: \enq
312: solves (\ref{extended euler},\ref{1introcontrainte})
313: when $(\mu,v,p)$ solves (\ref{varia1},\ref{varia2}).
314: Thus we conclude
315: that the constrained Vlasov system
316: (\ref{extended euler},\ref{1introcontrainte})
317: is a natural variational extension of the Euler equations.
318:
319:
320: \subsection{Approximate geodesics }\label{1section-apxgeo}
321: A general strategy to define approximate geodesics along a manifold $M$
322: (in our case $M=G(\Omega)$) embedded in a Hilbert space $H$
323: (here $H=L^2(\Omega,\Rd)$)
324: is to introduce a penalty parameter $\epsilon>0$
325: and the following $unconstrained$ dynamical system in $H$
326: \beq
327: \partial_{tt}{X} +\frac{1}{2\epsilon^2}\nabla_X\left(d^2(X,M))\right)
328: =0.
329: \label{1apxgeo}
330: \enq
331: In this equation,
332: the unknown $t\rightarrow X(t)$ is a curve in $H$,
333: $d(X,M)$ is the distance (in $H$) of $X$ to the manifold
334: $M$, i.e. in our case as $M=G(\Omega)$,
335: \beq
336: d(X,G(\Omega))=\inf_{g\in G(\Omega)}\|X-g\|_H,\label{1projection}
337: \enq
338: and, finally, $\nabla_X$ denotes the
339: gradient operator in $H$.
340: This penalty approach
341: has been used for the Euler equations
342: by the first author in \cite{Br3}.
343: It is similar-but not identical-
344: to Ebin's slightly compressible flow theory \cite{Eb}, and is
345: a natural extension of the theory of
346: constrained finite dimensional mechanical systems \cite{RU}.
347: The penalized system is formally hamiltonian in variables
348: $(X,\partial_t X)$ with
349: Hamiltonian (or energy) given by:
350: \be
351: E=\frac{1}{2}\|\partial_t X\|_H^2 + \frac{1}{2\epsilon^2}d^2(X,G(\Omega)).
352: \en
353: (Multiplying equation (\ref{1apxgeo}) by $\partial_t X$, we formally get
354: that the energy is conserved.)
355: Therefore it is plausible that the map $X(t)$ will remain close to $G(\Omega)$
356: if properly initialized at $t=0$.
357: A formal computation shows that, given a point $X$ for which there is
358: a unique closest point $\pi_X$ to $X$ in the $H$ closure
359: of $G(\Omega)$,
360: we have:
361: \beq
362: \nabla_X\left(d(X,G)\right)=\frac{1}{d(X,G)}(X-\pi_X).\label{1grad}
363: \enq
364: Thus the equation (\ref{1apxgeo}) formally becomes:
365: \beq
366: \label{1apxgeo2}
367: \partial_{tt} X + \frac{1}{\epsilon^2}(X-\pi_X)=0.
368: \enq
369: To understand why solutions to such a system may approach
370: geodesics along $G(\Omega)$ as
371: $\epsilon$ goes to 0, just recall that, in the simple framework of a
372: surface $S$ embedded in the 3 dimensional Euclidean space,
373: a geodesic $t\rightarrow s(t)$ along $S$
374: is characterized by the fact that
375: for every $t$, the plane defined by $\{\dot s(t), \ddot s(t)\}$ is
376: orthogonal to $S$.
377: In our case, $\partial_{tt} X(t)$ is nearly
378: orthogonal to $G(\Omega)$ thanks to
379: (\ref{1apxgeo2}), meanwhile $X(t)$ remains close to $G(\Omega)$.
380: \\
381: The approximate geodesic equation was introduced in \cite{Br3}
382: in order to allow a spatial approximation of $G(\Omega)$
383: by the group of permutations of $N$ points $A_j$ chosen to form a discrete
384: grid on $\Omega$. On such a discrete group, the concept of geodesics
385: becomes unclear meanwhile approximate geodesics still make sense.
386: They can be interpreted as trajectories of a cloud of $N$ particles $X_i$
387: moving in the Euclidean space ${\mathbb R}^{dN}$, which substitutes
388: for $H$. These particles solve the following coupled system of
389: harmonic oscillators
390: \[
391: \epsilon^2\frac{d^2 X_i}{dt^2}+ X_i-A_{\sigma_i}=0,
392: \]
393: where $\sigma$ is a time dependent permutation minimizing,
394: at each fixed time $t$,
395: $\Sigma \left|X_i -A_{\sigma(i)}\right|^2$ among all other permutations
396: of the first $N$ integers.
397: The convergence of this discrete model to the incompressible
398: Euler equations for well prepared initial data was proved in \cite{Br3}.
399: In order to study the continuous version (\ref{1apxgeo2}),
400: a specific study of the projection problem (\ref{1projection}) is needed.
401:
402:
403: \subsection{The polar decomposition Theorem}\label{1section-projection}
404:
405: Let us first recall a general measure theoretic definition:
406: \begin{defi}
407: Let $A$ and $B$ be two topological spaces, let
408: $\rho$ be a Borel finite
409: measure of $A$ and $X$ a Borel map $A\rightarrow B$,
410: we call the push-forward of $\rho$ by $X$ and note $X\#d\rho$ the Borel
411: measure $\eta$ on $B$ defined by
412: \be
413: \forall f\in C^0(B),\;\int_{B}f(y) d\eta(y) = \int_{A}f(X(x))d\rho(x).
414: \en
415: \end{defi}
416: Let us now consider the case of a bounded open subset $\Omega$ of
417: the Euclidean space $\Rd$ equipped with the Lebesgue
418: measure that we denote $dx$.
419: We say that a Borel map $s: \overline{\Omega}\rightarrow \overline{\Omega}$ is volume
420: (or Lebesgue measure) preserving
421: if $s\#dx=dx$, i.e. if for all $g\in C^0(\overline{\Omega})$ one has
422: $\int_{\Omega} g(x)dx =\int_{\Omega} g(s(x))dx,$
423: or equivalently, for any Borel subset $B$ of $\overline{\Omega}$ one has $|s^{-1}(B)|=|B|$.
424: The set of all measure preserving maps of $\Omega$ is
425: a closed subset of the Hilbert space $H=L^2(\Omega,\Rd)$
426: and will be denoted by $S(\Omega)$.
427: Notice that $S(\Omega)$ is only a semi-group for
428: the composition rule and contains the group
429: of volume preserving diffeomorphisms $G(\Omega)$.
430: It is known \cite{Ne} that, at least in the case when $\Omega$ is convex
431: and $d\geq 2$, $S(\Omega)$ is exactly the closure of $G(\Omega)$
432: in $L^2(\Omega,\Rd)$, which implies $d(.,G(\Omega))=d(.,S(\Omega)).$
433: \\
434: The polar decomposition Theorem for maps \cite{Br1} (extended to
435: Riemannian manifolds in \cite{Mc2})
436: will be crucial for
437: our analysis of the $VMA$ system:
438:
439: \begin{theo}\label{1brenier}
440: Let $\Omega$ be a bounded convex open subset of $\Rd$,
441: let $X\in L^2(\Omega;\Rd)$
442: and $\rho_X=X\#dx,$ where $dx$ is the Lebesgue measure on $\Omega.$
443: Assume $\rho_X$ to be a Lebesgue integrable function,
444: or, equivalently, $X$ to satisfy the non-degeneracy condition:
445: \beq
446: \forall E\subset \Rd \mbox{ Borel },\, |E|=0\Rightarrow |X^{-1}(E)|=0.
447: \label{1nondeg}
448: \enq
449: Then there exists a unique pair $(\nabla\Phi_X, \pi_X)$
450: where $\Phi_X$ is a convex function and
451: $\pi_X\in S(\Omega)$, such that
452: \beq
453: X=\nabla\Phi_X\circ \pi_X.\label{1polaire}
454: \enq
455: In this 'polar decomposition',
456: $\pi_X$ is also characterized as the unique
457: closest point to $X$ on $S(\Omega)$ in the $L^2$ sense
458: and
459: $\Phi_X$ is characterized to be (up to an additive constant)
460: the unique convex function
461: on $\Omega$ satisfying
462: \beq
463: \int_{\Rd}g(x)d\rho_X=\int_{\Omega}g(X(y))dy=\int_{\Omega}g(\nabla\Phi_X(y))dy\label{1trans1},
464: \enq
465: for any $ g \in C^0(\Rd)$ such that $\left|g(x)\right|\leq C(1+|x|^2)$.
466: \\
467: In addition, the Legendre-Fenchel transform $\Psi_X$ of $\Phi_X$ defined by
468: \beq \label{1defLegendre}
469: \Psi_X(x)=\sup_{y\in \Omega}\{x\cdot y -\Phi_X(y)\}
470: \enq
471: is Lipschitz continuous on $\Rd$, with Lipschitz constant bounded by
472: $\sup_{x\in\Omega}|x|$ and has the following properties~:
473: \\
474: $\nabla\Psi_X(x)\in\Omega$ holds true for $\rho_X$ a.e. $x$,
475: \beq
476: \int_{\Rd}g(\nabla\Psi_X)\rho_X(x)dx=\int_{\Omega}g(\nabla\Psi_X(X(x)))dx=\int_{\Omega}g(x)dx\label{1trans2}
477: \enq
478: for any $g \in C^0(\overline{\Omega})$, and
479: \beq
480: &&\nabla\Phi_X(\nabla\Psi_X(x))=x \;\;\;\rho_X(x)dx \;a.e,\label{1inverse1}\\
481: &&\nabla\Psi_X(\nabla\Phi_X(y))=y\;\;\;dy \;a.e,\label{1inverse2}\\
482: &&\pi_X(y)=\nabla\Psi_X(X (y))\;\;\;dy \;a.e.\label{1inverse3}
483: \enq
484:
485:
486: \end{theo}
487: We make here several remarks on Theorem \ref{1brenier}:
488:
489: \bigskip
490: \noindent
491: {\bf Link with the Monge-Amp\`ere equation}
492: We can interpret (\ref{1trans1})
493: as a weak version of the Monge-Amp\`ere equation:
494: \be
495: \rho_X(\nabla \Phi)\det D^2\Phi=1
496: \en and (\ref{1trans2}) can be seen as a weak version of another Monge-Amp\`ere equation:
497: \be
498: &&\det D^2\Psi=\rho_X\\
499: && \nabla\Psi \mbox{ maps supp}(\rho_X)\mbox{ in }\Omega.
500: \en
501:
502:
503:
504: \bigskip
505: \noindent
506: The pair $(\Phi_X, \Psi_X)$ depends in fact only of $\Omega$ and the measure $\rho_X=X\#dx,$
507: and if condition (\ref{1nondeg}) fails,
508: then existence and uniqueness of the projection $\pi_X$
509: may fail, but existence and uniqueness of $\nabla\Phi_X$ remain true.
510:
511: Theorem \ref{1brenier} and the subsequent remarks allow us to introduce the following
512: notation that will be used throughout the paper:
513:
514: \begin{defi}\label{1MaOmegarho}
515: Let $\Omega$ be a fixed bounded convex open set of $\Rd$, let $\rho$ be a positive measure
516: on $\Rd$ of total mass $|\Omega|$, absolutely continuous w.r.t the Lebesgue measure and such that $\int (1+|x|^2) d\rho(x) < +\infty$. We call
517: $\Phi[\Omega,\rho]$, or, in short, $\Phi[\rho]$,
518: the unique up to a constant convex function on $\Omega$ satisfying
519: \beq
520: \forall g\in C^0(\mathbb R^d)\cap L^1(d\rho{}), \int_{ {\mathbb R}^d}g(x)d\rho(x)\,=\,\int_{\Omega}g(\nabla\Phi[\Omega,\rho](y))dy.
521: \enq We call $\Psi[\Omega,\rho]$ its Legendre-Fenchel transform
522: satisfying
523: \beq
524: \forall g\in C^0(\mathbb R^d)\cap L^1(\Omega, dy), \int_{ {\mathbb R}^d}g(\nabla\Psi[\Omega,\rho](x))d\rho(x)\,=\,\int_{\Omega}g(y)dy.
525: \enq
526: \end{defi}
527: If no confusion is possible we may write $\Phi$ (resp. $\Psi$)
528: instead of $\Phi[\Omega,\rho]$ (resp. $\Psi[\Omega,\rho]).$
529:
530: \bigskip
531: \noindent
532: We will use some additional results from
533: \cite{Br1}. The first one establishes
534: the continuity of the polar decomposition:
535: \begin{theo}\label{1convergence}
536: Let $\rho$ be a Lebesgue integrable positive measure on $\Rd$, with total mass $\Omega$, such that $\int (1+|x|^2) d\rho < +\infty$.
537: Let $\rho_n$ be a sequence of Lebesgue integrable positive measures on $\Rd$, with total mass $\Omega$, such that
538: $\forall n$, \mbox{$\int (1+|x|^2) d\rho_n < +\infty$}.
539: Let $\Phi_n=\Phi[\Omega,\rho_n]$ and
540: $\Psi_n=\Psi[\Omega,\rho_n]$ be as in Definition \ref{1MaOmegarho}.
541: If for any $f\in C^0(\Rd)$ such that $|f(x)|\leq C(1+|x|^2),$ $\int f \ d\rho_n$ converges to $\int f \ d\rho $,
542: then
543: \begin{itemize}
544: \item $\Phi_n$ converges to $\Phi[\Omega,\rho]$
545: uniformly on each compact set of $\Omega$ and strongly in $W^{1,1}(\Omega),$
546: \item $\Psi_n$ converges to $\Psi[\Omega,\rho]$ uniformly on each compact set of $\Rd$ and strongly in $W^{1,1}(K)$ for every $K$ compact in $\Rd$.
547: \end{itemize}
548: \end{theo}
549: The second one provides a 'dual' definition of the distance between
550: a map $X$ and the semi-group $S(\Omega)$:
551: \begin{theo}\label{1distance}
552: Let $X\in L^2(\Omega;\Rd)$
553: and $\rho=X\#dx,$ where $dx$ is the Lebesgue measure on $\Omega.$
554: Assume $\rho$ to be a Lebesgue integrable function.
555: Then
556: $$
557: \frac{1}{2}d^2(X,S(\Omega))
558: =\int \left(|x|^2/2-\Psi[\Omega,\rho](x)\right)\rho(x)dx
559: +\int_{\Omega} \left(|y|^2/2-\Phi[\Omega,\rho](y)\right)dy\\
560: $$
561: $$
562: =\sup_{u,v}
563: \int \left(|x|^2/2-u(x)\right)\rho(x)dx +\int_{\Omega} \left(|y|^2/2-
564: v(y)\right)dy,
565: $$
566: where the supremum if performed over all pairs $(u,v)$ of continuous
567: functions on $\Rd$ such that $u(x)+v(y)\ge x\cdot y$ pointwise.
568: \end{theo}
569:
570: \subsection{The Vlasov-Monge-Amp\`ere system}
571:
572: Let us now derive the $VMA$ system as the kinetic formulation
573: of the approximate
574: geodesic equation (\ref{1apxgeo2}).
575: First, from the polar decomposition Theorem \ref{1brenier}, equation
576: (\ref{1apxgeo2}) reads
577: \beq
578: \partial_{tt} X(t,x) = \nabla\varphi(t,X(t,x)),\label{1odeapxgeo}
579: \enq
580: where
581: \beq
582: \label{1varphi}
583: \nabla\varphi(t,x)=\frac{\nabla\Psi[\Omega,\rho(t,\cdot)](x)-x}{\epsilon^2}
584: \enq
585: and $\Psi[\Omega,\rho]$ is as in Definition (\ref{1MaOmegarho}).
586: This means that $\nabla\varphi$ satisfies (\ref{1intromonge}) in
587: a weak form with the additional
588: condition that the range of $x\rightarrow x+\epsilon^2\nabla\varphi(t,x)$ is
589: contained in $\overline{\Omega}$.
590:
591: Next, let $f^0\ge 0$ be a given initial density function, that
592: we assume to be in $L^{\infty}(\Rd\times\Rd)$, compactly supported and
593: satisfying the compatibility condition
594: \beq
595: \int f^0(x,\xi)dxd\xi=|\Omega|.\label{1jauge}
596: \enq
597: For each $t\ge 0$, let us define $(x,\xi)\rightarrow f(t,x,\xi)$ to be
598: $f^0$ pushed forward
599: by the following ODE
600: \beq
601: &&\partial_t X(t,x,\xi)=\Xi(t,x,\xi)\\
602: &&\partial_{t} \Xi(t,x,\xi) = (\nabla\varphi)(X(t,x,\xi))\label{1od1}\\
603: && (X,\Xi)(t=0,x,\xi)=(x,\xi)\label{1od2}.
604: \enq
605: Then $f$ satisfies the following kinetic (or Liouville) equation
606: \beq
607: &&\frac{\partial f}{\partial t}+\nabla_x\cdot\left(\xi f\right)
608: +\nabla_{\xi}\cdot\left(\nabla\varphi f\right)=0\label{1principale}\\
609: &&f(0,\cdot,\cdot)=f^0,\label{1prinit}
610: \enq
611: which must be understood in the following weak sense
612: \beq
613: &&\forall g \in C^{\infty}_c([0,+\infty)\times \Rd \times \Rd),\nonumber\\
614: &&\int_0^\infty dt
615: \int_{\Rd\times\Rd} \left(\frac{\partial g} {\partial t} + \xi\cdot \nabla_x g
616: + \nabla\varphi\cdot\nabla_{\xi} g \right ) f dx d\xi\nonumber\\
617: && = -\int_{\Rd\times\Rd}
618: f_0(x,\xi)g(t=0,x,\xi)dx d\xi.
619: \enq
620: This linear Liouville equation is
621: nonlinearly coupled to equation (\ref{1varphi}),
622: where $\rho$ is linked to $f$ by equation (\ref{1rho}).
623: Finally, we have defined, through
624: (\ref{1varphi},\ref{1principale},\ref{1prinit}), the weak formulation
625: of the $VMA$ initial value problem.
626:
627: \noindent
628: The energy of the system is defined by
629: \beq
630: E(t)=&&\frac{1}{2}\int_{\Rd\times\Rd}f(t,x,\xi)|\xi|^2dxd\xi\nonumber\\
631: +&&\frac{1}{2\epsilon^2}\int_{\Rd}\rho(t,x)\left|\nabla\Psi[\Omega,\rho](t,x)-x\right|^2dx\label{1engdef}.
632: \enq
633:
634:
635:
636:
637:
638:
639:
640:
641:
642:
643: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
644: %%%%%%%%%%%%%%%%%%%%%% GLOBAL WEAK SOLUTION %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
645: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
646:
647:
648:
649: \section{Existence of global renormalized weak solutions}
650: The main result of this section is as follows:
651: \begin{theo}\label{1globalweak}
652: Let $(x,\xi)\rightarrow f^0(x,\xi)\ge 0$ be in
653: $L^{\infty}(\Rd\times\Rd)$, with compact support in both $x$
654: and $\xi$, satisfying condition (\ref{1jauge}).
655: \\
656: Then the $VMA$ system (\ref{1varphi},\ref{1principale},\ref{1prinit})
657: admits a global
658: weak solution $(f,\rho,\Psi)$, with $f\in L^{\infty}(\R^+\times \Rd\times\Rd)$
659: and $(\rho, \nabla\psi) \in \Linf([0,T]\times\Rd)$ for all $T>0$. In addition,
660: each such weak solution enjoys the following properties:
661: \begin{itemize}
662: \item $f$ is a continuous function of $t$, valued in
663: %$L^{\infty}(\Rd\times\Rd)$ weak star,
664: $L^p(\Rd\times\Rd)$, for every $1\leq p <\infty$
665: \item the density $\rho$ is a continuous function of $t$, valued in
666: %$L^{\infty}(\Rd)$ weak star,
667: $L^p(\Rd)$, for every $1\leq p <\infty$,
668: \item the support of $f(t,\cdot,\cdot)$ in $(x,\xi)$ is
669: compact, with a diameter growing no more than linearly in $t$.
670: \item the total energy defined by (\ref{1engdef}) is conserved,
671: \item the 'renormalization' property (in the sense of \cite{DL})
672: $$\frac{\partial g(f)}{\partial t}+\nabla_x\cdot\left(\xi g(f)\right)
673: +\nabla_{\xi}\cdot\left( \nabla\varphi g(f)\right)=0$$
674: holds true for all $g\in C^1(\mathbb R)$,
675: \item the trajectories of (\ref{1od1},\ref{1od2}) are uniquely defined
676: for almost every initial condition $(x,\xi)$,
677: \item $t\rightarrow f(t,\cdot,\cdot)$ is just $f^0$ pushed forward
678: along the trajectories of (\ref{1od1},\ref{1od2}).
679: \end{itemize}
680: \end{theo}
681: \noindent
682: {\bf Proof of Theorem \ref{1globalweak}:}
683:
684: \noindent
685: We build a sequence of approximate
686: solutions $(f_h,\Psi_h)_{h>0}$ by time discretization and let the
687: time step $h$ go to zero.
688: To handle the limiting process,
689: the non-linear terms will be treated with the help
690: of Theorem \ref{1convergence}.
691: More precisely if one can extract a subsequence
692: such that, for every $t$,
693: $f_h(t,\cdot,\cdot)$ converges weakly, then we can deduce from Theorem
694: \ref{1convergence} that the corresponding sequence $\nabla\Psi_h(t,\cdot)$
695: will converge strongly, and this will allow us pass to
696: the limit in the nonlinear term.
697: \subsection{Construction of a sequence of approximate solutions}
698: We consider $\eta\in C^{\infty}_c(\Rd)$ such that $\eta\geq 0$, $\int_{\Rd}\eta =1$ and
699: $\eta_h=\frac{1}{h^d}\eta(\frac{\cdot}{h})$. We then seek approximate solutions as
700: solutions of the approximate problem
701: \beq
702: && \frac{\partial f_h}{\partial t}+\xi\cdot\nabla_xf_h
703: +\frac{\nabla\Psi_h(x)-x}{\epsilon^2}\cdot\nabla_{\xi} f_h=0\\
704: &&f_h(0,x,\xi)=f_h^0(x,\xi)=f_0*_{x,\xi} \eta_h\otimes\eta_h\\
705: &&\Psi_h(t)=\eta_h * \Psi[\Omega,\rho(t=nh)] \mbox{ for } t\in [nh,(n+1)h[.
706: \enq
707: $\nabla\Psi_h$ being a smooth function of space this regularized equation admits a unique solution
708: that one builds by the method of characteristics.
709: Since the flow is
710: divergence-free in the phase space, the solution $f_h$ satisfies
711: \beq \label{1consnorme}
712: \forall p \in [1,+\infty],\;\|f_h(t)\|_{L^p(\Rd\times\Rd)}=\|f_h(0)\|_{L^p(\Rd\times\Rd)}.
713: \enq
714: By construction (through Theorem \ref{1brenier}),
715: $\nabla\Psi_h$ is valued in the convex bounded set $\overline{\Omega}$.
716: Suppose that $f^0(x,\xi)$ vanishes outside of the set
717: $\{x^2+\epsilon^2\xi^2\leq C^2\}$ for some constant $C>0$ fixed and
718: denote $R=\sup_{y\in \Omega}|y|$. Then we have
719: \begin{lemme}\label{1suppcpct}
720: $\forall t\geq 0,\; f_h(t,\cdot,\cdot)$ is supported in
721: $\{\sqrt{x^2+\epsilon^2\xi^2}\leq C+ Rt/\epsilon\}$.
722: \end{lemme}
723: {\bf Proof :}
724: We just write
725: $$\epsilon^2 \partial_{tt}{X} + X = \nabla\Psi_h(X)$$
726: in complex notation
727: $-i\epsilon\partial_t{Z}+Z=F$, where $Z=X+i\epsilon\partial_t{X}$ and $F=\nabla\Psi_h(X)$,
728: which is bounded by $R$. This leads to
729: $$Z(t)=Z(0)\exp(-it/\epsilon)+i\epsilon^{-1}
730: \int_0^t \exp(-i(t-s)/\epsilon)F(s)ds$$
731: ant the desired bound easily follows.
732: Notice here a sharp contrast with the classical $VP$ system, for which
733: the $\xi-$support of the solutions
734: cannot be controlled so easily (except in the one dimensional case).
735: \hfill $\Box$
736:
737: \bigskip
738: \noindent
739: {\bf Convergence of the sequence of approximate solutions}
740: \\
741: \noindent
742: Using (\ref{1consnorme}) and Lemma \ref{1suppcpct} there exists, for any $1<p<\infty$,
743: up to the extraction of a subsequence, $f \in L^p([0,T]\times \Rd \times \Rd)$
744: such that $f_h $ converges weakly to $f$ as $h\rightarrow 0$.
745:
746: It remains to show that the product $f_h \nabla\Psi_h$ converges to the good limit.
747: For this we need strong convergence of $\nabla\Psi_h.$ We already know that
748: $\nabla\Psi_h \in L^{\infty}([0,T]\times \Rd).$
749: We claim that for all $t>0$, $\nabla\Psi_h(t,\cdot)$
750: converges strongly to $\nabla\Psi(t,\cdot)$
751: in $L^q_{loc}(\Rd)$, $\forall q\in [1,+\infty[$. Indeed, such a strong convergence
752: of $\nabla\Psi_h$ follows from Theorem \ref{1convergence} provided
753: that we have for all $t>0$,
754: \beq
755: \int_{{\mathbb R}^d}g(x)\rho_h(t,x)dx\rightarrow
756: \int_{{\mathbb R}^d}g(x)\rho(t,x)dx,
757: \label{1MA}
758: \enq
759: for any $g\in C^0({\mathbb R}^d)$ such that $\int (1+|x|^2)g(x)dx<+\infty$.
760: Note first that from Lemma \ref{1suppcpct}, we can restrict ourselves here to test functions $g$ that are compactly supported.
761: Then we show that the sequence $\rho_h$ is relatively compact in
762: $C([0,T],L^p(\Rd)-w).$ This is done by the following lemma:
763: \begin{lemme} For all $T>0$, for all $p$ with $1\leq p< \infty$ the sequence
764: $f_h$ (resp. $\rho_h$) satisfies
765: \begin{itemize}
766: \item $f_h$ (resp. $\rho_h$) is a bounded sequence in $\Linf([0,T];L^p(\Rd\times\Rd))$
767: (resp. in $\Linf([0,T];L^p(\Rd))$,
768: \item $\dt f_h$ (resp. $\dt \rho_h$) is a bounded sequence in $\Linf([0,T];W^{-1,p}(\Rd\times\Rd)))$,
769: (resp. in $\Linf([0,T];W^{-1,p}(\Rd))$,
770: \end{itemize}
771: and one can extract from $f_h$ (resp. from $\rho_h$) a subsequence converging in $C([0,T],L^p(\Rd\times\Rd)-w)$ (resp. in $C([0,T],L^p(\Rd)-w)$).
772: \end{lemme}
773: {\bf Proof:} the first point uses equation (\ref{1consnorme}) and Lemma \ref{1suppcpct}.
774: The second point uses equation (\ref{1principale}) and the identity:
775: $$\dt \rho_h=-\nabla_x \cdot \int_{\Rd} \xi f_h d\xi,$$
776: with the fact that the $f_h$ are uniformly compactly supported in $x$ and $\xi$ (Lemma \ref{1suppcpct});
777: the last point is a classical
778: result of functional analysis (see \cite{Li} for example). $\hfill$ $\Box$
779:
780:
781: \bigskip
782: \noindent
783: This lemma and Lemma \ref{1suppcpct} yield (\ref{1MA}). Then using Theorem \ref{1convergence},
784: with $\rho$ the limit of a subsequence of $\rho_h$, we have convergence
785: of the sequence $\nabla\Psi_h$ to $\nabla\Psi[\Omega,\rho]$ in $C([0,T],L^p(\Rd))$.
786: We have extracted a subsequence $f_h$ such that
787: \begin{itemize}
788: \item $f_h$ converges in $C([0,T],L^p(\Rd\times\Rd)-w)$ for every $1\leq p <\infty$.
789: \item $\rho_h$ converges in $C([0,T],L^p(\Rd)-w)$ for every $1\leq p <\infty$.
790: \item $\nabla\Psi_h(t,\cdot)$ converges in $L^p(\Rd)$ for every $t$ and
791: for every $1\leq p <\infty$\label{1Psiconv}.
792: \end{itemize}
793: Thus the limit $(f,\nabla\Psi)$ satisfies equations (\ref{1principale}-\ref{1prinit})
794: and the first part of Theorem \ref{1globalweak} is proved.
795:
796: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
797: %%%%%%%%%%%%%%%%%%%%%%CONSERVATION OF ENERGY%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
798: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
799:
800:
801:
802: \subsection{Conservation of energy}\label{1energysection}
803: We now give a rigorous proof of the conservation of energy
804: following an argument going back to F. Otto (in an unpublished
805: work on the semi-geostrophic equations).
806: We recall the definition of the energy as
807: \be
808: E(t)=\frac{1}{2}\int_{\Rd\times\Rd}f(t,x,\xi)|\xi|^2dxd\xi+\frac{1}{2\epsilon^2}\int_{\mathbb R^d}\rho(t,x)|\nabla\Psi(t,x)-x|^2dx.
809: \en We call the first term the kinetic energy $E_c$ and the second term,
810: multiplied by $\epsilon^2$, the (normalized) potential energy $E_p.$
811: We have
812: \begin{prop}
813: Let $f$ be any solution of (\ref{1principale}) such that on every interval $[0,T]$, $f(t,\cdot,\cdot)$ is uniformly compactly supported in $|x|, |\xi|\leq R(T)$ for some function $R(T)$. Then the energy of the solution $f$ is conserved.
814: \end{prop}
815: {\bf Proof:}
816: From Theorem \ref{1distance},
817: we know that
818: $$
819: E_p(t)=\int \left(|x|^2/2-\Psi(t,x)\right)\rho(t,x)dx
820: +\int_{\Omega} \left(|y|^2/2-\Phi(t,y)\right)dy
821: $$
822: $$
823: =\sup_{u,v}
824: \int \left(|x|^2/2-u(x)\right)\rho(t,x)dx +\int_{\Omega} \left(|y|^2/2-
825: v(y)\right)dy,
826: $$
827: where the supremum if performed over all pairs $(u,v)$ of continuous
828: functions on $\Rd$ such that $u(x)+v(y)\ge x\cdot y$ pointwise.
829: Thus for each $t, t_0 \in \mathbb R+$, we have
830: \be
831: E_p(t)&&\geq \int \left(|x|^2/2-\Psi(t_0,x)\right)\rho(t,x)dx +\int_{\Omega} \left(|y|^2/2-\Phi(t_0,y)\right)dy,
832: \en
833: and this implies
834: \begin{eqnarray*}
835: E_p(t)-E_p(t_0)&&\geq \int_{\mathbb R^d} \left(|x|^2/2-\Psi(t_0,x)\right)(\rho(t,x)-\rho(t_0,x))dx\\
836: &&= \int_{t_0}^{t}\int_{\mathbb R^d}\partial_t\rho(s,x)\left(|x|^2/2-\Psi(t_0,x)\right)dxds\\
837: &&= \int_{t_0}^{t}\int_{\R^d\times \Rd}\xi f(s,x,\xi)\left(x-\nabla\Psi(t_0,x)\right)dxd\xi ds.
838: \end{eqnarray*}
839: Notice that the product in the second line is licit
840: since $\partial_t\rho$ is
841: in $W^{-1,p}$ for any $1\leq p < \infty$,
842: $f(t,\cdot,\cdot)$ and therefore $\rho(t,\cdot)$ are compactly supported in space uniformly on $[0,T]$,
843: and $\Psi-|x|^2/2$ is in $W^{1,\infty}_{loc}$.
844: Exchanging $t_0$ and $t$ we would have found
845: \be
846: E_p(t_0)-E_p(t)\geq\int_{t}^{t_0}\int_{\R^d\times\Rd}\xi f(s,x,\xi)\left(x-\nabla\Psi(t,x)\right)dxd\xi ds,
847: \en
848: moreover we have for the kinetic energy
849: \be
850: \epsilon^2(E_c(t)-E_c(t_0))=
851: \int_{t_0}^{t}\int_{\R^d\times\Rd} \xi f(t,x,\xi)\cdot(\nabla\Psi(s,x)-x)dxd\xi ds.
852: \en
853: Dividing by $t-t_0,t> t_0$ we find
854: \be
855: &&\epsilon^2\frac{E(t)-E(t_0)}{t-t_0}\\
856: &&\geq\frac{1}{t-t_0}\int_{t_0}^{t} \int_{\R^d\times\Rd}\xi f(s,x,\xi)\cdot(\nabla\Psi(s,x)-\nabla\Psi(t_0,x))dxd\xi ds
857: \en
858: and
859: \be
860: &&\epsilon^2\frac{E(t)-E(t_0)}{t-t_0}\\
861: &&\leq\frac{1}{t-t_0}\int_{t_0}^{t}\int_{\Rd\times\Rd} \xi f(s,x,\xi)\cdot(\nabla\Psi(t,x)-\nabla\Psi(s,x))dxd\xi ds.
862: \en
863: We know from \ref{1Psiconv} that $\nabla\Psi(t,.)$ converges strongly in $L^p_{loc}(\Rd), 1\leq p<\infty$ to $\nabla\Psi(t_0,.)$ as $t$ goes to $t_0$, and so the right hand sides of the above inequalities converges to 0
864: and we conclude that
865:
866: $$\lim_{t>t_0}\frac{E(t)-E(t_0)}{t-t_0}= 0.$$
867: We could take $t<t_0$ and find the same result. Finally we conclude that
868: \be
869: \frac{dE}{dt}\equiv 0.
870: \en
871: $\hfill$ $\Box$
872:
873:
874:
875: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
876: %%%%%%%%%%%%%%%%%%%%%% RENORMALIZATION %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
877: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
878:
879:
880: \subsection{Renormalized solutions and existence of characteristics}
881: The study of renormalized solutions for transport equations
882: has been introduced in \cite{DL} for
883: vector fields in $W^{1,1}$ with bounded divergence.
884: These results have been extended by Bouchut \cite{Bo}
885: to the case of Vlasov-type
886: equations with acceleration field in $BV$
887: (A recent result of L. Ambrosio, \cite{Am}, has extended the existence of renormalized solutions to transport equations with vector fields in $BV$ and with bounded divergence).
888: The fact that solutions of (\ref{1principale}, \ref{1prinit}) are renormalized solutions is an immediate consequence of the following theorem:
889:
890: \begin{theo}\label{1Bouchut} $(${\rm F. Bouchut}$)$
891: \\
892: Let $f \in L^{\infty}(]0,T[,L^{\infty}_{loc}( \Rd \times \Rd))$ satisfy
893: \begin{eqnarray*}
894: \frac{\partial f}{\partial t}+\nabla_x\cdot\left(\xi f\right)
895: +\nabla_{\xi}\cdot\left( E(t,x)f\right)=0,
896: \end{eqnarray*}
897: with $E(t,x)\in L^1(]0,T[;L^1_{loc}(\Rd))\cap L^{1}(]0,T[;BV_{loc}(\Rd))$,
898: \\
899: then, for
900: any $g\in C^1(\R)$,
901: $$
902: \frac{\partial g(f)}{\partial t}+\nabla_x\cdot\left(\xi g(f)\right)
903: +\nabla_{\xi}\cdot\left( E(t,x)g(f)\right)=0,
904: $$
905: and for every $1\leq p <\infty$, $f$ belongs to $C(]0,T[, L^p_{loc}(\Rd\times\Rd))$.
906:
907: \end{theo}
908:
909:
910: In our case the BV bound on the
911: acceleration $\nabla\Psi$
912: is a direct consequence of the fact that $\Psi$ is a globally
913: Lipschitz convex function.
914: This result implies the strong time continuity results for $f$ and $\rho$ in Theorem \ref{1globalweak}.
915: Finally, as in \cite{DL}, it can be deduced from
916: the renormalization property that
917: \\
918: 1) for almost every initial condition $(x,\xi)$,
919: there is a unique trajectory solving (\ref{1od1},\ref{1od2}),
920: \\
921: 2) $t\rightarrow f(t)$ is just $f^0$ pushed forward along these trajectories.
922:
923: A complete proof is given in appendix.
924:
925: Remark: From the renormalization property it follows that,
926: once the potential $\Psi(t,x)$ is known, there exists a unique solution to
927: (\ref{1principale})
928: in $\Linf_{t,x,\xi}$. Of course, this does not imply at all
929: the uniqueness of weak solutions to the
930: Vlasov-Monge-Amp\`ere system!
931: This paragraph ends the proof of Theorem \ref{1globalweak}.
932:
933:
934: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
935: %%%%%%%%%%%%%%%%%%%%%%%%%%%%% STRONG SOLUTIONS %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
936: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
937:
938:
939:
940: \section{Strong solutions}
941: In this section we show existence of strong solutions over a finite time intervall. To do so, we need regularity estimates for solutions of
942: Monge-Amp\`ere equation. We will get rid of the difficulties that may arise at the free boundary of the set $\{\rho>0\}$ by considering the periodic case.
943: Note that for the Vlasov-Poisson system, existence of global smooth solutions
944: has been proved (see \cite{Pf});
945: in the present case, due to the non-linearity of the Monge-Amp\`ere equation,
946: we were only able to obtain a result for finite time.
947:
948: \subsection{The periodic Vlasov-Monge-Amp\`ere system}
949: \subsubsection*{Polar factorization of maps in a periodic domain}
950: The polar decomposition Theorem has been generalized in \cite{Mc2} to general Riemannian manifolds,
951: while the particular case of the flat torus $\Td=\Rd/\mathbb Z^d$ had been addressed in \cite{Co}.
952: \begin{defi}
953: We say that a mapping $Y: \Rd \rightarrow \Rd$ is
954: $\mathbb Z^d$ additive if the mapping $x\rightarrow Y(x)-x$ is $\mathbb Z^d$ periodic.
955: The set of all measurable $\mathbb Z^d$ additive mappings is denoted by ${\cal P}$.
956: For each $x\in \Rd$ we call $\hat x$
957: the class of $x$ in $\Rd/\mathbb Z^d$, and for any $X\in {\cal P}$, $\hat X$ the mapping
958: of $\Td$ into itself defined by
959: $$\forall x \in \Rd, \hat X(\hat x)=\hat{X(x)}.$$
960: \end{defi}
961: We may say if no confusion is possible additive instead of $\mathbb Z^d$ additive.
962: Then the following theorem can be deduced from the results of \cite{Co} and \cite{Mc2}:
963:
964:
965: \begin{theo}\label{1McCann}
966: Let $X: \Rd \rightarrow \Rd$ be additive and
967: assume that $\rho_X= X\#dx$ has a density in $L^1([0,1]^d)$.
968: Then there exists a unique pair $(\nabla\Phi_X, \pi_X)$ such that
969: $$X=\nabla\Phi_X\circ \pi_X$$
970: where
971: $\Phi_X$ is a convex function and
972: $\Phi_X(x)-|x|^2/2$ is $\Zd$ periodic,
973: $\pi_X: \Rd \rightarrow \Rd$ is additive and
974: $\hat \pi_X$ is Lebesgue measure preserving in $\Td$.
975: Moreover we have
976: $$\|X- \pi_X\|_{L^2([0,1]^d)}=\|\hat X- \hat \pi_X\|_{L^2(\Td)}
977: $$
978: and, $\Psi_X$ denoting the Legendre transform of $\Phi_X$, we have
979: $$\pi_X=\nabla\Psi_X\circ X.$$
980: \end{theo}
981:
982: \noindent
983: Remark : The pair $(\Phi_X, \Psi_X)$ is uniquely defined by the density
984: $\rho_X=X\#dx$.
985:
986: \medskip
987: \noindent
988: Notice that the periodicity of $\Phi_X(x)-|x|^2/2$
989: implies that $\nabla\Phi_X$ and $\nabla\Psi_X$ are $\mathbb Z^d$ additive, and that
990: $\Psi_X -|x|^2/2$ is also $\mathbb Z^d$ periodic.
991: As in the previous case, we introduce the following notation:
992: \begin{defi}\label{1MaTdrho}
993: Let $\rho$ be a probability measure on $\Td$, with density in $L^1(\Td)$. We denote $\Phi[\rho]$
994: (resp. $\Psi[\rho]$) the unique up to a constant convex function such that
995: \beq
996: && \Phi[\rho]-|x|^2/2 \mbox{ is } \Zd \mbox{ periodic },\\
997: &&\forall f \in C^0(\Td),\;\int_{\Td} f(\hat{\nabla\Phi}[\rho](x))dx=\int fd\rho
998: \enq
999: (resp. its Legendre fenchel transform).
1000: \end{defi}
1001: $\Psi[\rho]$ will thus be a generalized
1002: solution of the Monge-Amp\`ere equation \\
1003: $\det D^2\Psi=\rho.$
1004: \\Next the results of Caffarelli (\cite{Ca1}, \cite{Ca2}, \cite{Ca3})
1005: on the regularity of solutions
1006: of the Monge-Amp\`ere equation yield the following theorem:
1007:
1008: \begin{theo}\label{1regper}
1009: Let $\rho>0$ be a $ C^{\alpha}(\Td)$
1010: probability density on $\Td$, for some $\alpha \in ]0,1[$.
1011: \\
1012: Then $\Psi=\Psi[\rho]$ (see Definition \ref{1MaTdrho}) is a classical solution of
1013: \begin{eqnarray*}
1014: \det D^2\Psi=\rho\label{1mongeper}
1015: \end{eqnarray*}
1016: and satisfies:
1017: \begin{eqnarray*}
1018: &&\|\nabla\Psi(x)-x\|_{\Linf}\leq C(d)=\sqrt d /2\\
1019: &&\|D^2\Psi\|_{C^{\alpha}}\leq K(m,M,\|\rho\|_{C^{\alpha}})
1020: \end{eqnarray*}
1021: where $m=\inf\rho$ and $M=\sup\rho$.
1022: \end{theo}
1023: This theorem is an adaptation of the regularity results stated above, whose complete proof
1024: is given in appendix.
1025: \subsubsection*{The periodic Vlasov-Monge-Amp\`ere system}
1026: We now seek $f: (t,x,\xi) \in (\Td\times \Rd \times [0,T]) \rightarrow f(t,x,\xi)\in \R^+$, for some $T>0$,
1027: solution of the initial value problem for the periodic Vlasov-Monge-Amp\`ere $(VMA_p)$ system
1028: \beq
1029: &&\frac{\partial f}{\partial t}+\nabla_x\cdot\left(\xi f\right)
1030: +\frac{1}{\epsilon^2}\nabla_{\xi}\cdot\left((\nabla\Psi[\rho](x) - x)f\right)=0\label{1principaleper}\\
1031: &&f(0,\cdot,\cdot)=f^0\label{1prinitper},
1032: \enq
1033: for a given $f^0$ satisfying the compatibility condition
1034: \beq
1035: \int f^0(x,\xi)dxd\xi= 1.\label{1jaugeper}
1036: \enq
1037: The macroscopic density $\rho$ is still related to $f$ by equation (\ref{1defrho}), and $\Psi[\rho]$ is as in Definition \ref{1MaTdrho}.
1038:
1039:
1040:
1041: \subsection{Existence of local strong solutions}
1042:
1043: We mention first that the proof of existence of global weak solutions adapts
1044: with minor changes to the periodic case, and that the obtain for the periodic $(VMA_p)$ system the same result as Theorem \ref{1globalweak}.
1045:
1046: Our result in this section is the following:
1047: \begin{theo}\label{1strong}
1048: Let $f_0 \in W^{1,\infty}(\Td\times\Rd),$ be such that:
1049: \beq
1050: &&\exists C_0 >0:\; f_0 \equiv 0\text{ for } |\xi|\geq C_0,\label{1pasvide1}\\
1051: &&\exists m >0:\;\rho_0 (x)=\int_{\Rd}f_0(x,\xi)d\xi\geq m\;\forall x\in \Td\label{1pasvide2},
1052: \enq
1053: then there exists $T>0$ and a solution $f$ to the $VMA_p$ system
1054: (\ref{1principaleper},\ref{1prinitper}), in the space
1055: $W^{1,\infty}([0,T]\times\Td \times \Rd)$.
1056: \end{theo}
1057:
1058:
1059: \bigskip
1060: \noindent
1061: {\bf Proof of Theorem \ref{1strong}:}
1062: First we deduce from Theorem \ref{1regper}:
1063: \begin{cor}\label{1regularite}
1064: Let $\rho,\Psi=\Psi[\rho]$ be as in Theorem \ref{1regper}. Then, we have
1065: $$\|D^2\Psi\|_{\Linf(\Td)}\leq C(m,M,\|\nabla\rho\|_{\Linf(\Td)}),$$
1066: and we can define
1067: $$K(m,M,l)=\sup\{\|D^2\Psi[\rho] \|_{L^{\infty}(\Td)};
1068: \;\;\|\nabla\rho\|_{\Linf(\Td)}\leq l,\;\;\; m\leq\rho\leq M\}<\infty.$$
1069: \end{cor}
1070: We see that in order to use Theorem \ref{1regper} we need $\rho$ to be bounded
1071: away from 0.
1072: In the following lemma, we show that under suitable assumptions on
1073: the initial data, it is possible to enforce locally in time
1074: the condition $0<m\leq\rho\leq M.$
1075:
1076: \begin{lemme}\label{1densite}
1077: Let $f\in \Linf([0,T]\times\Td\times\Rd)$ satisfy
1078: \begin{eqnarray*}
1079: &&\frac{\partial f}{\partial t}+\nabla_x\cdot\left(\xi f\right)
1080: +\nabla_{\xi}\cdot\left(E(t,x) f\right)=0\\
1081: &&f(0,.,.)=f^0,
1082: \end{eqnarray*}
1083: with $E\in L^1([0,T],BV(\Td))$ and
1084: \be
1085: \|E\|_{\Linf([0,T]\times\Td)}\leq F,
1086: \en
1087: let the initial condition $f_0$ be such that
1088: $$a(x,\xi)\leq f(0,x,\xi)\leq b(x,\xi),$$
1089: with $\rho_a(x)=\int a(x,\xi)d\xi\geq m>0$ and
1090: $\rho_b(x)=\int b(x,\xi)d\xi\leq M<\infty$ and $a,b$ satisfying
1091: \be
1092: |\nabla_{x,\xi}(a,b)|\leq \frac{c}{1+|\xi|^{d+2}}.
1093: \en
1094: Then there exists a constant $R>0$ depending on $m,M,c,F$, such that
1095: \be
1096: (\rho_a(x)-Rt)\leq\rho(t,x)\leq(\rho_b(x)+Rt).
1097: \en
1098: \end{lemme}
1099: The proof of the lemma is given in appendix.
1100:
1101:
1102:
1103: \subsubsection{Construction of approximate solutions}
1104:
1105: Let us consider $(t,x)\rightarrow E(t,x)$ a smooth vector-field on $\Td$, and write
1106: \be
1107: T_E(f)=\frac{\partial f}{\partial t}+\nabla_x\cdot\left(\xi f\right)
1108: +\nabla_{\xi}\cdot\left( E\,f\right).
1109: \en
1110: If $f$ satisfies $T_E(f)=0$, we have
1111: \be
1112: &&T_E\nabla_xf=-(\nabla_xE)\cdot\nabla_{\xi}f\label{1Tx}\\
1113: &&T_E\nabla_{\xi}f=-\nabla_xf\label{1Tv}\\
1114: &&T_E\partial_tf=-\partial_tE\cdot\nabla_{\xi}f,
1115: \en
1116: and therefore
1117: \beq\label{1Gronwall}
1118: \frac{d}{dt}\|\nabla_{x,\xi}f\|_{\Linf}\leq \|\nabla_{x,\xi}f\|_{\Linf}(1+\|\nabla_x E\|_{\Linf})
1119: \enq
1120: which implies
1121: \be
1122: \|\nabla_{x,\xi}f(t)\|_{\Linf}\leq \|\nabla_{x,\xi}f(t=0)\|_{\Linf}\exp \left(\int_{0}^{t}(1+\|\nabla_x E(s)\|_{\Linf})ds\right).
1123: \en
1124:
1125: \medskip
1126: \noindent
1127: Now let $f_0$ be given as in Theorem \ref{1strong}, satisfying (\ref{1pasvide1},\ref{1pasvide2}).
1128: Thanks to Lemma \ref{1densite} it is possible to find $t_1,m,M$ such that for any $f$ satisfying
1129: \be
1130: &&T_E(f)=0\\
1131: && f(t=0)=f_0,
1132: \en
1133: for any field $E\in L^1([0,t_1],BV(\Td))$ satisfying
1134: $\|E\|_{\Linf([0,t_1]\times \Td)}\leq \sqrt{d}/(2\epsilon^2)$,
1135: we have
1136: \beq
1137: &&m\leq\rho(t,\cdot)\leq M,\;\;\forall t\in [0,t_1]\label{1cond1}\\
1138: &&|\xi|_{max} \leq C_1=10 C_0\label{1cond2},
1139: \enq with $f$ supported in $\{|\xi|\leq |\xi|_{max}\}$ and with $C_0$ as in
1140: Theorem \ref{1strong}, so that we have for $0\leq t\leq t_1$:
1141: \beq\label{1nablarhof}
1142: \|\nabla\rho\|_{\Linf}\leq \omega_d C_1^d\|\nabla_x f\|_{\Linf},
1143: \enq
1144: $\omega_d$ being the volume of the unit ball of $\Rd$.
1145: Then we construct a family of approximate solutions $(f_h,\Psi_h)$
1146: to (\ref{1principaleper}),
1147: in the same spirit as we did for weak solutions,
1148: by solving
1149: \be
1150: && \frac{\partial f_h}{\partial t}+\xi\cdot\nabla_xf_h
1151: +\frac{\nabla\Psi_h(x)-x}{\epsilon^2}\cdot\nabla_{\xi} f_h=0\\
1152: &&f_h(t=0)=f_0\\
1153: &&\Psi_h(t)=\Psi[\rho(t=nh)] \mbox{ for } t\in [nh,(n+1)h[.
1154: \en
1155: Note that we have neither mollified the term $\nabla\Psi$ nor the initial condition and that $\|\nabla\Psi_h-x\|_{\Linf}\leq C(d)=\sqrt{d}/2$.
1156: Now choose $l=10\|\nabla_{x,\xi}f_0\|_{\Linf}\omega_d C_1^d$, if for some $t=nh\leq t_1-h$ we have
1157: \be
1158: \|\nabla_{x,\xi}f^h(t=nh)\|_{\Linf} \leq \frac{l}{\omega_d C_1^d}
1159: \en
1160: this implies, thanks to (\ref{1nablarhof}), that
1161: \be
1162: \|\nabla_x\rho^h(t=nh)\|_{\Linf}\leq l,
1163: \en
1164: and conditions (\ref{1cond1},\ref{1cond2}) are satisfied because $t\leq t_1$.
1165: Then if we denote $K=K(m,M,l)$ as in Corollary \ref{1regularite}, we have for $nh\leq t< nh+h,$
1166: \be
1167: \frac{d}{dt}\|\nabla_{x,\xi}f^h\|_{\Linf}\leq (K+1)\|\nabla_{x,\xi}f^h\|_{\Linf},
1168: \en
1169: and then
1170: \be
1171: \|\nabla_{x,\xi}f^h(t=nh+h)\|_{\Linf} \leq \|\nabla_{x,\xi}f^h(t=nh+h)\|_{\Linf}\exp{(K+1)h}.
1172: \en
1173: So if we define $T$ as
1174: \be
1175: T=\min \{t_1,t_2\},
1176: \en
1177: with $\exp((K+1)t_2)=10$,
1178: we have for $0\leq t \leq T$,
1179: \be
1180: &&\|\nabla_{x,\xi}f^h\|_{\Linf}\leq 10\|\nabla_{x,\xi}f_0\|_{\Linf}\\
1181: &&\|\nabla\rho^h\|_{\Linf}\leq l\\
1182: &&m\leq\rho\leq M\\
1183: &&\|D^2\Psi^h\|_{\Linf}\leq K.
1184: \en
1185:
1186: Thus we can extract a subsequence converging to a strong solution of (\ref{1principaleper},\ref{1prinitper}). Then we argue as in section 2 to show that all terms converge to the correct limit. This ends the proof of Theorem \ref{1strong}.
1187:
1188: $\hfill\Box$
1189:
1190:
1191:
1192:
1193:
1194:
1195:
1196:
1197: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1198: %%%%%%%%%%%%%%%%%%EULER INCOMPRESSIBLE%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1199: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1200:
1201: \section{Asymptotic analysis}
1202: \subsection{Convergence to the Euler equation}\label{1asymptotic}
1203: In this section we justify that the Vlasov-Monge-Amp\`ere system describes approximate geodesics on volume preserving transformations: indeed we will show that
1204: weak solutions of this system converge to a solution
1205: of the incompressible Euler equations $(E)$ as the parameter $\epsilon$
1206: goes to 0, at least for well prepared initial data.
1207: We restrict ourselves to the space periodic case, the macroscopic density $\rho$ is still defined by
1208: (\ref{1defrho}) and the convex potentials $\Phi[\rho],\Psi[\rho]$ are still as in Definition \ref{1MaTdrho}.
1209:
1210: %\be
1211: %\forall g \in C^0(\Td),\,&&\int_{\Td} g(x)d\rho(x)=\int_{\Td} g(\nabla\Phi[\rho](x))dx,\\
1212: % &&\int_{\Td} g(\nabla\Psi(x))d\rho(x)=\int_{\Td} g(x)dx.
1213: %\en
1214:
1215:
1216:
1217: For sake of simplicity, we slightly modify our notations and introduce the following rescaled potentials
1218: \be
1219: &&\tilde\varphi[\rho]=\frac{|x|^2/2-\Psi[\rho]}{\epsilon},\\
1220: &&\varphi[\rho]=\frac{\Phi[\rho]-|x|^2/2}{\epsilon},
1221: \en
1222: so that
1223: \be
1224: \nabla\varphi[\rho]=\nabla\tilde\varphi[\rho]\circ \nabla\Phi[\rho],
1225: \en
1226: and the $VMA_p$ system takes the following form:
1227: \begin{eqnarray}
1228: &&\frac{\partial f}{\partial t}+\xi\cdot\nabla_x f
1229: -\frac{\nabla\tilde\varphi[\rho]}{\epsilon}\cdot\nabla_{\xi}f=0
1230: \label{1principale2}\\
1231: &&f(0,\cdot,\cdot)=f_0\label{1initiale}.
1232: \end{eqnarray}
1233: The energy is given by
1234: \begin{eqnarray}\label{1energyvma}
1235: E(t)&&=\frac{1}{2}\int f(t,x,\xi)|\xi|^2 dxd\xi +
1236: \frac{1}{2}\int|\nabla\varphi|^2 dx\\
1237: &&=\frac{1}{2}\int f(t,x,\xi)|\xi|^2 dxd\xi +
1238: \frac{1}{2}\int\rho|\nabla\tilde\varphi|^2 dx\nonumber.
1239: \end{eqnarray}
1240: It has been shown in section \ref{1energysection} that the energy is conserved. The Euler equations for incompressible fluids $(E)$ reads:
1241: \beq
1242: &&\dt v + v\cdot\nabla v =-\nabla p\label{1euler1}\\
1243: && \nabla\cdot v = 0\label{1euler2}.
1244: \enq
1245: We shall here consider a smooth solution of $E$ and a weak solution of
1246: $VMA_p$, with `well prepared initial data',
1247: meaning that the initial data of both systems are close a time 0.
1248: Then we will show that as time evolves, both solutions stay close to each other.
1249: \begin{theo}
1250: \label{1neutre}
1251: Let $f$ be a weak solution of (\ref{1principale2}, \ref{1initiale}) with finite energy, let
1252: $(t,x)\rightarrow v(t,x)$ be a smooth $C^2([0,T]\times\Td)$ solution of (\ref{1euler1},\ref{1euler2}) for $t\in[0,T],$ and $p(t,x)$ the corresponding pressure, let
1253: \begin{eqnarray*}
1254: H_\epsilon(t)=\frac{1}{2}\int f(t,x,\xi)|\xi-v(t,x)|^2 dx d\xi + \frac{1}{2}\int|\nabla\varphi(t,x)|^2 dx,
1255: \end{eqnarray*}
1256: then
1257: \begin{eqnarray*}
1258: H_{\epsilon}(t)\leq C\exp(Ct)(H_{\epsilon}(0) + \epsilon^2),\;\forall t \in [0,T].
1259: \end{eqnarray*}
1260: $C$ depends only on
1261: $T, \sup_{0\leq s\leq T} \left\{ \|v(s,.), p(s,.), \partial_t p(s,.), \nabla p(s,.)\|_{W^{1,\infty}(\Td)}\right\}$.
1262:
1263: \end{theo}
1264:
1265: \medskip
1266: \noindent
1267: Remark 1: This estimate is enough to compare the weak solutions $f$ to
1268: the $VMA_p$ system (for well prepared initial data) and the smooth solutions $v$ of the Euler equations.
1269: For instance, $\int f(t=0,x,\xi)d\xi\equiv 1$
1270: implies \mbox{$\varphi(t=0,x)\equiv 0$} and therefore,
1271: $$\int|\xi-v(t=0,x)|^2 f(t=0,x,\xi) dxd\xi\le C_0\epsilon^2
1272: $$
1273: implies
1274: $$
1275: \sup_{t\in [0,T]}
1276: \int|\xi-v(t,x)|^2 f(t,x,\xi)dxd\xi\le C_T\epsilon^2,
1277: $$
1278: where $C_T$ depends only on $C_0$, $T$ and $v$.
1279: \\
1280: Remark 2: We see that we consider nearly monokinetic initial data for
1281: the $VMA_p$ system.
1282: \subsubsection*{Proof of Theorem \ref{1neutre}}
1283: We shall show that
1284: \begin{eqnarray}\label{1dtH}
1285: \frac{d}{dt}H_{\epsilon} =
1286: &&-\int f(t,x,\xi)(\xi-v)\nabla v(\xi-v)\nonumber\\
1287: &&+\int f(t,x,\xi)\frac{1}{\epsilon}v\cdot\nabla\tilde\varphi\nonumber\\
1288: &&-\int f(t,x,\xi)(v-\xi)\cdot\nabla p,
1289: \end{eqnarray}
1290: where we will use the notation
1291: \be
1292: u \ \nabla v \ w= \sum_{i,j=1}^d u^i \partial_{i}v^j w^j.
1293: \en
1294: The proof of this identity is postponed to the end of the section.
1295:
1296: Now we look at all terms of the right hand side. All the constants that we denote by $C$ are controlled as in Theorem \ref{1neutre}.
1297: We set
1298: \be
1299: &&T_1=-\int f(t,x,\xi)(\xi-v)\nabla v(\xi-v),\\
1300: &&T_2=\int f(t,x,\xi)\frac{1}{\epsilon}v\cdot\nabla\tilde\varphi,\\
1301: &&T_3=-\int f(t,x,\xi)(v-\xi)\cdot\nabla p.
1302: \en
1303: First we have $T_1\leq C H_{\epsilon}.$
1304: For $T_2$ we have
1305: \begin{eqnarray*}
1306: T_2=\frac{1}{\epsilon} \int \rho v \cdot\nabla\tilde\varphi =&&
1307: \frac{1}{\epsilon}\int v(\nabla\Phi[\rho])
1308: \cdot\nabla\tilde\varphi(\nabla\Phi[\rho])\\
1309: =&& \frac{1}{\epsilon}\int v(x+\epsilon\nabla\varphi)\cdot\nabla\varphi\\
1310: =&& \frac{1}{\epsilon}\int v\cdot\nabla\varphi + (v(x+\epsilon\nabla\varphi)
1311: -v(x))\cdot\nabla\varphi\\
1312: \leq &&0 + C \int\left|\nabla\varphi\right|^2 \leq C H_{\epsilon},
1313: \end{eqnarray*}
1314: we have used that $v$ is divergence-free thus its integral against any gradient is zero.
1315: Next we have the following lemma:
1316: \begin{lemme}
1317: \label{1rofi}Let $G:\Td\rightarrow \R$ be Lipschitz continuous such that $\displaystyle \int_{\Td} G =0$, then for all $R>0$, one has
1318: $$|\int \rho G| \leq \frac{1}{2}\|\nabla G\|_{\Linf}(\frac{1}{R}\epsilon^2 + R H_{\epsilon}).$$
1319: \end{lemme}
1320: {\bf Proof:} We just write a Taylor expansion of $G$:
1321: \begin{eqnarray*}
1322: &&\left|\int (\rho-1) G\right|=\left|\int(G(x+\epsilon\nabla\varphi)-G(x)\right|\\
1323: \leq &&\epsilon\|\nabla G\|_{\Linf} \|\nabla\varphi\|_{L^1}\leq \epsilon\|\nabla G\|_{\Linf} H_{\epsilon}^{1/2}\leq \frac{1}{2}\|\nabla G\|_{\Linf} (\frac{1}{R}\epsilon^2 + R H_{\epsilon}).
1324: \end{eqnarray*}
1325: $\hfill$ $\Box$
1326:
1327: \bigskip
1328: \noindent
1329: Again, since $v$ is divergence-free, $\int v\cdot\nabla p=0,$ thus from Lemma \ref{1rofi} we have
1330: \be
1331: -\int\rho v\cdot\nabla p\leq C (\epsilon^2 + H_{\epsilon})\nonumber.
1332: \en
1333: We remind that $$\dt \rho(t,x) = -\nabla_x\cdot\int f(t,x,\xi)\xi d\xi.$$
1334: Since it costs no generality to suppose that for all $t\in [0,T]$,
1335: $\int p(t,x)dx \equiv 0$, we obtain that
1336: \be
1337: \int f(t,x,\xi)\xi\cdot\nabla p
1338: &=&\int \frac{\partial\rho}{\partial t} p\\
1339: &=&\frac{d}{dt}\int\rho p - \int \rho\frac{\partial p}{\partial t}\nonumber\\
1340: &\leq & C(\epsilon^2 + H_{\epsilon})-\frac{dQ}{dt} \nonumber\\
1341: \en
1342: again using Lemma \ref{1rofi}, where $\;\displaystyle Q(t)=-\int \rho p.$ Thus
1343: \be
1344: T_3\leq C(H_{\epsilon}+\epsilon^2) - \frac{dQ}{dt}
1345: \en
1346: and we have the following inequality:
1347: \beq
1348: \frac{d}{dt}(H_{\epsilon}+Q)\leq CH_{\epsilon}+ O(\epsilon^2).\label{1qwe}
1349: \enq
1350: Moreover, using Lemma \ref{1rofi},
1351: \beq
1352: |Q(t)|\leq C\epsilon^2 + H_{\epsilon}(t)/2,\label{1Q}
1353: \enq
1354: thus
1355: \beq
1356: H_{\epsilon}+Q\geq H_{\epsilon}/2- C\epsilon^2,
1357: \enq
1358: and we can transform (\ref{1qwe}) in
1359: \begin{eqnarray}
1360: \frac{d}{dt}(H_{\epsilon}+Q)\leq C(H_{\epsilon}+Q)+ C\epsilon^2.
1361: \end{eqnarray}
1362: Gronwall's lemma then yields
1363: $$H_{\epsilon}(t)+Q(t)\leq (H_{\epsilon}(0)+Q(0)+Ct\epsilon^2)\exp(Ct).$$
1364: Using again (\ref{1Q}) we obtain
1365: \begin{eqnarray}
1366: H_{\epsilon}(t)\leq C(H_{\epsilon}(0)+\epsilon^2)\exp(Ct),
1367: \end{eqnarray}
1368: which achieves the proof of Theorem \ref{1neutre}.
1369:
1370: $\hfill$ $\Box$
1371:
1372: \bigskip
1373: \noindent
1374: {\bf Proof of identity (\ref{1dtH}):}
1375:
1376: \noindent
1377: We first notice that,
1378: for all $g\in C^1(\mathbb R\times\Td)$ , we have:
1379: $$\frac{d}{dt}\int\rho(t,x) g(t,x) dx=\int\int f(t,x,\xi)(\partial_t g(t,x) + \xi\cdot\nabla g(t,x)) d\xi dx.$$ We also use the conservation of energy defined by (\ref{1energyvma}). Then we get
1380: \begin{eqnarray*}
1381: \frac{d}{dt}H_{\epsilon}=&&\frac{d}{dt}\frac{1}{2}\int f(t,x,\xi)(|v|^2- 2\xi\cdot v)dx d\xi\\
1382: =&&\int f(t,x,\xi)(\dt v\cdot v - \dt v \cdot\xi)
1383: - \frac{1}{2}\int \nabla_x\cdot(f(t,x,\xi)\xi)(|v|^2- 2\xi\cdot v) \\
1384: && + \frac{1}{2}\int \nabla_\xi\cdot(\frac{1}{\epsilon}\nabla\tilde\varphi f(t,x,\xi))(|v|^2- 2\xi\cdot v).
1385: \end{eqnarray*}
1386: Integrating by part, we get
1387: \begin{eqnarray*}
1388: \frac{d}{dt}H_{\epsilon}=&&\int f(t,x,\xi)(\dt v\cdot v - \dt v \cdot\xi)+ \int f(t,x,\xi)\xi \nabla v (v-\xi)\\
1389: &&+\int f(t,x,\xi)\frac{1}{\epsilon}\nabla\tilde\varphi\cdot v.
1390: \end{eqnarray*}
1391: The first two terms can be rewritten as
1392: \begin{eqnarray*}
1393: &&\int f(t,x,\xi)(\dt v\cdot v - \dt v \cdot\xi)+
1394: \int f(t,x,\xi)\xi \nabla v (v-\xi)\\
1395: =&&-\int f(t,x,\xi)(v-\xi) \nabla v (v-\xi)
1396: +\int f(t,x,\xi)\dt v\cdot(v - \xi)\\
1397: && + \int f(t,x,\xi)v \nabla v (v-\xi)\\
1398: =&&-\int f(t,x,\xi)(v-\xi) \nabla v (v-\xi)
1399: + \int f(t,x,\xi)(v-\xi)\cdot(\dt v + v\cdot\nabla v),
1400: \end{eqnarray*} and finally using equation (\ref{1euler1}) we conclude.
1401:
1402: $\hfill$ $\Box$
1403: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1404: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1405: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1406: %%%%%%%%%%%%%%%%%%%% CONVERGENCE VERS EP %%%%%%%%%%%%%%%%%%%%%%%%%%
1407: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1408: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1409:
1410: \subsection{Comparison with the Euler-Poisson system}\label{1section-Euler-Poisson}
1411: Here we show that, as mentioned in the introduction,
1412: the Euler-Poisson ($EP$) system
1413: is a more accurate
1414: approximation to the Vlasov Monge-Amp\`ere
1415: system than the Euler equations, as $\epsilon$ goes to zero.
1416:
1417:
1418:
1419: \bigskip
1420: \noindent
1421: {\bf The $EP$ system}
1422: Let us recall
1423: that the (pressureless) Euler-Poisson system
1424: describes the motion of a continuum of electrons
1425: on a neutralizing background of ions through electrostatic interaction.
1426: Let $\bar v$ and $\bar\rho$ be the velocity and density of electrons.
1427: Let $\bar\varphi$ be the (rescaled) electric potential.
1428: Under proper scaling, these functions of $x\in \Rd$ and $t>0$ satisfy the Euler-Poisson
1429: system:
1430: \beq
1431: &&\dt \bar v + \bar v\cdot\nabla \bar v =-\frac{1}{\epsilon}\nabla\bar\varphi
1432: \label{1eulerpoisson1}\\
1433: &&\dt \bar\rho + \nabla\cdot(\bar\rho \bar v)=0\label{1eulerpoisson2}\\
1434: && 1-\epsilon\Delta\bar\varphi=\bar\rho\label{1eulerpoisson3}.
1435: \enq
1436: The so-called 'quasi-neutral' limit $\epsilon\rightarrow 0$ of similar systems
1437: has been studied for example in \cite{Gr} and \cite{CG}, and convergence
1438: results have been established using pseudo-differentials energy estimates.
1439: For well-prepared initial data, solutions of $EP$ are expected to behave
1440: as solutions of Euler incompressible equations.
1441: This fact is proved by the second author in his PhD thesis (\cite{These}, Chap 2), see also \cite{L1}.
1442: We give here the complete result that we will use herafter.
1443: We will denote by $\bar v\ep$ (resp. $f\ep)$ the solutions
1444: of the $EP$ (resp. $VMA_p$) system with parameter $\epsilon$.
1445: \begin{theo}\label{1cvep}
1446: Let $v$ be a solution of (\ref{1euler1},\ref{1euler2})
1447: on $[0,T]\times \Td$,
1448: with initial data $v_0$, and satisfying $v\in \Linf([0,T], H^s(\Td))$ for some $s\geq s_0(d)$. There for some $s'>0$, $s'< s$, if
1449: $(\bar v_0\ep, \bar\rho_0\ep)$ is such that the sequences
1450: \be
1451: \frac{\bar v_0\ep-v_0}{\epsilon}, \hspace{1cm} \frac{\bar\rho_0\ep-1}{\epsilon^2}
1452: \en
1453: are bounded in $H^{s'}(\Td)$, then there exists $T\epu>0$ with $\liminf_{\epsilon\rightarrow 0}T_{\epsilon}\geq T$ and a sequence $(\bar v\ep,\bar\rho\ep)$ of solutions to the $EP$ system on $[0,T_{\epsilon}[$
1454: with initial data $(\bar v_0\ep, \bar\rho_0\ep)$, belonging to $\Linf([0,T\epu], H^{s'}(\Td))$. Moreover, for $\epsilon$
1455: small enough, the sequences
1456: \be
1457: \frac{\bar v\ep-v}{\epsilon}, \hspace{1cm} \frac{\bar\rho\ep-1}{\epsilon^2}
1458: \en
1459: are bounded in $\Linf([0,T], H^{s'}(\Td))$. Finally, $s'$ goes to $+\infty$ as $s$ goes to $+\infty$.
1460: \end{theo}
1461:
1462:
1463:
1464: \subsubsection*{Assumptions}
1465: Here we consider $v$ a solution to $E$ (\ref{1euler1}, \ref{1euler2})
1466: with initial data $v_0$, a sequence $f\ep$ of solutions of $VMA_p$
1467: (\ref{1principale2},\ref{1initiale}) with initial data $f\ep_0$, and a sequence $(\bar v\ep, \bar\rho\ep)$ solutions of
1468: $EP$ (\ref{1eulerpoisson1}, \ref{1eulerpoisson2}, \ref{1eulerpoisson3}) with initial data $(\bar v \ep_0, \bar\rho\ep_0)$.
1469: We still define $H\epu$ as in Theorem \ref{1neutre}:
1470: \be
1471: H\epu(t)=\frac{1}{2}\int f\ep(t,x,\xi)|\xi- v(t,x)|^2 dx d\xi
1472: + \frac{1}{2}\int|\nabla\varphi\ep|^2 dx.
1473: \en
1474: We introduce the following assumptions:
1475:
1476: \begin{enumerate}
1477: \item[{\bf H0}] $v$ solution of $E$ satisfies, for some $C_0>0$, $\ds\|v\|_{\Linf([0,T] H^s(\Td))} \leq C_0$, and $s$ is large enough so that $s'$ in Theorem \ref{1cvep} satisfies $\ds s \geq s'>[\frac{d}{2}] + 2$.
1478:
1479:
1480: \item[{\bf H1}] The sequence $(\bar v_0\ep , \bar\rho_0\ep)$ of initial data of $EP$
1481: is such that, for some $C_1>0$,
1482: \be
1483: \sup_{\epsilon > 0}\left\{\frac{1}{\epsilon} \|\bar v\ep_0-v\|_{H^{s'}(\Td)}, \
1484: \frac{1}{\epsilon^2}\|\bar\rho\ep-1\|_{H^{s'}(\Td)}\right\} \leq C_1.
1485: \en
1486:
1487:
1488: \item[{\bf H2}] The sequence $f\ep_0$ satisfies $H\epu(0)\leq C_2\epsilon^2$ for some $C_2>0$.
1489:
1490: \end{enumerate}
1491:
1492: \bigskip
1493: \noindent
1494: {\bf H0}, {\bf H1}, {\bf H2} imply that
1495: \begin{enumerate}
1496: \item There exists $\tilde C_0$ such that
1497: \beq\label{1boundei}
1498: \|v\|_{\Linf([0,T],W^{2,\infty}(\Td))}\leq \tilde C_0.
1499: \enq
1500: \item From Theorem \ref{1neutre}, there exists $\tilde C_1$ such that
1501: \beq\label{1boundvma}
1502: H\epu(t)\leq \tilde C_1\epsilon^2 \mbox{ for } t\in[0,T].
1503: \enq
1504: \item From Theorem \ref{1vmavp} and Sobolev imbeddings, there exists $\tilde C_2$ such that
1505: \beq\label{1boundep}
1506: \sup_{\epsilon < \epsilon_0} \left\{ \left\|\frac{\bar v\ep-v}{\epsilon}, \hspace{.5cm} \frac{\bar\rho\ep-1}{\epsilon^2} \right\|_{\Linf([0,T],W^{2,\infty}(\Td))}\right\} \leq \tilde C_2.
1507: \enq
1508: \end{enumerate}
1509:
1510:
1511: \bigskip
1512: \noindent
1513: We are now ready to prove the following result:
1514:
1515:
1516: \begin{theo}
1517: \label{1vmavp}
1518: Let $f\ep_0,\bar v\ep_0,\bar\rho\ep_0, v, T$ be as above, satisfying assumptions {\bf H0}, {\bf H1}, {\bf H2}.
1519: Define
1520: \begin{eqnarray*}
1521: G_\epsilon(t)=\frac{1}{2}\int f\ep(t,x,\xi)|\xi-\bar v\ep(x)|^2 dx d\xi
1522: + \frac{1}{2}\int|\nabla\varphi\ep-\nabla\bar\varphi\ep|^2 dx.
1523: \end{eqnarray*}
1524: Then there exists $C>0$ such that
1525: \begin{eqnarray*}
1526: G_{\epsilon}(t)\leq C\exp(Ct)(G_{\epsilon}(0) + \epsilon^3),\;\forall t \in [0,T]
1527: \end{eqnarray*}
1528: where C depends on $s', C_0, C_1, C_2, T$.
1529:
1530:
1531:
1532:
1533: \end{theo}
1534: \bigskip
1535: \noindent
1536: Remark: the theorem shows that the distance between solutions of the $(EP)$ system
1537: and the $VMA_p$ system measured with $G\epu$ is like $O(\epsilon^3)$
1538: whereas Theorem \ref{1neutre} showed
1539: that the distance between the solution of
1540: the Euler equation and the $VMA_p$ system was like $O(\epsilon^2)$.
1541: Note also that $G\epu$ and $H\epu$ can both be interpreted as the square of a distance.
1542:
1543:
1544: \bigskip
1545: \noindent
1546: {\bf Proof of Theorem \ref{1vmavp}:}
1547: For notational simplicity, we drop most $\epsilon$'s.
1548: Proceeding as in (\ref{1dtH}) and noticing that:
1549: \be
1550: \Dt\int_{\Td}|\nabla\bar\varphi|^2=\frac{1}{\epsilon}\int_{\Td}\bar \rho \bar v
1551: \cdot\nabla\bar\varphi
1552: \en
1553: we obtain the following identity:
1554: \begin{eqnarray}\label{1dtH2}
1555: \frac{d}{dt}G_{\epsilon} =
1556: &&-\int f(t,x,\xi)(\xi-\bar v)\nabla \bar v(\xi-\bar v)\nonumber\\
1557: &&+\int f(t,x,\xi)\frac{1}{\epsilon}\bar v\cdot\nabla\tilde\varphi
1558: -\int f(t,x,\xi)\frac{1}{\epsilon}\bar v \cdot\nabla \bar\varphi \nonumber\\
1559: &&+\int f(t,x,\xi)\frac{1}{\epsilon}\xi\cdot\nabla \bar\varphi
1560: +\int \frac{1}{\epsilon}\bar\rho\bar v\cdot\nabla\bar\varphi\nonumber\\
1561: &&- \Dt \int \nabla\bar\varphi\cdot\nabla\varphi.
1562: \end{eqnarray}
1563: Then we notice
1564: \be
1565: \int f(t,x,\xi)\frac{1}{\epsilon}\xi\cdot\nabla \bar\varphi
1566: =\Dt\left(\int \frac{1}{\epsilon} \rho\bar\varphi\right)-\frac{1}{\epsilon}\int\rho\dt\bar\varphi.
1567: \en
1568: Next, we have the following lemma:
1569: \begin{lemme}\label{1Taylor}
1570: Define for any $\theta\in C^2(\Td)$
1571: \be
1572: &&<\nabla\theta>(x)=\int_{0}^{1}\nabla\theta(x+s\epsilon\nabla\varphi(x))ds,\\
1573: &&<\nabla^2 \theta>(x)=\int_{0}^{1}(1-s)\nabla^2 \theta(x+s\epsilon\nabla\varphi(x))ds.
1574: \label{1def<>}
1575: \en
1576: Then
1577: \be
1578: \int \rho\theta &=&\int\theta+\epsilon\int <\nabla\theta>\cdot\nabla\varphi \\
1579: &=& \int\theta+\epsilon\int \nabla\theta\cdot\nabla\varphi + \epsilon^2\int <\nabla^2 \theta>\nabla\varphi\nabla\varphi.
1580: \en
1581: \end{lemme}
1582: {\bf Proof:} The proof just uses the Taylor expansion and the identity \\
1583: $\int \rho\theta=\int\theta(x+\epsilon \nabla\varphi)$.
1584:
1585: $\hfill \Box$
1586:
1587: \bigskip
1588: \noindent
1589: Using Lemma \ref{1Taylor}, we get
1590: \be
1591: &&\frac{1}{\epsilon}\int \rho \dt\bar\varphi \\
1592: =&&\frac{1}{\epsilon}\int\dt\bar\varphi+\int\dt\nabla\bar\varphi\cdot\nabla\varphi
1593: +\epsilon\int<\dt \nabla^2\bar\varphi>\nabla\varphi\nabla\varphi.
1594: \en
1595: We claim that, under our assumptions, we have
1596: $$\|\dt \nabla^2\bar\varphi\|_{\Linf([0,T']\times\Td)}\leq C.$$
1597:
1598: \noindent
1599: Proof: from (\ref{1eulerpoisson2}), we have
1600: \be
1601: \dt\bar\rho=-\bar\rho\nabla\cdot \bar v - \bar v\cdot \nabla\bar\rho.
1602: \en
1603: Using (\ref{1boundep}), we obtain that $\ds\|\dt\bar\rho\|_{H^{s'-1}}\leq C \epsilon$. Since $H^{s'}(\Td)$ is continuously embedded in
1604: $W^{2,\infty}(\Td)$, $H^{s'-1}(\Td)$ is continuously embedded in $\Linf(\Td)$.\\
1605: Then, using (\ref{1eulerpoisson3})
1606: and classical elliptic regularity, we have
1607: $$\epsilon\|\dt \nabla^2\bar\varphi\|_{H^{s'-1}}\leq C \|\dt\bar\rho\|_{H^{s'-1}},$$
1608: and the desired result follows.
1609:
1610: $\hfill\Box$
1611:
1612: \bigskip
1613: \noindent
1614: This implies, using (\ref{1boundvma}), that
1615: \be
1616: \left | \epsilon\int<\dt \nabla^2\bar\varphi>\nabla\varphi\nabla\varphi \right |\leq C\epsilon^3.
1617: \en
1618: Next,
1619: \be
1620: &&\int\dt\nabla\bar\varphi\cdot\nabla\varphi=-\int \dt\Delta\bar\varphi\varphi\\
1621: &&=\frac{1}{\epsilon}\int \dt\bar\rho\varphi=\frac{1}{\epsilon}\int\bar\rho\bar v \cdot\nabla\varphi.
1622: \en
1623: Using again Lemma \ref{1Taylor}, we get
1624: \be
1625: &&\Dt \int \nabla\bar\varphi\cdot\nabla\varphi\\
1626: =&&\frac{1}{\epsilon}\Dt\left(\int \rho \bar\varphi-
1627: \epsilon^2\int<\nabla^2\bar\varphi>\nabla\varphi\nabla\varphi\right)
1628: \en
1629: and for the same reasons we have $\|\nabla^2\bar\varphi\|_{\Linf([0,T]\times\Td))}\leq C\epsilon$. This yields
1630: $$Q(t)=\epsilon\int<\nabla^2\bar\varphi>\nabla\varphi\nabla\varphi=O(\epsilon^4).$$
1631: Moreover, it does not cost to set $\int \bar\varphi\equiv 0$ and deduce
1632: \be
1633: \int f(t,x,\xi)\frac{1}{\epsilon}\xi\cdot\nabla \bar\varphi- \Dt \int \nabla\bar\varphi\cdot\nabla\varphi
1634: =-\frac{1}{\epsilon}\int\bar\rho\bar v \cdot\nabla\varphi+O(\epsilon^3)+\Dt Q.
1635: \en
1636: Thus the remaining terms are
1637: \be
1638: R=\frac{1}{\epsilon}\int \left[\rho\nabla\tilde\varphi -\rho\nabla\bar\varphi
1639: +\bar\rho\nabla\bar\varphi-\bar\rho\nabla\varphi\right]\cdot\bar v.
1640: \en
1641: Calculations that we postpone to the end of the proof show that
1642: \beq\label{1dtH3}
1643: R\leq&&\int (\nabla\varphi-\nabla\bar\varphi)\nabla\bar v (\nabla\varphi-\nabla\bar\varphi)\nonumber
1644: + C\int |\nabla\varphi-\nabla\bar\varphi|^2
1645: \\
1646: &&-\demi\int \nabla\cdot\bar v(|\nabla\bar\varphi|^2-2\nabla\varphi\cdot\nabla\bar\varphi ) + C\epsilon^3.
1647: \enq
1648: with $C$ depending on $\ds\|\nabla^2\bar v \|_{\Linf([0,T]\times\Td)}$ and $\ds\epsilon^{-1}\|\nabla^3\bar\varphi\|_{\Linf([0,T]\times\Td)}$, therefore uniformly bounded thanks to (\ref{1boundep}).
1649: Finally we obtain
1650: \be
1651: \Dt G\epu\leq&&-\int f(t,x,\xi)(\xi-\bar v)\nabla \bar v(\xi-\bar v)+(\nabla\varphi-\nabla\bar\varphi)\nabla\bar v (\nabla\varphi-\nabla\bar\varphi)\\
1652: && -\demi\int (\nabla\cdot\bar v)(|\nabla\bar\varphi|^2-2\nabla\bar\varphi\cdot \nabla\varphi)+ C\int |\nabla\varphi-\nabla\bar\varphi|^2\\
1653: && + C\epsilon^3 + \Dt Q\;\;\;
1654: \en
1655: with $\left|Q(t)\right|\leq C\epsilon^4$ for $t\in[0,T]$.
1656: From (\ref{1boundep}) we have $\ds\|\nabla\cdot v\|_{\Linf([0,T]\times\Td)}\leq C\epsilon$ and $\ds\|\nabla\bar\varphi\|_{\Linf([0,T]\times\Td)}\leq C\epsilon$,
1657: whereas (\ref{1boundvma}) yields $\ds\int \left|\nabla\varphi\right|^2 \leq C\epsilon^2$.
1658: Note that we also have
1659: \be
1660: &&-\int f(t,x,\xi)(\xi-\bar v)\nabla \bar v(\xi-\bar v)+(\nabla\varphi-\nabla\bar\varphi)\nabla\bar v (\nabla\varphi-\nabla\bar\varphi)\\
1661: &&+ C\int |\nabla\varphi-\nabla\bar\varphi|^2\leq C G\epu.
1662: \en
1663: We conclude that
1664: \be
1665: \Dt (G\epu-Q)\leq C ((G\epu-Q)+ \epsilon^3),
1666: \en
1667: and the conclusion of Theorem \ref{1vmavp} follows by Gronwall's lemma.
1668:
1669: $\hfill \Box$
1670:
1671:
1672:
1673:
1674: \bigskip
1675: \noindent
1676: {\bf Proof of identity (\ref{1dtH3}):}
1677: Here we have to compute:
1678: \be
1679: R=\frac{1}{\epsilon}\int \bar v(x+\epsilon\nabla\varphi)\cdot\nabla\varphi - (\bar v \nabla\bar \varphi)(x+\epsilon\nabla\varphi)
1680: + (1-\epsilon \Delta \bar\varphi)(\bar v \cdot\nabla\bar \varphi - \bar v\cdot\nabla\varphi)
1681: \en
1682: Using Lemma \ref{1Taylor} we have:
1683: \be
1684: R=&&\frac{1}{\epsilon}\int \bar v\cdot\nabla\varphi -\bar v \cdot\nabla\bar \varphi+\bar v \cdot\nabla\bar \varphi - \bar v\cdot\nabla\varphi\\
1685: &&+\int \nabla\bar v \cdot\nabla\varphi \nabla\varphi -\nabla(\bar v \nabla\bar \varphi)\nabla\varphi
1686: - \bar v \nabla\bar \varphi\Delta\bar \varphi +\bar v \nabla\varphi\Delta\bar \varphi\\
1687: &&+\int(<\nabla\bar v>-\nabla\bar v)\nabla\varphi\nabla\varphi
1688: - \epsilon<\nabla^2(\bar v \nabla\bar\varphi)>\nabla\varphi\nabla\varphi.
1689: \en
1690: We see that the first line cancels. Then we show that the last line is bounded by $C\epsilon^3$.
1691: \\
1692: This is obvious for the last term since from (\ref{1boundei}, \ref{1boundvma}) we have
1693: $\ds\|\bar v\|_{W^{2,\infty}} \leq C$, and $\ds \|\nabla \bar\varphi \|_{W^{2,\infty}}\leq C\epsilon$.
1694: \\
1695: Then for the first term we have the following lemma:
1696: \begin{lemme}
1697: \label{1delta}
1698: We define
1699: $$\Delta=\int( <\nabla\bar v>(x) - \nabla \bar v(x))\nabla\varphi\nabla\varphi dx,$$
1700: then one has:
1701: $$\left|\Delta\right|\leq C\epsilon^{10/3}+ C \int|\nabla\varphi-\nabla\bar\varphi|^2 dx.$$
1702: \end{lemme}
1703: {\bf Proof:}
1704: \noindent
1705: First we show that if $\Theta(R)=\int_{\{\left|\nabla\varphi\right|\geq R\}}\left|\nabla\varphi\right|^2$,
1706: $$\Theta(R)\leq C \int\left|\nabla\varphi-\nabla\bar\varphi\right|^2 +\frac{C\epsilon^4}{R^2}.$$
1707:
1708: \noindent
1709: Proof: $\int \left|\nabla\varphi\right|^2\leq C\epsilon^2$, implies that
1710: $$\mbox{meas}\{\left|\nabla\varphi\right|\geq R\}\leq C(\frac{\epsilon}{R})^2.$$
1711: Since $\left|\nabla\bar\varphi(t,x)\right|\leq \epsilon$ for $(t,x)\in [0,T'x]\times\Td$
1712: \begin{eqnarray*}
1713: &&\Theta(R)\leq\int_{\{\left|\nabla\varphi\right|\geq R\}}\left|\nabla\bar\varphi\right|^2+\int_{\{\left|\nabla\varphi\right|\geq R\}}\left|\nabla\varphi-\nabla\bar\varphi\right|^2\\
1714: &&\leq \frac{C\epsilon^4}{R^2}+\int\left|\nabla\varphi-\nabla\bar\varphi\right|^2.
1715: \end{eqnarray*}
1716:
1717: \medskip
1718: \noindent
1719: Then we have
1720: \begin{eqnarray*}
1721: &&\Delta\leq C \Theta(R)+ \int_{\left|\nabla\varphi\right|\leq R}\left|<\nabla\bar v>(x) - \nabla\bar v(x)\right|\nabla\varphi\nabla\varphi\\
1722: \mbox{ with }&& \left|<\nabla\bar v>(x) - \nabla\bar v(x)\right|\leq C \epsilon\left|\nabla\varphi\right|\\
1723: \mbox{thus } &&\Delta\leq C\epsilon\int_{\left|\nabla\varphi\right|\leq R}\left|\nabla\varphi\right|^3 + C\Theta(R)\\
1724: &&\leq C\left(\epsilon R \int\left|\nabla\varphi\right|^2 +\frac{\epsilon^4}{R^2}+\int \left|\nabla\varphi-\nabla\bar\varphi\right|^2\right)\\
1725: &&\leq C\left(\epsilon^3 R + +\frac{\epsilon^4}{R^2}+\int \left|\nabla\varphi-\nabla\bar\varphi\right|^2\right)
1726: \en
1727: for all R, so for $R=\epsilon^{(1/3)}$ one obtains:
1728: \be
1729: &&\Delta\leq C\epsilon^{10/3}+C\int \left|\nabla\varphi-\nabla\bar\varphi\right|^2.
1730: \end{eqnarray*}
1731: $\hfill \Box$
1732:
1733: \bigskip
1734: \noindent
1735: Thus we have shown that $R=S+O(\epsilon^3)$,
1736: and $\displaystyle S=\Sigma_{k=1}^{6}T_k$
1737: where each $T_k$ is given by:
1738: \be
1739: && T_1= \partial_j\bar v_i\partial_j\varphi\partial_i\varphi\\
1740: && T_2= -\partial_j\bar v_i\partial_j\varphi\partial_i\bar\varphi\\
1741: && T_3= -\bar v_i\partial_{ij}\bar\varphi\partial_j\varphi\\
1742: && T_4= \partial_j\bar v_i\partial_j\bar\varphi\partial_i\bar\varphi\\
1743: && T_5= \bar v_i\partial_{ij}\bar\varphi\partial_j\bar\varphi\\
1744: && T_6= \bar v_i\partial_{jj}\bar\varphi\partial_i\varphi\\
1745: \en
1746: where we have used Einstein's convention for repeated indices.
1747: First we have
1748: \be
1749: T_5=-\demi\int(\nabla\cdot \bar v) \left|\nabla\bar\varphi\right|^2
1750: \en
1751: \be
1752: T_1+T_2+T_4=\int \partial_j\bar v_i(\partial_j\varphi-\partial_j\bar\varphi)(\partial_i\varphi-\partial_i\bar\varphi)+ T_7
1753: \en
1754: with $T_7=\int \partial_j\bar v_i\partial_j\bar\varphi\partial_i\varphi$.
1755: \be
1756: T_6=-\int \partial_i\bar v_i\partial_{jj}\bar\varphi\varphi + \bar v_i\partial_{ijj}\bar\varphi\varphi
1757: \en
1758: and
1759: \be
1760: -\int \bar v_i\partial_{ijj}\bar\varphi\varphi=\int \partial_j\bar v_i\partial_{ij}\bar\varphi\varphi
1761: +\bar v_i\partial_{ij}\bar\varphi\partial_j\varphi
1762: \en
1763: thus
1764: \be
1765: T_6=\int -(\nabla\cdot\bar v)\Delta\bar\varphi\;\varphi + T_8 -T_3
1766: \en
1767: with $T_8=\int\partial_j\bar v_i\partial_{ij}\bar\varphi\varphi $.
1768: Then
1769: \be
1770: T_8 = &&-\int \partial_j\bar v_i\partial_j\bar\varphi\partial_i\varphi + \partial_{ij}\bar v_i\partial_j\bar\varphi\varphi\\
1771: &&=-T_7+ \int \nabla\cdot\bar v(\Delta \bar\varphi\varphi+\nabla\bar\varphi \nabla\varphi)
1772: \en
1773: and finally we obtain
1774: \be
1775: S(t)=&&\int \nabla\bar v (\nabla\bar\varphi-\nabla\varphi)(\nabla\bar\varphi-\nabla\varphi)
1776: -\demi(\nabla\cdot \bar v)\left| \nabla\bar\varphi-\nabla\varphi\right|^2\\
1777: &&+\demi\int(\nabla\cdot \bar v)\left| \nabla\varphi\right|^2
1778: \en
1779: and the identity (\ref{1dtH3}) is proved.
1780:
1781: $\hfill \Box$
1782: \eject
1783: \newpage
1784:
1785:
1786:
1787:
1788:
1789: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1790: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1791:
1792: \section{Appendix}
1793: \vspace{1cm}
1794:
1795:
1796: \subsection*{Existence and uniqueness of solutions to second order ODE's with BV field}
1797: In this section we prove existence and a.e. uniqueness for ordinary differential
1798: equations of the form:
1799: \beq\label{1ODE}
1800: \Dt\left(\begin{array}{c} X\\ V\end{array}\right)=\left(\begin{array}{c} V \\ E(t,X)\end{array}\right)
1801: \enq
1802: for $X\in\Td,\;Y\in \Td$, and where the field $E$ belongs to $\Linf(]0,T[\times\Td)\cap L^1(]0,T[,BV(\Td))$. We work in the flat torus for simplicity, but our results are still
1803: valid in an open subset of $\Rd$.
1804: This result is an adaptation of the proof of \cite{DL} that uses the result
1805: of \cite{Bo} on renormalized solutions of transport equations.
1806: \\
1807: Remark: After this proof was written, the authors learned of a result by L. Ambrosio (\cite{Am}) that extends the results of \cite{DL} to transport equations when the vector field is in $BV$ with bounded divergence.
1808:
1809:
1810: \subsubsection*{Renormalized solutions for Vlasov equations with BV field}
1811: Theorem 3.4 in \cite{Bo} adapted to the periodic case sates that if
1812: $f \in L^{\infty}(]0,T[\times \Td \times\Rd)$ satisfies:
1813: \begin{eqnarray}\label{1kinebouchut}
1814: \frac{\partial f}{\partial t}+\nabla_x\cdot\left(\xi f\right)
1815: +\nabla_{\xi}\cdot\left( E(t,x)f\right)=0,
1816: \end{eqnarray}
1817: with $E(t,x)\in L^1(]0,T[\times\Td)\cap L^{1}(]0,T[,BV(\Td))$, then for all $g$ Lipschitz continuous we have
1818: \begin{eqnarray*}
1819: \frac{\partial g(f)}{\partial t}+\nabla_x\cdot\left(\xi g(f)\right)
1820: +\nabla_{\xi}\cdot\left( E(t,x)g(f)\right)=0.
1821: \end{eqnarray*}
1822: The property of renormalization implies that
1823: \begin{itemize}
1824: \item solutions to (\ref{1kinebouchut})
1825: with initial data in $L^{\infty}_{loc}(\Td \times\Rd)$ belong to
1826: \\$C(]0,T[, L^p_{loc}(\Td \times\Rd))$ for any $1\leq p <\infty$,
1827: \item solutions to (\ref{1kinebouchut}) with prescribed initial data in
1828: $L^{\infty}(\Td \times\Rd)$ are a.e. unique,
1829: \item if $E_n$ converges to $E$ in $L^1(]0,T[\times \Td)$ then the solutions of (\ref{1kinebouchut}) with $E_n$ instead of $E$ converge to the solution of (\ref{1kinebouchut}).
1830: \end{itemize}
1831: We notice that equation (\ref{1principaleper}) satisfies the assumptions of the Theorem,
1832: and thus will have the renormalization property. This renormalization
1833: property was used in \cite{DL} to obtain a.e. uniqueness for solutions
1834: of the corresponding ODE's. Indeed, the ODE
1835: \be
1836: &&\dt X(t,s,x)=b(t,X)\\
1837: &&X(s,s,x)=x
1838: \en
1839: is associated to the transport equation:
1840: \be
1841: \dt u +b(t,x).\nabla u=0
1842: \en
1843: whose solutions satisfy for all $(t,s)\in]0,T[$
1844: \be
1845: u(t,X(t,s,x))=u(s,x).
1846: \en
1847: We extend this consequence to the case of second order equations,
1848: with $BV$ acceleration field.
1849: To the kinetic equation
1850: \beq\label{1cineapp}
1851: \dt f +\xi\cdot\nabla_x f + E(t,x)\cdot\nabla_{\xi} f=0
1852: \enq
1853: we associate the second order ODE (\ref{1ODE}) which can rewritten as
1854: $\partial_{tt}{X}=E(t,X)$. The result is then the following:
1855:
1856:
1857: \begin{theo}\label{1theo9}
1858: Let $E(t,x) \in \Linf(]0,T[\times \Td)\cap L^1(]0,T[,BV(\Td))$,\\
1859: then the ODE
1860: \begin{eqnarray}\label{1ODE2}
1861: &&\partial_{tt}{X}(t,s,x,\xi)=E(t,X)\\
1862: &&(X(s,s,x,\xi),\partial_t{X}(s,s,x,\xi))=(x,\xi)
1863: \end{eqnarray}
1864: admits an a.e. unique solution.
1865: \end{theo}
1866: Remark: Here almost everywhere must be understood for the Lebesgue measure of $\R^6$.
1867: \\
1868: {\bf Proof of Theorem \ref{1theo9}:}
1869: We know that through equation (\ref{1ODE}) equation (\ref{1ODE2})
1870: can be considered as a first order differential equation.
1871: Let us first consider the
1872: case where $E$ is smooth. Note $Y\in \Td\times\Rd $ (resp. $y$) for $(X,V)$ (resp. for $(x, \xi)$)
1873: and $B\in \Rd\times\Rd$ for $(\xi,E)$. Then for all $s\in ]0,T[$, $Y$ solves:
1874: \begin{eqnarray}
1875: &&\dt Y(t,s,y)=B(t,Y(t,s,y))\label{1edo}\\
1876: &&Y(s,s,y)=y\label{1edoin}
1877: \end{eqnarray}
1878: Then for all $t,t_1,t_2,t_3\in ]0,T[$ we have the following:
1879: \begin{eqnarray*}
1880: &&Y(t_3,t_2,Y(t_2,t_1,y))=Y(t_3,t_1,y)\\
1881: &&Y(t,t,y)=y\\
1882: &&Y(t_1,t_2,Y(t_2,t_1,y))=y.
1883: \end{eqnarray*}
1884: Differentiating the last equation with respect to $t_2$ yields:
1885: \begin{eqnarray}
1886: &&\partial_{s}Y(t,s,y)+\nabla_y Y(t,s,y)\cdot B(s,y)=0\label{1transport}\\
1887: &&Y(t,t,y)=y\label{1transportin}.
1888: \end{eqnarray}
1889: $Y_t(s,y)=Y(t,s,y)$ thus solves a transport equation which is nothing but equation (\ref{1cineapp}). Using Theorem \ref{1Bouchut} we know that for all $g:\R^{2d}\rightarrow\R$ Lipschitz continuous, $g(t,s,y)=g_0(Y(t,s,y))$ is the unique solution of
1890: \begin{eqnarray}
1891: &&\partial_{s}g(t,s,y)+\nabla_y g(t,s,y)\cdot B(s,y)=0\label{1transg}\\
1892: &&g(t,t,y)=g_0(y)\label{1transgin}.
1893: \end{eqnarray}
1894: Now we show existence and uniqueness for solutions of (\ref{1edo},\ref{1edoin}).
1895: Let $t$ and $s$ be fixed. Let us consider a regularization $E_n$ of the the field $E$ and set
1896: $B_n=(\xi,E_n)$. We have
1897: \begin{itemize}
1898: \item$t\rightarrow Y_{1,n}(t,s,y)$ that satisfies (\ref{1edo},\ref{1edoin})
1899: \item$s\rightarrow Y_{2,n}(t,s,y)$ that satisfies (\ref{1transport},\ref{1transportin}).
1900: \end{itemize}
1901: From the stability Theorem 2.4 in \cite{DL} we know that the whole sequence \\$t\rightarrow Y_{2,n}(t,s,.)$ converges in $C(]0,T[,L^p_{loc}(\Rd\times\Td))$ to $t\rightarrow Y_{2}(t,s,.)$, the unique renormalized solution of (\ref{1transport},\ref{1transportin}). Thus for fixed $t$ the whole sequence $Y_{2,n}(t,s,.)$ converges strongly in $L^p_{loc}(\Rd\times\Td)$. Now since for every $n$ we have $Y_{1,n}(t,s,y)=Y_{2,n}(t,s,y)$
1902: the same property holds for $Y_{1,n}(s,t,.).$ Now we can pass to the limit in the term $B_n(t,Y_{1,n}(t,s,y))$. Indeed, by density of $C_c^{\infty}$ functions in $L^1,$ if we have $E_s\in C_c^{\infty}$ approximating $E$ then
1903: \be
1904: &&\|B(t,Y_n(t,s,y))-B(t,Y(t,s,y))\|_{L^1}\\
1905: &\leq&\|B(t,Y_n(t,s,y))-B_s(t,Y_n(t,s,y))\|_{L^1}\\
1906: &+&\|B_s(t,Y_n(t,s,y))-B_s(t,Y(t,s,y))\|_{L^1}\\
1907: &+&\|B(t,Y(t,s,y))-B_s(t,Y(t,s,y))\|_{L^1}
1908: \en
1909: The second term goes to 0 because of the strong convergence of $Y_n$, the first and the third go to 0 because $Y$ and $Y_n$ are measure preserving mappings, and so for example $\|B(t,Y(t,s,y))-B_s(t,Y(t,s,y))\|_{L^1}=\|B(t,y)-B_s(t,y)\|_{L^1}$.
1910: So finally we have
1911: \be
1912: &&\|B_n(t,Y_n(t,s,y))-B(t,Y(t,s,y))\|_{L^1}\\
1913: \leq&&\|B_n(t,Y_n(t,s,y))-B(t,Y_n(t,s,y))\|_{L^1}\\
1914: &&+\|B(t,Y_n(t,s,y))-B(t,Y(t,s,y))\|_{L^1}
1915: \en
1916: that goes to 0 and we can pass to the limit in the equation (\ref{1edo},\ref{1edoin}) and the existence of a solution
1917: to (\ref{1edo},\ref{1edoin}) is proved.
1918:
1919: \medskip
1920: \noindent
1921: To obtain uniqueness, we argue as in \cite{DL}. Any function of the form $g_0(Y(t,s,y))$ is a solution of
1922: (\ref{1transg},\ref{1transgin}),
1923: thus by uniqueness of the solution of the transport equation we obtain uniqueness of the ODE.
1924:
1925: $\hfill\Box$
1926:
1927: \subsubsection*{A remark on ODE's of second order}
1928: In this section, we want to solve the Cauchy problem for:
1929: \begin{eqnarray*}
1930: &&\partial_{tt}{X}(t,x)=E(t,X)\\
1931: &&(X(0,x)\; ,\partial_t{X}(0,x))=(x,v(x))
1932: \end{eqnarray*}
1933: with $E$ as above. We are thus interested in monokinetic initial data.
1934:
1935: \begin{theo}
1936: for all $v^0(x)$ vector field on $\Td$, and for Lebesgue almost every $\delta v\in\mathbb R^d$, there exists an a.e. unique solution to
1937: \begin{eqnarray*}
1938: &&\partial_{tt}{X}(t,x)=E(t,X(t,x))\\
1939: &&(X(0,x)\;,\partial_t{X}(0,x))=(x,v^0(x)+\delta v)
1940: \end{eqnarray*}
1941: \end{theo}
1942:
1943: \medskip
1944: \noindent
1945: {\bf Proof:} Let $g(x,\xi)$ be the indicator function of the set of those $(x;\xi)$ such that the trajectory coming from x is not well defined. We just have to prove that for a.e. $\delta v\in \mathbb R^d$ we have $\int g(x,v^0(x)+\delta v)dx=0,$ which is true because
1946: $$\int g(x,v^0(x)+\xi)dx d\xi=\int g(x,\xi)dx d\xi=0.$$
1947:
1948:
1949: \paragraph{Stability}
1950: Using the fact that for $E_n$ converging to $E$ in $L^1$ with\\
1951: $E\in L^1(]0,T[,BV(\Td))$, we have $X_n(t,x,v)\rightarrow X(t,x,v)$
1952: in $C([0,T],L^p)$, we have then, for all t, for almost every $\delta v,$ $X_n(t,x,v^0(x)+\delta v) \rightarrow X(t,x,v^0(x)+\delta v)$ in $L^p$. Thus we have
1953:
1954: \begin{theo}
1955: If $E_n$ converges to $E$ in $L^1$ let $X_n$ be solution of
1956: \begin{eqnarray*}
1957: &&\partial_{tt}{X_n}(t,x)=E_n(t,X_n(t,x))\\
1958: &&(X_n(0,x),\partial_t{X_n}(0,x))=(x,v^0(x)+\delta v)
1959: \end{eqnarray*}
1960: then for all t, for almost every $\delta v,$ $X_n$ converges in $L^p(\mathbb R^3)-s$ to a solution (unique for almost every $\delta v$) of
1961: \begin{eqnarray*}
1962: &&\partial_{tt}{X}(t,x)=E(t,X)\\
1963: &&(X(0,x),\partial_t{X}(0,x))=(x,v^0(x)+\delta v).
1964: \end{eqnarray*}
1965: \end{theo}
1966:
1967: \subsection*{Control of macroscopic density in kinetic\\ equations}
1968: We prove here Lemma \ref{1densite}:
1969: \begin{lemme}\label{1densite2}
1970: Let $f\in \Linf([0,T]\times\Td\times\Rd)$ satisfy
1971: \begin{eqnarray}
1972: &&\frac{\partial f}{\partial t}+\nabla_x\cdot\left(\xi f\right)
1973: +\nabla_{\xi}\cdot\left(E(t,x) f\right)=0\label{1cinetik1ap}\\
1974: &&f(0,.,.)=f^0\label{1cinetik2ap}
1975: \end{eqnarray}
1976: with $E\in L^1([0,T]; BV(\Td))$ and
1977: \beq
1978: \|E\|_{\Linf([0,T]\times\Td)} \leq F.\label{1Ebv}
1979: \enq
1980: Let the initial condition $f_0$ be such that:
1981: $$a(x,\xi)\leq f(0,x,\xi)\leq b(x,\xi),$$
1982: with $\rho_a(x)=\int a(x,\xi)d\xi\geq m>0$ and $\rho_b(x)=\int b(x,\xi)d\xi\leq M<\infty$ and $a,b$ satisfying
1983: \beq
1984: |\nabla_{x,\xi}a,b|\leq \frac{c}{1+|\xi|^{d+2}}\label{1lipab}.
1985: \enq
1986: Then there exists a constant $R>0$ such that
1987: \be
1988: (\rho_a(x)-Rt)\leq\rho(t,x)\leq(\rho_b(x)+Rt).
1989: \en
1990: \end{lemme}
1991: {\bf Proof: }
1992: First suppose that the force field and the initial data are smooth. For equation (\ref{1cinetik1ap},\ref{1cinetik2ap}) we can exhibit characteristics $(x,\xi)(t;t_0,x_0,\xi_0),$ giving the evolution of the particles in the phase space. We have $f(t,x,\xi)=f(t_0,x_0,\xi_0).$ Since the initial data is compactly supported and the force field is bounded in the $\Linf$ norm, we have
1993: $$|\xi-\xi_0|\leq F|t-t_0|,$$
1994: $$|x-x_0|\leq (|\xi_0|+\frac{F}{2}|t-t_0|)|t-t_0|.$$
1995:
1996: If for $t=0$ we have $a(x,\xi)\leq f(0,x,\xi)\leq b(x,\xi)$ then
1997: \begin{eqnarray*}
1998: &&\underline{A}(t,x,\xi)\leq f(t,x,\xi)\leq\overline{B}(t,x,\xi)\\
1999: &&\underline{A}(t,x,\xi)=\inf_{|\sigma_1|,|\sigma_2|\leq 1}a(x+|t-t_0|(\xi +\frac{F}{2}|t-t_0|)\sigma_1, \xi+F|t-t_0|\sigma_2)\\
2000: &&\overline{B}(t,x,\xi)=\sup_{|\sigma_1|,|\sigma_2|\leq 1}b(x+|t-t_0|(\xi +\frac{F}{2}|t-t_0|)\sigma_1, \xi+F|t-t_0|\sigma_2).
2001: \end{eqnarray*}
2002: Using (\ref{1lipab}) and integrating in $\xi$ we find thus a constant $R=R(F,C,d)$ such that for $t-t_0\leq 1$ we have:
2003: \be
2004: \rho_a(x)-R |t-t_0|\leq \rho(t,x)\leq \rho_b(x)+R|t-t_0|.
2005: \en
2006:
2007: \medskip
2008: \noindent
2009: Next we need to show that the solution of the regularized equation converges to the solution we are studying: this result comes from the uniqueness of the solution to (\ref{1cinetik1ap},\ref{1cinetik2ap}) which is a consequence of the renormalization property. Indeed since $E$ is bounded in $BV$ the system (\ref{1cinetik1ap},\ref{1cinetik2ap}) admits a unique renormalized solution and the sequence of approximate solutions converge in $C([0,T],L^p_{x,\xi})$ for $1\leq p < \infty$ thus the bounds obtained above are preserved.
2010:
2011: $\hfill$ $\Box$
2012:
2013: \subsection*{Regularity of the polar factorization on the flat torus}
2014: Here we deduce from \cite{Mc2}, \cite{Co} and \cite{Ca1}, \cite{Ca2}, \cite{Ca3}
2015: the Theorem \ref{1regper}.
2016: \begin{theo}\label{1regperap}
2017: If $\rho \in C^{\alpha}(\Td)$ with $0<m\leq\rho\leq M$ is a probability measure on $\Td$ then $\Psi=\Psi[\rho]$ (see Definition \ref{1MaTdrho}) is a classical solution of
2018: \begin{eqnarray}
2019: \det D^2\Psi=\rho\label{1mongeperap}
2020: \end{eqnarray}
2021: and satisfies:
2022: \begin{eqnarray}
2023: &&\|\nabla\Psi(x)-x\|_{\Linf}\leq C(d)=\sqrt d /2\\
2024: &&\|D^2\Psi\|_{C^{\alpha}}\leq K(m,M,\|\rho\|_{C^{\alpha}})
2025: \end{eqnarray}
2026: \end{theo}
2027: {\bf Proof of Theorem \ref{1regperap}: }Consider $\rho$ a $\Zd$ periodic probability measure, satisfying
2028: \beq
2029: 0<m\leq\rho\leq M,\label{1mM}
2030: \enq
2031: and $\Phi[\rho]$ as in Definition \ref{1MaTdrho}.
2032: First it is shown in \cite{Co} that
2033: \beq
2034: |\nabla\Phi[\rho](x)-x|\leq C(d).\label{1LinfCo}
2035: \enq
2036: It follows that the strict convexity argument of \cite{Ca1} applies: indeed
2037: if $\Phi=\Phi[\rho]$ is not strictly convex its graph contains a line and this
2038: contradicts (\ref{1LinfCo}). Moreover since $\Phi-|x|^2/2$ is globally Lipschitz and periodic there exists $N(d)$ such that $\|\Phi-|x|^2/2\|_{\Linf}\leq N(d)$.
2039: It follows then that there exists $0 <r(d)\leq R(d)$ and $M(d)$ such that
2040: \beq
2041: B(r(d))\subset \{\Phi-\Phi(0) \leq M(d)\} \subset B(R(d))
2042: \enq
2043: It remains to show that our solution is a solution in the Aleksandrov sense
2044: of the Monge-Amp\`ere equation
2045: \be
2046: m\leq \det D^2 \Phi\leq M.
2047: \en
2048: This is a direct consequence with minor changes (to adapt to the periodic case) of Lemma 2 of \cite{Ca3}.
2049: Then, normalizing $\Phi$ to $\tilde\Phi=\Phi-\Phi(0)-M(d)$
2050: it follows that $\tilde\Phi$ is a solution of
2051: \be
2052: &&\rho(\nabla\tilde\Phi)\det D^2\tilde\Phi=1\\
2053: &&\tilde\Phi = 0\;\;\text{ on }\partial\Omega\\
2054: &&B(r(d))\subset \Omega \subset B(R(d))
2055: \en
2056: Thus the interior regularity results of \cite{Ca2} apply uniformly to
2057: all $\Phi[\rho]$ with $\rho$ satisfying (\ref{1mM}) and $\|\rho\|_{C^{\alpha}(\Td)}$
2058: bounded and Theorem \ref{1regper} follows.
2059:
2060: $\hfill$ $\Box$
2061:
2062:
2063:
2064:
2065: \subsection*{Acknowledgments}
2066: The authors acknowledge the support
2067: of the European IHP network "HYKE" HPRN-CT-2002-00282 and
2068: the LRC CEA-Cadarache/UNSA.
2069:
2070:
2071: \bibliography{vma-gafa-biblio}
2072: \end{document}
2073:
2074:
2075:
2076:
2077:
2078:
2079:
2080:
2081:
2082:
2083:
2084:
2085:
2086:
2087:
2088:
2089:
2090:
2091:
2092:
2093:
2094:
2095:
2096:
2097:
2098:
2099: