math0701857/lost.tex
1: \documentclass[leqno]{amsart}
2:        
3: \usepackage{amsthm,amsmath,amsfonts,amssymb}
4: %\usepackage[notref,notcite]{showkeys}
5: 
6: \def\C{{\mathbb C}}
7: \def\R{{\mathbb R}}
8: \def\N{{\mathbb N}}
9: \def\Z{{\mathbb Z}}
10: \def\T{{\mathbb T}}
11: \def\Q{{\mathbb Q}}
12: \def\Sch{{\mathcal S}} 
13: \def\virgp{\raise 2pt\hbox{,}}
14: \def\bv{{\bf v}}
15: \def\bu{{\bf u}}
16: \def\bw{{\bf w}}
17: \def\({\left(}
18: \def\){\right)}
19: \def\<{\left\langle}
20: \def\>{\right\rangle}
21: \def\le{\leqslant}
22: \def\ge{\geqslant}
23: \def\les{\lesssim}
24: 
25: \def\Eq#1#2{\mathop{\sim}\limits_{#1\rightarrow#2}}
26: \def\Tend#1#2{\mathop{\longrightarrow}\limits_{#1\rightarrow#2}}
27: 
28: \def\d{{\partial}}
29: \def\a{{\tt a}}
30: \def\b{\beta}
31: \def\g{\gamma}
32: \def\eps{\varepsilon}
33: \def\h{h}
34: \def\l{\lambda}
35: \def\om{\omega}
36: \def\si{{\sigma}}
37: \def\F{\mathcal F}
38: \def\O{\mathcal O}
39: 
40: \DeclareMathOperator{\RE}{Re}
41: \DeclareMathOperator{\IM}{Im}
42: \DeclareMathOperator{\Op}{\mathrm{Op}}
43: \DeclareMathOperator{\DIV}{\mathrm{div}}
44: 
45: \def\defn{\mathrel{:=}}
46: 
47: \theoremstyle{plain}
48: \newtheorem{theorem}{Theorem}[section]
49: \newtheorem{lemma}[theorem]{Lemma}
50: \newtheorem{corollary}[theorem]{Corollary}
51: \newtheorem{proposition}[theorem]{Proposition}
52: 
53: 
54: \theoremstyle{definition}
55: \newtheorem{definition}[theorem]{Definition}
56: 
57: \theoremstyle{remark}
58: \newtheorem{remark}[theorem]{Remark}
59: 
60: \numberwithin{equation}{section}
61: 
62: 
63: \title[Loss of regularity for NLS]{Loss of regularity for
64:  supercritical nonlinear Schr\"odinger equations}
65:  \author[T. Alazard]{Thomas Alazard}
66: \address{CNRS \& Universit\'e
67:   Paris-Sud\\ Math\'ematiques\\ B\^at. 425\\ 91405 
68:   Orsay cedex\\ France}
69: \email{Thomas.Alazard@math.cnrs.fr}
70: \author[R. Carles]{R{\'e}mi Carles}
71: \address{CNRS \& Universit\'e Montpellier~2\\ Math\'ematiques\\ CC 051\\ 
72:    Place Eug\`ene Bataillon\\ 34095
73:   Montpellier cedex 5\\ France}
74: \email{Remi.Carles@math.cnrs.fr}
75: 
76: \begin{document}
77: 
78: 
79: 
80: \begin{abstract}
81: We consider the nonlinear Schr\"odinger equation with defocusing,
82: smooth, 
83: nonlinearity. Below the critical Sobolev regularity, it is known that
84: the Cauchy problem is ill-posed. We show that this is even worse,
85: namely that there is a loss of regularity, in the spirit of the
86: result due to 
87: G.~Lebeau in the case of the wave equation. As a consequence, the
88: Cauchy problem for energy-supercritical equations is not well-posed
89: in the sense of Hadamard. 
90: We  reduce the problem to a
91: supercritical WKB analysis. For super-cubic, smooth nonlinearity,
92: this analysis is new, and relies on the introduction of a modulated
93: energy functional \emph{\`a la} Brenier. 
94: \end{abstract}
95: 
96: \subjclass{Primary 35Q55; Secondary 35A07, 35B33, 35B65,  76Y05, 81Q05, 81Q20}
97: \keywords{Nonlinear Schr\"odinger equation; Cauchy problem;
98:   supercritical analysis; semi-classical analysis; compressible Euler
99:   equation;  modulated energy estimates}
100: \thanks{Support by the ANR project SCASEN is acknowledged.} 
101: \maketitle
102: 
103: 
104: 
105: 
106: \section{Introduction}
107: \label{sec:intro}
108: 
109: We consider the following defocusing nonlinear Schr\"odinger equation
110: on $\R^n$: 
111: \begin{equation}
112:   \label{eq:nls}
113:   i\d_t \psi+\frac{1}{2}\Delta \psi = |\psi|^{2\si}\psi\quad ;\quad \psi_{\mid
114:   t=0}=\varphi, 
115: \end{equation}
116: where $\si\ge 1$ is an integer, so that the nonlinearity is
117: smooth. It is well-known that the critical Sobolev regularity
118: corresponds to the value given by scaling arguments,
119: \begin{equation*}
120:   s_c\defn \frac{n}{2} - \frac{1}{\si}\cdot 
121: \end{equation*}
122: Throughout this paper, we assume $s_c>0$. If $\varphi\in H^s(\R^n)$ with
123: $s \ge s_c$, then the Cauchy problem \eqref{eq:nls} is locally
124: well-posed in $H^s(\R^n)$ \cite{CW90}. On the other hand, if
125: $s<s_c$, then the Cauchy problem \eqref{eq:nls} is ill-posed
126: \cite{CCT2} (see also the appendices in \cite{BGTENS,CaARMA}). 
127: The
128: worst phenomenon proved in \cite{CCT2} is the norm inflation. For
129: $0<s<s_c$, one can 
130: find a sequence $(\psi^\h)_{0<\h\le 1}$ of solutions to
131: \eqref{eq:nls} and $0<t^\h\to 0$, such that $\varphi^\h \in
132: \Sch(\R^n)$  and  
133: \begin{equation*}
134:   \|\varphi^\h\|_{H^s}\Tend \h 0 0 \quad ;\quad
135:   \|\psi^\h(t^\h)\|_{H^s}\Tend \h 0 +\infty.
136: \end{equation*}
137: In this paper, we prove the stronger result:
138: \begin{theorem}\label{theo:new}
139: Let $\si\ge 1$. Assume that $s_c=n/2-1/\si >0$, and let
140: $0<s<s_c$. 
141: There exists a family $(\varphi^\h)_{0<\h \le 
142:   1}$ in ${\mathcal S}({\mathbb R}^n)$ with 
143: \begin{equation*}
144:   \|\varphi^\h\|_{H^{s}({\mathbb R}^n)} \to 0 \text{ as
145:   }\h \to 0, 
146: \end{equation*}
147: a solution $\psi^\h$ to
148: \eqref{eq:nls} and $0<t^\h \to 0$, such that: 
149: \begin{equation*}
150:   \|\psi^\h(t^\h)\|_{H^{k}({\mathbb R}^n)} \to +\infty \text{ as }\h \to
151:  0\, , \ \forall k> \frac{s}{1+\si(s_c-s)}\cdot
152: \end{equation*}
153: \end{theorem}
154: Note that this result is not bound to the case $x\in \R^n$: for
155: instance, it remains valid on a compact manifold, see
156: \S\ref{sec:geometrie}.
157: \smallbreak
158: 
159: In the case $\si =1$ and $n\ge 3$, this result was established in
160:   \cite{CaARMA}. It followed from a supercritical WKB
161:   analysis for the cubic nonlinear Schr\"odinger equation, which had
162:   been justified by E.~Grenier~\cite{Grenier98}. For $\si \ge 2$,
163:   adapting the results of \cite{Grenier98} seems to be a much more
164:   delicate issue, and a rigorous analysis in this setting for $n\le 3$
165:   has been given very recently \cite{AC-BKW}. An important remark, in
166:   the proof of Theorem~\ref{theo:new} that we present here, is that it
167:   is not necessary to justify WKB analysis as precisely as in
168:   \cite{Grenier98}, or \cite{AC-BKW}, to obtain this
169:   loss of regularity. From this point of view, our proof is very
170:   simple. On the other hand, it can be considered as highly
171:   nonlinear: it relies on a \emph{quasilinear} analysis, as opposed
172:   to the \emph{semilinear} analysis in \cite{CCT2} (see also
173:   Remark~\ref{rem:semiquasi} at the end of this paper). In the opinion of
174:   the authors, the proof of Theorem~\ref{theo:new} is at least as
175:   interesting as the result itself. 
176: \begin{remark}
177:  Shortly after this work was completed, an alternative proof was given
178:  by L.~Thomann \cite{ThomannAnalytic}, based on the justification of
179:  WKB analysis in an 
180:  analytic setting. This approach allows to consider focusing
181:  nonlinearities (the nonlinearity is treated as a semilinear
182:  perturbation in spaces based on analytic regularity), unlike the
183:  method followed in the present paper. On 
184:  the other hand, the virial identity shows that for supercritical
185:  focusing nonlinearities, blow-up can happen for arbitrary small data
186:  in $H^s$ ($s<s_c$) and arbitrary small times
187:  (see e.g. \cite[Exercise~3.63]{TaoDisp}).       
188: \end{remark}
189: \smallbreak
190: 
191: 
192: This result is to be compared with the main result in
193: \cite{Lebeau05}, which we recall with notations adapted to make the
194: comparison with the Schr\"odinger case easier. For supercritical wave
195: equations  
196: \begin{equation*}
197:   \(\d_t^2-\Delta\)u + u^{2\si+1}=0,
198: \end{equation*}
199: G.~Lebeau shows
200: that one can find a \emph{fixed} initial datum in $H^s$, and a
201: sequence of times $0<t^\h\to 0$, such that the $H^k$ norms of the
202: solution are unbounded along the sequence $t^\h$, for $k\in
203: ]I(s),s]$. The expression for  $I(s)$ is related to the critical
204: Sobolev exponent
205: \begin{equation*}
206:   s_{\text{sob}}=\frac{n}{2}\frac{\si}{\si+1}\virgp
207: \end{equation*}
208: which corresponds to the embedding
209: $H^{s_{\text{sob}}}(\R^n)\subset L^{2\si +2}(\R^n)$. In
210: \cite{Lebeau05}, we find:
211: \begin{equation}\label{eq:Ileb}
212:   I(s)=1\text{ if } 1<s\le s_{\text{sob}}\quad ;\quad 
213: I(s)=\frac{s}{1+\si(s_c-s)}\text{ if } s_{\text{sob}}\le s<s_c.
214: \end{equation}
215: Note that we have
216: \begin{equation}
217:   \label{eq:sature}
218: \frac{s_{\text{sob}}}{1+\si(s_c-s_{\text{sob}})}=1.
219: \end{equation}
220: The approach in \cite{Lebeau05} consists in using an \emph{anisotropic}
221: scaling, as opposed to the isotropic scaling used in
222: \cite{Lebeau01,CCT2}. Compare Theorem~\ref{theo:new} with
223: the approach of \cite{Lebeau05}. Recall that
224: \eqref{eq:nls} has two important (formally) conserved quantities: mass
225: and energy,
226: \begin{equation}
227:   \label{eq:conserv}
228:   \begin{aligned}
229: &M(t) = \int_{\R^n}|\psi(t,x)|^2dx\equiv M(0),\\
230: &E(\psi(t)) = \frac{1}{2}\int_{\R^n}|\nabla \psi(t,x)|^2dx
231: +\frac{1}{\si+1}\int_{\R^n}|\psi(t,x)|^{2\si+2}dx\equiv E(\varphi).
232: \end{aligned}
233: \end{equation}
234: In view of \eqref{eq:sature}, we obtain, for
235: $H^1$-supercritical nonlinearities: 
236: \begin{corollary}\label{cor:energy}
237: Let $n\ge 3$ and $\si >\frac{2}{n-2}$. 
238: There exists  $(\varphi^\h)_{0<\h \le 
239:   1}$ in ${\mathcal S}({\mathbb R}^n)$ with 
240: \begin{equation*}
241: \| \varphi^\h \|_{H^1}+ \| \varphi^\h\|_{L^{2\si+2}}\rightarrow 0
242: \text{ as
243:   }\h \to 0, 
244: \end{equation*}
245: a solution $\psi^\h$ to
246: \eqref{eq:nls} and $0<t^\h \to 0$, such that: 
247: \begin{equation*}
248:   \|\psi^\h(t^\h)\|_{H^{k}({\mathbb R}^n)} \to +\infty \text{ as }\h \to
249:  0\, , \ \forall k>1.
250: \end{equation*}
251: \end{corollary}
252: We thus get the analogue
253: of the result of G.~Lebeau when $I(s)=1$, with the drawback that we
254: consider a \emph{sequence} of initial data only. The information that
255: we don't have for Schr\"odinger equations, and which is available for
256: wave equations, is the finite speed of propagation, that is used in
257: \cite{Lebeau05} to construct a fixed initial datum; see also the
258: discussion in \S\ref{sec:geometrie}.  On the other hand,
259: our approach involves an \emph{isotropic} scaling, which is recalled
260: and generalized in Section~\ref{sec:reduc}. Moreover, our range for $k$ is
261: broader when $1<s<s_{\rm sob}$, and also, we allow the range $0<s\le
262: 1$, for which no analogous result is available for the wave
263: equation. Note that  
264: unlike in \cite{Lebeau05}, we perform no linearization in our
265: analysis (the properties of the analogous linearized operator are not
266: as  interesting
267: in the case of Schr\"odinger equations): despite the fact that
268: for fixed $\eps$, \eqref{eq:nls} is a 
269: semilinear equation, we consider a quasilinear system to
270: prove our main result. 
271: \smallbreak
272: 
273: Before going further into details, let us focus on the notion of solution
274: to \eqref{eq:nls}. In view of Theorem~\ref{theo:new}, we assume that
275: the initial data are in the 
276: Schwartz class: $\varphi\in \Sch(\R^n)$. Then \eqref{eq:nls} has a
277: local smooth solution: for all $s>n/2$, there exists $T_s>0$ such
278: that \eqref{eq:nls} has a unique solution $\psi \in
279: C([-T_s,T_s];H^s)$.  If $n\le 2$, then 
280: \eqref{eq:nls} has a global smooth solution, $\psi \in C(\R;H^s)$ for
281: all $s\ge 0$, and the identities \eqref{eq:conserv} hold for all
282: time. The same is 
283: true when $n=3$ and $\si =1$. These results are established in
284: \cite{GV84}. In the $H^1$-critical three dimensional case ($n=3$ and
285: $\si=2$), it is proved in \cite{CKSTTAnnals} that solutions with $H^s$
286: regularity ($s>1$) remain in $H^s$ for all time; the same is true in
287: the four dimensional case ($n=4$ and $\si=1$), from \cite{RV}. 
288: On the other hand, if the nonlinearity is
289: $H^1$-supercritical ($\si > \frac{2}{n-2}$), then it is
290: not known in general whether the solution remains smooth for all time
291: or not. In Theorem~\ref{theo:new}, for $\si=1$, the solution $\psi^h$
292: is a smooth 
293: solution, that remains smooth up to time $t^h$, thanks to a result due
294: to E.~Grenier \cite{Grenier98}.  \emph{A priori}, the solution we
295: consider in Theorem~\ref{theo:new} is a weak solution:
296: \begin{definition}\label{def:weak}
297:   Let $\varphi \in H^1\cap L^{2\si+2}(\R^n)$. A (global) weak solution to
298:   \eqref{eq:nls}  is a function $\psi \in C(\R;L^2)\cap
299:   L^\infty(\R; H^1\cap L^{2\si+2})\cap C_w (\R; H^1\cap L^{2\si+2})$
300:   solving \eqref{eq:nls} in 
301:   ${\mathcal D}'(\R\times \R^n)$, and such that:
302:   \begin{itemize}
303:   \item $\|\psi(t)\|_{L^2}= \|\varphi\|_{L^2}$, $\forall t\in \R$.
304: \item $E(\psi(t))\le E(\varphi)$, $\forall t\in \R$.
305:   \end{itemize}
306: \end{definition}
307: From \cite{GV85c}, for $\varphi \in \Sch(\R^n)$, \eqref{eq:nls} has a
308: global weak solution. The proof in \cite{GV85c} is based on Galerkin
309: method. We use a different construction, as in \cite{Lebeau05}, which
310: is described in Section~\ref{sec:og}. Note that when the
311: nonlinearity is $H^1$-subcritical, then the weak solution is unique,
312: and coincides with the strong solution. Recall also that the existence
313: of blowing-up solutions in the $H^1$-supercritical case is open so
314: far. On the other hand, if the nonlinearity is focusing, many
315: results are available (see \cite{TaoDisp} for an overview of the
316: subject, and similar problems for other dispersive equations).  
317: \smallbreak
318: 
319: Note that in view of
320: Definition~\ref{def:weak}, Corollary~\ref{cor:energy} is sharp. 
321: \smallbreak
322: 
323: As in \cite{CaARMA}, the idea for the proof of Theorem~\ref{theo:new}
324: consists in reducing the analysis to a supercritical WKB analysis,
325: for an equation of the form:
326: \begin{equation}
327:   \label{eq:nlssemi}
328:   i\eps \d_t u^\eps +\frac{\eps^2}{2}\Delta u^\eps =
329:   |u^\eps|^{2\si}u^\eps\quad ;\quad u^\eps(0,x) = a_0(x). 
330: \end{equation}
331: The parameter $\eps$ goes to zero. The above equation is
332: supercritical as far as geometrical optics is concerned: if one plugs
333: an approximate solution of the form
334: \begin{equation*}
335:   v^\eps \sim e^{i\phi/\eps}\(\a_0+\eps \a_1+\eps^2 \a_2+\ldots\)
336: \end{equation*}
337: into the equation, then closing the systems of equations for
338: $\phi,\a_0,\a_1,\ldots$ is a very delicate issue (see
339: e.g. \cite{CaBKW}). In the case 
340: $\si=1$, this issue was resolved by E.~Grenier
341: \cite{Grenier98}. However, the argument in \cite{Grenier98}
342: relies very strongly on the fact that the nonlinearity is defocusing,
343: and cubic at the origin. In \cite{AC-BKW}, we have
344: proposed an approach that justifies  WKB analysis in Sobolev spaces for
345: \eqref{eq:nlssemi} for any $\si\ge 2$, in space dimension $n\le 3$
346: (higher dimensions could also be considered with the same proof, up to
347: considering sufficiently large values of $\si$). In
348: \cite{PGX93} and \cite{ThomannAnalytic}, WKB analysis was justified in spaces
349: based on analytic regularity, in a periodical setting and for analytic
350: manifolds, respectively. As noticed in
351: \cite{AC-BKW}, analytic regularity is necessary to justify WKB
352: analysis with a focusing nonlinearity. 
353: The analytic regularity essentially allows to
354: view the nonlinearity as a semilinear perturbation, and to construct
355: an approximate solution that solves \eqref{eq:nlssemi} up to a source
356: term of order $e^{-\delta/\eps}$ for some $\delta>0$. Yet,
357: such a justification is not needed to prove
358: Theorem~\ref{theo:new}; see \S\ref{sec:reduc}. In this paper, we use
359: a functional 
360: that yields sufficiently many informations to infer
361: Theorem~\ref{theo:new}. This functional may be viewed as a
362: generalization of the one used in \cite{LinZhang} in the cubic case,
363: following an idea introduced by Y.~Brenier \cite{BrenierCPDE}. The
364: general form for this modulated energy functional was announced in
365: \cite{LinZhang}. However, we will see that making the corresponding
366: analysis rigorous is not straightforward, since we consider weak
367: solutions. 
368: %Finally, note that the discussion presented in
369: %\S\ref{sec:geometrie} certainly fails in the case of analytic
370: %regularity, since compactly supported  approximate solutions are
371: %involved. 
372: \smallbreak
373: 
374: The main idea of the proof of Theorem~\ref{theo:new} consists in
375: noticing that for an $\eps$-independent initial datum $a_0$ in
376: \eqref{eq:nlssemi}, the solution $u^\eps$ becomes $\eps$-oscillatory
377: for times of order $\O(1)$ as $\eps \to 0$. This phenomenon is typical
378: of supercritical r\'egimes, as far as geometrical optics is concerned
379: (see also \cite{CG05}). This crucial step is stated in
380: Theorem~\ref{theo:reduc}, which in turn is proved thanks to the above
381: mentioned modulated energy functional.  
382: 
383: \smallbreak
384: We end this introduction with a remark concerning the study of the 
385: Cauchy problem for \eqref{eq:nls}. 
386: From \cite{CCT2}, 
387: it is known that the Cauchy problem is not well posed in $H^s(\R^n)$
388: for $0<s<s_{c}$.  
389: Yet, one can try to solve the Cauchy problem by searching the
390: solutions in a larger space. Denote $H^\infty = \cap_{s>0}
391: H^s(\R^n)$. Recall the notion of well-posedness in the sense of Hadamard:
392: \begin{definition}\label{def:WP}
393: Let $s\ge k\ge 0$. The Cauchy problem for \eqref{eq:nls} 
394: is well posed from $H^{s}(\R^n)$ to $H^{k}(\R^n)$ if, for all bounded
395: subset $B\subset H^{s}(\R^n)$, there exist  
396: $T>0$ and a Banach space $X_T\hookrightarrow C([0,T];H^{k}(\R^n))$
397: such that: \\
398: (1) For all $\varphi\in B\cap H^\infty $, \eqref{eq:nls} has a unique
399:   solution $\psi \in
400:   C([0,T];H^\infty)$. \\
401: (2) The mapping $\varphi\in (H^\infty,\| \cdot \|_{B})
402:   \mapsto \psi\in X_T$ is continuous. 
403: \end{definition}
404: The following result is a direct consequence of our analysis (see
405: Remark~\ref{rem:hadamard}).
406: \begin{corollary}\label{cor:CauchyEnd}
407: Let $n\ge 1$ and $\si\ge 1$ be such that $\sigma>2/n$. 
408: The Cauchy problem for \eqref{eq:nls} is not well posed from
409: $H^s(\R^n)$ to $H^k(\R^n)$ for all $(s,k)$ such that 
410: $$
411: 0<s<s_c=\frac{n}{2}-\frac{1}{\sigma}\virgp \qquad k>
412: \frac{s}{1+\si(s_c-s)}\cdot $$
413: \end{corollary}
414: 
415: 
416: 
417: \section{Reduction of the problem}
418: \label{sec:reduc}
419: 
420: Let $0<s<s_c$ and $a_0\in \Sch(\R^n)$. For a sequence $\h$ aimed at
421: going to zero, 
422: consider the family of initial data
423: \begin{equation}\label{eqs:varphi}
424:   \varphi^\h(x) =\h^{s-\frac{n}{2}}a_0\(\frac{x}{\h}\). 
425: \end{equation}
426: Let $\eps = \h^{\si(s_c-s)}$. By assumption, $\eps$ and $\h$ go
427: simultaneously to zero. Define the function $u^\eps$ by the relation:
428: \begin{equation}\label{eqs:ueps}
429:   u^\eps (t,x) = \h^{\frac{n}{2}-s} \psi^\h \( \h^2 \eps t,\h x\). 
430: \end{equation}
431: Then \eqref{eq:nls} is equivalent to \eqref{eq:nlssemi}.
432: Note that we have the relation:
433: \begin{equation*}
434:   \|\psi^\h(t)\|_{\dot H^m} = \h^{s-m}\left\|u^\eps\(
435:   \frac{t}{\h^2\eps}\)\right\|_{ \dot H^m}.
436: \end{equation*}
437: Our aim is to show that for some $\tau>0$ independent of $\eps$, 
438: \begin{equation}\label{eq:epsosc}
439:  \liminf_{\eps \to 0}\eps^k \left\|u^\eps\(
440:   \tau\)\right\|_{ \dot H^k}>0,\quad \forall k\ge 0. 
441: \end{equation}
442: Back to $\psi$, this will yield $t^\h= \tau \h^2\eps$ and 
443: \begin{equation*}
444:   \|\psi^\h(t^\h)\|_{\dot H^k} \gtrsim \h^{s-k}\eps^{-k}=
445:   \h^{s-k(1+\si(s_c-s))}.
446: \end{equation*}
447: To complete the above reduction, note that in view of
448: Theorem~\ref{theo:new}, we only have to prove \eqref{eq:epsosc} for
449: $k\in ]0,1]$. Indeed, for $k>1$, there exists $C_k>0$ such that 
450: \begin{equation*}
451:   \|f\|_{\dot H^1}\le C_k \|f\|_{L^2}^{1-1/k}\|f\|_{\dot
452:   H^k}^{1/k},\quad \forall f\in H^k(\R^n).
453: \end{equation*}
454: This inequality is straightforward thanks to Fourier analysis. Note
455: also that thanks to the conservation of mass for $u^\eps$, we have:
456: \begin{equation*}
457:   \|u^\eps(t)\|_{\dot H^1}\le C_k \|a_0\|_{L^2}^{1-1/k}\|u^\eps(t)\|_{\dot
458:   H^k}^{1/k}.
459: \end{equation*}
460: 
461: 
462: Up to replacing $a_0$ with $|\log \h|^{-1}a_0$ (for instance), the
463: analysis of this section shows that Theorem~\ref{theo:new} follows
464: from:
465: \begin{theorem}\label{theo:reduc}
466: Let $n\ge 1$, $a_0\in \Sch(\R^n)$ be non-trivial, and $\si\ge 1$.
467: There exists a solution
468: $u^\eps$ to \eqref{eq:nlssemi} and $\tau>0$ such that for all $k\in ]0,1]$,
469: \begin{equation*}
470:   \liminf_{\eps \to 0} \left\lVert \lvert\eps D_{x}\rvert^k u^\eps\(
471:   \tau\)\right\rVert_{L^2}>0,\quad \text{where }D_x =-i\nabla.
472: \end{equation*}
473: \end{theorem}
474: \begin{remark}\label{rem:hadamard}
475: As we will see, the previous conclusion holds for all family of smooth 
476: solutions $u^\eps$ defined on a time interval independent of
477: $\eps$. In particular,  
478: Corollary~\ref{cor:CauchyEnd} also 
479: follows from this analysis. To see this, suppose, by
480: contradiction, that the Cauchy problem is well  
481: posed from $H^{s}(\R^n)$ to $H^{k}(\R^n)$. 
482: Since the family of initial data given by \eqref{eqs:varphi} 
483: is bounded in $H^s(\R^n)$, the first point in Definition~\ref{def:WP}
484: implies that  
485: the solutions $\psi^\h$ are defined for a time interval $[0,T]$
486: independent of~$h$. As a result, the function  
487: $u^\eps$, as given by \eqref{eqs:ueps}, is defined for $t\in
488: [0,T/(\eps h^2)]$ with value in $H^\infty(\R^n)$, and hence on the
489: fixed  time interval $[0,T]$. Then, Theorem~\ref{theo:og} implies that
490: there exists $\tau>0$ such that  
491: $\liminf \| |\eps D_{x}|^{k}u^\eps(\tau)|\|_{L^2}>0$. Back to
492: $\psi^\h$ this yields the existence of a sequence $\tau^h$ such that  
493: $\| \psi^h (\tau^h)\|_{H^k}$ tends to $+\infty$, which contradicts the
494: continuity given by the second point of  
495: the definition.
496: \end{remark}
497: 
498: \begin{remark}\label{rema:Iunif}
499:   If we could prove Theorem~\ref{theo:reduc} for $k=1$ only, then back
500:   to $\psi^h$, this would yield Theorem~\ref{theo:new} for $I(s)<k\le s$,
501:   where $I(s)$ is given by \eqref{eq:Ileb}, like in
502:   \cite{Lebeau05}. 
503: \end{remark}
504: Consider the case $k=1$, and recall that the conservation of energy for
505: $u^\eps$ reads, as long as $u^\eps$ is a strong solution of
506: \eqref{eq:nlssemi}:  
507: \begin{equation*}
508:   E^\eps(t) = \frac{1}{2}\int_{\R^n}|\eps\nabla u^\eps(t,x)|^2dx
509: +\frac{1}{\si+1}\int_{\R^n}|u^\eps(t,x)|^{2\si+2}dx\equiv E^\eps(0).
510: \end{equation*}
511: At time $t=0$, the first term (kinetic energy) is of order
512: $\O(\eps^2)$, while the second (potential energy) is dominating, of
513: order $\O(1)$. Therefore, the game consists in showing that there
514: exists $\tau>0$, time at which the kinetic energy is of the order of
515: the total (initial) energy as $\eps\to 0$. 
516: \smallbreak
517: 
518: Some important features of the proof of this result 
519: can be revealed by analyzing the linear case with variable
520: coefficients: 
521: \begin{equation*}
522:   i\eps \d_t u^\eps +\frac{\eps^2}{2}\Delta u^\eps =
523:   V(x) u^\eps\quad ;\quad u^\eps(0,x) = a_0(x). 
524: \end{equation*}
525: Introduce the operator $H^\eps\defn -(\eps^2 / 2)\Delta + V(x)$, 
526: so that $u^\eps (t)=e^{-i t H^\eps/\eps}a_0$. 
527: Now, let $\Op_\eps (q)$ be a semiclassical pseudo-differential operator 
528: with symbol $q(x,\xi)\in S^1_{1,0}$. 
529: Since $e^{i t H^\eps/\eps}$ is unitary, by means of Egorov's Theorem
530: (see \cite{Martinez}), we obtain 
531: \begin{align*}
532: \| \Op_\eps (q)u^\eps\|_{L^2}
533: =\| e^{i t H^\eps/\eps}\Op_\eps (q)e^{-i t H^\eps/\eps} a_0 \|_{L^2}=
534: \|\Op_\eps (q\circ \Phi_t)a_{0}\|_{L^2}+\mathcal{O}(\eps),
535: \end{align*}
536: where $\Phi_{t}$ is the Hamiltonian flow associated with $H^\eps$. 
537: For small times, one can
538: relate $\Phi_t$ to the solution $\phi(t,x)$ of the Hamilton--Jacobi equation:
539: \begin{equation*}
540: \partial_t \phi +\frac{1}{2}|\nabla\phi|^2+V(x)=0\quad;\quad \phi(0,x)=0,
541: \end{equation*}
542: by the identity $\Phi_t(x,\xi)=(X(t,x)+t\xi\, , \,
543: \xi+(\nabla\phi)(t,X(t,x))),$ 
544: where $X$ satisfies 
545: \begin{equation*}
546:   \partial_{t}X(t,x)=(\nabla\phi)(t,X(t,x))\quad ;\quad X(0,x)=x.
547: \end{equation*}
548: Hence, with $q(x,\xi)=i\xi$, we infer
549: $$
550: \| \eps\nabla u^\eps(t)\|_{L^2} = \| (\nabla\phi)(t,X(t,x))
551: a_0\|_{L^2}+\O(\eps), 
552: $$
553: so that the kinetic energy is of order $\O(1)$ provided that
554: $(\nabla\phi)(t,X(t,\cdot)) a_0\neq 0$.   
555: 
556: \smallbreak
557: 
558: The previous argument can be made explicit for  
559: the harmonic oscillator 
560: \begin{equation}
561:   \label{eq:nlslinear}
562:   i\eps \d_t u^\eps_\ell +\frac{\eps^2}{2}\Delta u^\eps_\ell =
563:  \frac{|x|^2}{2} u^\eps_\ell\quad ;\quad u^\eps_\ell(0,x) = a_0(x). 
564: \end{equation}
565: \begin{lemma}\label{lem:linear}
566: Let $n\ge 1$, and $a_0\in \Sch(\R^n)$ (non-trivial). There exists
567: $\tau>0$ such that 
568: the solution $u^\eps_\ell$ to 
569: \eqref{eq:nlslinear} satisfies
570: \begin{equation*}
571:   \liminf_{\eps \to 0} \left\lVert \eps \nabla u^\eps_\ell\(
572:   \tau\)\right\rVert_{L^2}>0.
573: \end{equation*}
574: \end{lemma}
575: \begin{proof}
576: The standard WKB approach yields, at leading order, the following
577: approximate solution:
578: \begin{equation*}
579:   v^\eps_\ell(t,x) = a_\ell(t,x)e^{i\phi_\ell(t,x)/\eps},
580: \end{equation*}
581: where $\phi_\ell$ and $a_\ell$ are given by an eikonal equation and a
582: transport equation. 
583: Since we consider an harmonic
584: oscillator, we can compute $\phi_\ell$ and $a_\ell$ explicitly:
585: \begin{align*}
586:   &\d_t \phi_\ell +\frac{1}{2}|\nabla \phi_\ell|^2
587:   +\frac{|x|^2}{2}=0 ;\  \phi_{\ell \mid t=0}=0:
588: \quad \phi_\ell(t,x)
589:   = \frac{-|x|^2}{2}\tan t.\\ 
590:   &\d_t a_\ell +\nabla \phi_\ell\cdot \nabla a_\ell +\frac{1}{2}a_\ell
591:   \Delta \phi_\ell=0  ;\  a_{\ell \mid t=0}=a_0:
592:  \ a_\ell(t,x)=
593:   \frac{1}{(\cos 
594:   t)^{n/2}}a_0\(\frac{x}{\cos t}\). 
595: \end{align*}
596: Energy estimates then yield (see for instance \cite[\S 3]{CaBKW} for more
597: details): 
598: \begin{equation*}
599:   \|\eps\nabla u^\eps_\ell -\eps\nabla
600:   v^\eps_\ell\|_{L^\infty([0,T];L^2)} \le C_T \eps,\quad \forall T\in
601:   \left[ 0,\frac{\pi}{2}\right[. 
602: \end{equation*}
603: Since 
604: \begin{equation*}
605:  \liminf_{\eps \to 0} \left\lVert \eps \nabla v^\eps_\ell\(
606:   t\)\right\rVert_{L^2}= \sin t \left\lVert x a_0
607:   \right\rVert_{L^2}, \quad \forall t\in
608:   \left[ 0,\frac{\pi}{2}\right[, 
609: \end{equation*}
610: the lemma follows easily. 
611: \end{proof}
612: The strategy for proving Theorem~\ref{theo:reduc} is the same: we
613: compare with the limit system. For nonlinear  
614: Schr\"odinger equation, the eikonal equation which
615: gives the phase is coupled  
616: to the transport equation: the limiting system reads
617: \begin{equation}
618:   \label{eq:limiteS}
619:   \left\{
620:     \begin{aligned}
621:      \d_t \phi +\frac{1}{2}|\nabla \phi|^2+ |a|^{2\si}=0
622: \quad &;\quad \phi_{\mid t =0}=0,\\
623: \d_t a +\nabla\phi\cdot \nabla a +\frac{1}{2}a
624: \Delta \phi = 0\quad   &;\quad a_{\mid
625:   t=0}=a_0. 
626:  \end{aligned}
627: \right.
628: \end{equation}
629: By introducing $v=\nabla\phi$, one can transform this system into a
630: quasilinear system of nonlinear equations.  
631: An important feature of the system thus obtained is that it does not 
632: enter into the classical framework of symmetric hyperbolic systems
633: for $\si\ge 2$.  
634: Nevertheless, one can solve the Cauchy problem~\eqref{eq:limiteS} for
635: all $\si\ge 1$  
636: by a nonlinear change of variable. This is done in
637: \S\ref{sec:muk} following  
638: an idea due to T.~Makino, S.~Ukai and S.~Kawashima~\cite{MUK86}. 
639: \smallbreak
640: 
641: 
642: For the general case $\si\ge 1$, we establish a modulated energy
643: estimate, following the 
644: pioneering  
645: work of Y.~Brenier \cite{BrenierCPDE}. The idea consists in obtaining
646: an estimate  
647: for the $L^{2}$ norm of $\eps\nabla a^\eps$ where $a^\eps$ is the
648: modulated unknown function $a^\eps \defn u^\eps e^{-i\phi/\eps}$.  
649: It is found that 
650: $a^\eps$ satisfies
651: \begin{align*}
652: i\eps \Bigl( \partial_t a^\eps +\nabla\phi\cdot\nabla a^\eps
653: +\frac{1}{2}a^\eps \Delta\phi\Bigr) 
654: + \frac{\eps^2}{2}\Delta a^\eps
655: = \( \left| a^\eps \right|^{2\si} - \left|a\right|^{2\si}\) a^\eps .
656: \end{align*}
657: For $\si=1$, one can obtain estimates uniform in $\eps$, that is 
658: $$
659: \left\| \eps\nabla a^\eps\right\|_{L^\infty([0,T];L^2)} 
660: + \left\| |a^\eps|^2-|a|^2 \right\|_{L^\infty([0,T];L^2)}
661: =\O(\eps),
662: $$
663: by an integration by parts argument. Again, this is based on the
664: hyperbolicity in the case $\si=1$ (see \cite{LinZhang} for an
665: application  
666: of this idea to the {G}ross--{P}itaevskii equations). 
667: Using a modulated energy functional adapted to our problem, we prove
668: the estimate (see
669: Theorem~\ref{theo:og} below): 
670: $$
671: \left\| \eps\nabla a^\eps\right\|_{L^\infty([0,T];L^2)} 
672: + \left\| \(|a^\eps|^2-|a|^2\)\(|a^\eps|^{\si-1}
673:   + |a|^{\si -1}\)  \right\|_{L^\infty([0,T];L^2)}
674: =\O(\eps).
675: $$
676: This is enough to prove Theorem~\ref{theo:reduc} for $k=1$. Note that
677: this suffices to infer
678: Corollary~\ref{cor:energy}.  Finally,
679: to cover the range $k\in ]0,1]$, we microlocalize the previous
680: estimate  by means of wave packets operator. 
681: 
682: 
683: 
684: \section{The limiting system}
685: \label{sec:muk}
686: 
687: Being optimistic, one would try to mimic the approach of E.~Grenier
688: \cite{Grenier98}, and write the solution $u^\eps$ to
689: \eqref{eq:nlssemi} as 
690: $  u^\eps = a^\eps e^{i\phi^\eps /\eps}$, where
691: \begin{equation}
692:   \label{eq:ideal}
693:   \left\{
694:     \begin{aligned}
695:      \d_t \phi^\eps +\frac{1}{2}|\nabla \phi^\eps|^2+ |a^\eps|^{2\si}=0
696: \quad &;\quad \phi^\eps_{\mid t =0}=0,\\
697: \d_t a^\eps +\nabla\phi^\eps\cdot \nabla a^\eps +\frac{1}{2}a^\eps
698: \Delta \phi^\eps = i\frac{\eps}{2}\Delta a^\eps\quad   &;\quad a^\eps_{\mid
699:   t=0}=a_0. 
700:  \end{aligned}
701: \right.
702: \end{equation}
703: Considering the unknown $v^\eps = \nabla \phi^\eps$ instead of
704: $\phi^\eps$, the first step in the analysis would be to solve
705: \begin{equation}
706:   \label{eq:idealv}
707:   \left\{
708:     \begin{aligned}
709:      \d_t v^\eps +v^\eps\cdot \nabla v^\eps+ \nabla\(|a^\eps|^{2\si}\)=0
710: \quad &;\quad v^\eps_{\mid t =0}=0,\\
711: \d_t a^\eps +v^\eps\cdot \nabla a^\eps +\frac{1}{2}a^\eps
712: \DIV v^\eps = i\frac{\eps}{2}\Delta a^\eps\quad   &;\quad a^\eps_{\mid
713:   t=0}=a_0. 
714:  \end{aligned}
715: \right.
716: \end{equation}
717: In \cite{Grenier98}, E.~Grenier considers the unknown $\bu^\eps =
718: (v^\eps,\RE a^\eps,\IM a^\eps)\in \R^{n+2}$. It solves a partial
719: differential equation of the form
720: \begin{equation*}
721:   \d_t \bu^\eps +\sum_{j=1}^n
722:   A_j(\bu^\eps)\d_j \bu^\eps
723:   = \frac{\eps}{2} L  
724:   \bu^\eps .
725: \end{equation*}
726: In the case $\si=1$, the left-hand side of
727: the above equation defines a symmetric quasilinear hyperbolic system in
728: the sense of Friedrichs, with a constant symmetrizer. The linear
729: operator $L$ corresponds to the term $i\Delta$ on the right hand side
730: of \eqref{eq:idealv}: it is skew-symmetric, and does not appear in the
731: energy estimates. Therefore, one can construct a smooth solution to
732: \eqref{eq:idealv} on some time interval $[0,T]$ with $T>0$ independent
733: of $\eps$. In the case $\si\ge 2$,  the symmetrizer of
734: \cite{Grenier98} would become
735: \begin{equation*}
736:   S=\left(
737:     \begin{array}[l]{cc}
738:       \frac{1}{4\si |a^\eps|^{2\si -2}}I_n&0\\
739:  0& I_2
740:     \end{array}
741: \right).
742: \end{equation*}
743: For $a^\eps \in L^2(\R^n)$, this matrix is not uniformly bounded, and
744: this is why the analysis in \cite{Grenier98} is restricted to
745: nonlinearities which are defocusing, and cubic at the origin.
746: \smallbreak
747: 
748: This apparent lack of hyperbolicity is not a real problem for the
749: homogeneous nonlinearity that we consider, provided that we analyze
750: the limiting system only:
751: \begin{equation}
752:   \label{eq:limite}
753:   \left\{
754:     \begin{aligned}
755:      \d_t \phi +\frac{1}{2}|\nabla \phi|^2+ |a|^{2\si}=0
756: \quad &;\quad \phi_{\mid t =0}=0,\\
757: \d_t a +\nabla\phi\cdot \nabla a +\frac{1}{2}a
758: \Delta \phi = 0\quad   &;\quad a_{\mid
759:   t=0}=a_0. 
760:  \end{aligned}
761: \right.
762: \end{equation}
763: The above restriction remains apparently valid for this
764: system: in the presence of vacuum (zeroes of $a$), the symmetrizer $S$
765: is singular. This may lead to a loss of regularity in the energy
766: estimates. However, we shall see that thanks to the special structure of
767: \eqref{eq:limite}, we can construct solutions to \eqref{eq:limite} in
768: Sobolev spaces of sufficiently large order.  
769: Following an idea due to T.~Makino, S.~Ukai and S.~Kawashima
770: \cite{MUK86}, we prove: 
771: \begin{lemma}\label{lem:muk}
772:   Let $a_0\in \Sch(\R^n)$. There exists $T>0$ such that
773:   \eqref{eq:limite} has a unique solution $(\phi,a)\in
774:   C^\infty([0,T]\times \R^n)^2$, with $(\phi,a)\in
775:   C([0,T],H^s)^2$ for all $s\ge 0$. Moreover, $ \<x\>^s\nabla \phi \in
776:   C([0,T],L^2)$ for all $s\ge 0$, where $\<x\>=(1+|x|^2)^{1/2}$.   
777: \end{lemma}
778: \begin{proof} 
779:  Differentiating the first equation in \eqref{eq:limite}, we first
780:  consider:
781: \begin{equation}
782:   \label{eq:limitev}
783:   \left\{
784:     \begin{aligned}
785:      \d_t v +v\cdot \nabla v+ \nabla\(|a|^{2\si}\)=0
786: \quad &;\quad v_{\mid t =0}=0,\\
787: \d_t a +v\cdot \nabla a +\frac{1}{2}a
788: \DIV v = 0\quad   &;\quad a_{\mid
789:   t=0}=a_0. 
790:  \end{aligned}
791: \right.
792: \end{equation}
793: Adapting the idea of \cite{MUK86}, consider the unknown
794: $(v,u)=(v,a^\si)$. Even though the map $a\mapsto a^\si$ is not
795: bijective, this will suffice
796: to prove the lemma. The pair $(v,u)$  solves:
797: \begin{equation}
798:   \label{eq:limitev+}
799:   \left\{
800:     \begin{aligned}
801:      \d_t v +v\cdot \nabla v+ \nabla\(|u|^{2}\)=0
802: \quad &;\quad v_{\mid t =0}=0,\\
803: \d_t u +v\cdot \nabla u +\frac{\si}{2}u
804: \DIV v = 0\quad   &;\quad u_{\mid
805:   t=0}=a_0^\si\in \Sch(\R^n). 
806:  \end{aligned}
807: \right.
808: \end{equation}
809: This system is hyperbolic symmetric, with a constant symmetrizer. To
810: see this, denote $\bu=(v,\RE u,\IM u)^T\in
811: \R^{n+2}$. Equation~\eqref{eq:limitev+} is of the form  
812: \begin{equation*}
813: \d_t \bu +\sum_{j=1}^n A_j(\bu)\d_j \bu =0,
814: \end{equation*}
815: where the matrices $A_j \in {\mathcal M}_{n+2}(\R)$ are such that
816: $SA_j$ are symmetric, for 
817: \begin{equation*}
818: S=\(
819: \begin{array}{cc}
820:  \si {\rm I}_n   &  0  \\
821:    0  &  4     {\rm I}_2
822: \end{array}
823: \)\in {\mathcal S}_{n+2}(\R).
824: \end{equation*}
825: From classical theory on hyperbolic symmetric quasilinear systems (see
826:   e.g. \cite{AlinhacGerardUS,Taylor3}), 
827: there exist  $T>0$ and a unique solution $(v,u)\in C^\infty([0,T]\times
828: \R^n)^2$, which is in $C([0,T],H^s)^2$ for all $s\ge 0$.  
829:   The fact that $\<x\>^s v\in
830:   C([0,T],L^2)$ follows easily by considering the momenta of $u$ and
831:   $v$. Now that $v$ is known,
832:   we define 
833:   $a$ as the solution of the transport equation
834:   \begin{equation*}
835:    \d_t a +v\cdot \nabla a +\frac{1}{2}a
836: \DIV v = 0\quad   ;\quad a_{\mid
837:   t=0}=a_0.  
838:   \end{equation*}
839: The function $a$ has the regularity announced in
840: Lemma~\ref{lem:muk}. We check that $a^\si$ solves the second equation
841: in \eqref{eq:limitev+}. Since $v$ is a smooth coefficient, by
842: uniqueness for this linear equation, we have
843: $u=a^\si$. Therefore, $(v,a)$ solves \eqref{eq:limitev}. To conclude, we
844: notice that $v$ is irrotational, so there exists $\widetilde \phi$
845: such that $v=\nabla \widetilde \phi$. Setting $\phi=\widetilde \phi +
846: F$, where $F=F(t)$ is a function of time only, $(\phi,a)$ solves
847: \eqref{eq:limite}. Uniqueness follows from the uniqueness for
848: \eqref{eq:limitev+}. 
849: \end{proof}
850: \begin{remark}
851:   The proof shows that if we assume only $a_0\in H^s(\R^n)$ with
852:   $s>n/2+1$, then $u,v \in C([0,T];H^s)$. We infer $a \in
853:   C([0,T];H^{s-1})$: the possible loss of regularity due to the lack
854:   of hyperbolicity for \eqref{eq:limite} remains limited.  
855: \end{remark}
856: \begin{remark}
857: The nonlinear change of unknown function, $u=a^\si$, suggests that the
858: above approach cannot be adapted to study \eqref{eq:idealv}, since we
859: have to deal with the term $i\Delta a^\eps$, and prevent the loss of
860: regularity that it may cause in the energy estimates. 
861: \end{remark}
862: 
863: 
864: 
865: 
866: 
867: 
868: 
869: 
870: 
871: 
872: 
873: 
874: 
875: 
876: \section{Semi-classical limit}
877: \label{sec:og}
878: Introduce the hydrodynamic variables:
879:   \begin{equation*}
880:     \rho = |a|^2\quad ;\quad \rho^\eps = |u^\eps|^2 \quad ;\quad
881:     J^\eps = \IM\(\eps \overline u^\eps \nabla u^\eps\). 
882:   \end{equation*}
883: The main result of this section is:
884: \begin{theorem}\label{theo:og}
885:   Let $n\ge1$, and $\si\ge 1$ be an integer. Let $(v,a)\in
886:   C([0,T];H^\infty)^2$ given by Lemma~\ref{lem:muk}, where $v=\nabla
887:   \phi$. Then we have the following estimate:
888:   \begin{align*}
889:     \left\| (\eps\nabla -iv)u^\eps\right\|_{L^\infty([0,T];L^2)}^2 
890: + \left\| \(\rho^\eps-\rho\)^2\((\rho^\eps)^{\si-1}+\rho^{\si-1} \)
891:     \right\|_{L^\infty([0,T];L^1)} 
892: =\O(\eps^2).  
893:   \end{align*}
894: \end{theorem}
895: Note that the above quantities are well-defined for weak
896: solutions. We outline the argument in a formal proof, which is then made
897: rigorous. 
898: 
899: \begin{proof}[Formal proof]
900: For $y\ge 0$, denote 
901: \begin{align*}
902:   &f(y)=y^\si\quad ;\quad F(y)=\int_0^y f(z)dz = \frac{1}{\si +1}y^{\si
903:   +1}\quad ;\\
904: & G(y)= \int_0^y zf'(z)dz= yf(y)-F(y) =
905:   \frac{\si}{\si+1}y^{\si +1}.  
906: \end{align*}
907: We check that
908: $(\rho^\eps,J^\eps)$ satisfies, for $\si\ge 1$:
909: \begin{equation}
910:   \label{eq:hydroq}
911:   \left\{
912:     \begin{aligned}
913:       &\d_t \rho^\eps + \DIV J^\eps =0.\\
914: &\d_t J^\eps_j+\frac{\eps^2}{4}\sum_k \d_k\(4\RE \d_j \overline u^\eps
915: \d_k u^\eps - \d_{jk}^2 \rho^\eps\) +\d_j G 
916: (\rho^\eps) =0. 
917:     \end{aligned}
918: \right.
919: \end{equation}
920: As suggested in \cite[Remark~1, (2)]{LinZhang}, introduce the
921: modulated energy functional:
922: \begin{equation*}
923:   H^\eps(t) = \frac{1}{2}\int_{\R^n} \left| (\eps \nabla
924:   -iv)u^\eps\right|^2dx + \int_{\R^n} \(F(\rho^\eps) -
925:   F(\rho)-(\rho^\eps -\rho)f(\rho)\)dx. 
926: \end{equation*}
927: Denote 
928: \begin{equation*}
929:   K^\eps(t)= \frac{1}{2}\int_{\R^n} \left| (\eps \nabla
930:   -iv)u^\eps\right|^2dx . 
931: \end{equation*}
932: Integrations by parts, which are studied in more detail below,  yield:
933: \begin{equation*}
934:   \frac{d }{dt}H^\eps(t) = \O\(K^\eps +\eps^2\) -\int_{\R^n}
935:   \(G(\rho^\eps)-G(\rho) -(\rho^\eps-\rho)G'(\rho)\) \DIV v\,  dx. 
936: \end{equation*}
937: We check that there exists $K>0$ such that 
938: \begin{equation*}
939:   \left\lvert G(\rho^\eps)-G(\rho) -(\rho^\eps-\rho)G'(\rho)\right
940:   \rvert\le K \left\lvert F(\rho^\eps)-F(\rho) -(\rho^\eps-\rho)F'(\rho)\right
941:   \rvert . 
942: \end{equation*}
943: Therefore,
944: \begin{equation*}
945:   H^\eps (t ) \le H^\eps (0 ) + C \int_0^t \(
946:   H^\eps (s )+\eps^2\)ds . 
947: \end{equation*}
948: We infer by Gronwall lemma that $H^\eps (t ) =\O(\eps^2)$
949: so long as it is defined.
950: Finally, Taylor's formula yields, since $F''(y)=f'(y)=\si y^{\si-1}$:
951: \begin{equation*}
952:   F(\rho^\eps)-F(\rho) -(\rho^\eps-\rho)F'(\rho) =
953:   \si \(\rho^\eps-\rho\)^2\int_0^1 (1-\theta)\( \rho
954:   +\theta\(\rho^\eps-\rho\) \)^{\si-1}d\theta.
955: \end{equation*}
956: The estimate, for all $\theta\in [0,1]$,
957: \begin{equation*}
958:  \( \rho
959:   +\theta\(\rho^\eps-\rho\) \)^{\si-1} = \( (1-\theta)\rho
960:   +\theta\rho^\eps \)^{\si-1} \ge (1-\theta)^{\si-1}\rho^{\si-1} +
961:   \theta^{\si-1} \(\rho^\eps\)^{\si-1}
962: \end{equation*}
963: shows that
964: there exists $c>0$ such that 
965: \begin{equation}\label{eq:convexfinal}
966:  H^\eps (t )\ge K^\eps(t) + c\int_{\R^n} (\rho^\eps-\rho)^2 \(
967:   (\rho^\eps)^{\si-1} + \rho^{\si-1}\)dx.
968:  \end{equation}
969: The result of Theorem~\ref{theo:og} follows. 
970: \end{proof}
971: 
972: \begin{proof}[Rigorous proof] 
973:   In general, the above integrations by parts do not make sense for all
974:   $t\in [0,T]$, since we consider weak solutions only. Note however
975:   that for $\si \ge 2$ and $n\le 3$, the analysis in \cite{AC-BKW}
976:   shows that we can work with strong solutions, so the following
977:   analysis is not needed in this case (for $\si=1$ and $n\ge 1$, the
978:   same holds true, from \cite{Grenier98}). Also, if one is just
979:   interested in proving Corollary~\ref{cor:CauchyEnd} by
980:   contradiction, no further justification is needed for integrations
981:   by parts, and one can skip the end of this section. 
982: 
983: We work on a sequence of global strong solutions,
984:   converging to a weak solution. For $(\delta_m)_m$ a sequence of positive
985:   numbers going to zero,  introduce the saturated
986:   nonlinearity, defined for $y\ge 0$: 
987:   \begin{equation*}
988:     f_m(y)= \frac{y^\si}{1+\(\delta_m y\)^{\si}}\cdot 
989:   \end{equation*}
990: Note that $f_m$ is a symbol of degree $0$. 
991: For fixed $m$, we have a global strong solution $u^\eps_m\in
992: C(\R;H^1)$ to:
993: \begin{equation}
994:   \label{eq:nlssemireg}
995:   i\eps \d_t u^\eps_m +\frac{\eps^2}{2}\Delta u^\eps_m =
996:   f_m\(|u^\eps_m|^2\) u^\eps_m\quad ;\quad u^\eps_m(0,x) = a_0(x). 
997: \end{equation}
998: As $m\to \infty$, the sequence $(u^\eps_m)_m$ converges to a weak
999: solution of \eqref{eq:nlssemi} (see \cite{GV85c,Lebeau05}). 
1000: For $y\ge 0$, introduce also
1001: \begin{align*}
1002:   F_m(y)=\int_0^y f_m(z)dz\quad ;\quad
1003:  G_m(y)= \int_0^y zf'_m(z)dz= yf_m(y)-F_m(y) .  
1004: \end{align*}
1005: The mass and energy associated to $u_m^\eps$ are conserved: 
1006: \begin{align*}
1007:   &M_m^\eps(t)=\int |u^\eps_m(t,x)|^2dx \equiv \|a_0\|_{L^2}^2.\\
1008: &E^\eps_m(t) = \frac{1}{2}\|\eps\nabla u^\eps_m(t)\|_{L^2}^2 +
1009: \int_{\R^n} F_m \( |u^\eps_m(t,x) |^2\) dx \equiv E^\eps_m(0). 
1010: \end{align*}
1011: Moreover, the solution is in $H^2(\R^n)$ for all time: $u^\eps_m\in
1012: C(\R;H^2)$. To see this, we use an idea due to T.~Kato
1013: \cite{Kato87,KatoNLS}, and consider $\d_t u^\eps_m$. Energy estimates
1014: show that $\d_t u^\eps_m \in C(\R;L^2)$, since $f_m$ is a symbol of
1015: degree $0$. Using
1016: \eqref{eq:nlssemireg} and the boundedness of $f_m$, we
1017: infer $\Delta u^\eps_m \in C(\R;L^2)$. 
1018: \smallbreak
1019: 
1020: We consider the hydrodynamic variables:
1021:   \begin{equation*}
1022:     \rho^\eps_m = |u^\eps_m|^2 \quad ;\quad
1023:     J^\eps_m = \IM\(\eps \overline u^\eps_m \nabla u^\eps_m\). 
1024:   \end{equation*}
1025: From the above discussion, we have:
1026: \begin{equation}
1027:   \label{eq:reghydrom}
1028:   \rho^\eps_m (t)\in
1029: W^{2,1}(\R^n)\text{ and }J^\eps_m(t)\in 
1030: W^{1,1}(\R^n),\quad \forall t\in \R.
1031: \end{equation}
1032: The analogue of \eqref{eq:hydroq} is:
1033: \begin{equation}
1034:   \label{eq:hydroqm}
1035:   \left\{
1036:     \begin{aligned}
1037:       &\d_t \rho^\eps_m + \DIV J^\eps_m =0.\\
1038: &\d_t (J^\eps_m)_j+\frac{\eps^2}{4}\sum_k \d_k\(4\RE \d_j \overline u^\eps_m
1039: \d_k u^\eps_m - \d_{jk}^2 \rho^\eps_m\) +\d_j G_m
1040: (\rho^\eps_m) =0. 
1041:     \end{aligned}
1042: \right.
1043: \end{equation}
1044: Introduce the modulated energy functional ``adapted to
1045: \eqref{eq:nlssemireg}'':
1046: \begin{equation*}
1047:   H^\eps_m(t) = \frac{1}{2}\int_{\R^n} \left| (\eps \nabla
1048:   -iv)u^\eps_m\right|^2dx + \int_{\R^n} \(F_m(\rho^\eps_m) -
1049:   F_m(\rho)-(\rho^\eps_m -\rho)f_m(\rho)\)dx. 
1050: \end{equation*}
1051: Notice that this functional is not exactly adapted to
1052: \eqref{eq:nlssemireg}, since the limiting quantities (as $\eps \to 0$)
1053: $\rho$ and $v$ are constructed with the nonlinearity $f$ and not the
1054: nonlinearity $f_m$. We also distinguish the kinetic part:
1055: \begin{equation*}
1056:   K^\eps_m(t) = \frac{1}{2}\int_{\R^n} \left| (\eps \nabla
1057:   -iv)u^\eps_m\right|^2dx. 
1058: \end{equation*}
1059: Thanks to the conservation of energy for $u^\eps_m$, we have:
1060: \begin{align*}
1061:   \frac{d}{dt}K^\eps_m =& -\frac{d}{dt} \int F_m(\rho^\eps_m)dx
1062:   +\frac{1}{2}\int |v|^2 \d_t \rho^\eps_m +\int \rho^\eps_m v \cdot
1063:   \d_t v\\
1064: &-\int J^\eps_m \cdot \d_t v - \int v\cdot \d_t J^\eps_m. 
1065: \end{align*}
1066: Using Lemma~\ref{lem:muk}, \eqref{eq:reghydrom} and
1067: \eqref{eq:hydroqm}, (licit) integrations by parts yield:
1068: \begin{align*}
1069:   \frac{d}{dt}K^\eps_m =& -\frac{d}{dt} \int F_m(\rho^\eps_m)dx
1070:   -\frac{1}{2}\int |v|^2 \DIV J^\eps_m - \sum_{j,k}\int \rho^\eps_m v_j v_k
1071:   \d_j v_k \\
1072: &- \int \rho^\eps_m \nabla f(\rho)\cdot v 
1073: + \int (v\cdot \nabla v)\cdot J^\eps_m + \int  \nabla f(\rho)\cdot
1074:   J^\eps_m\\
1075:  -\sum_{j,k} \int& \d_k v_j \RE \( \eps \d_j \overline
1076:   u^\eps_m \eps \d_k u^\eps_m \)
1077:  -\frac{\eps^2}{4}\int \nabla\(\DIV
1078:   v\)\cdot \nabla \rho^\eps_m +\int \rho^\eps_m v\cdot \nabla
1079:   f_m(\rho^\eps_m). 
1080: \end{align*}
1081: Proceeding as in \cite{LinZhang}, we have:
1082: \begin{align*}
1083:   \eps^2\int \nabla \(\DIV v\)\cdot \nabla \rho^\eps_m &=  \eps\int
1084:   \nabla \(\DIV v\)\cdot \(\overline u^\eps_m \eps\nabla u^\eps_m +
1085:   u^\eps_m \eps\nabla\overline u^\eps_m \)\\
1086: &= \eps\int
1087:   \nabla \(\DIV v\)\cdot\( \overline u^\eps_m (\eps\nabla-iv)u^\eps_m
1088:   +u^\eps_m \overline {(\eps\nabla-iv)u^\eps}_m \)\\
1089: & =\O\( K^\eps_m +\eps^2\), 
1090: \end{align*}
1091: where we have used the conservation of mass and Young's
1092: inequality. From now on, we use the convention that the constant
1093: associated to the notation $\O$ is independent of $m$ 
1094: and $\eps$. We have written the time derivative of $K^\eps_m$ as the
1095: sum of nine terms. The first one corresponds to the conservation of
1096: the energy, and will be canceled by the first term of the time
1097: derivative of $H^\eps_m-K^\eps_m$. We have just bounded the eighth
1098: one. We consider next the sum of four of the seven remaining terms:
1099: the second, third, fifth and seventh,
1100: \begin{align*}
1101:   &\int\big( -\frac{1}{2} |v|^2 \DIV J^\eps_m - \sum_{j,k} \rho^\eps_m v_j v_k
1102:   \d_j v_k +  (v\cdot \nabla v)\cdot J^\eps_m 
1103: -\sum_{j,k}  
1104:   \d_k v_j \RE \( \eps \d_j \overline 
1105:   u^\eps_m \eps \d_k u^\eps_m \) \big)\\
1106: &= \sum_{j,k}\int \Big( v_k \d_j v_k
1107:   \(J^\eps_m\)_j - \lvert u_m^\eps\rvert^2 v_jv_k \d_jv_k + v_j
1108:   \d_jv_k  \(J^\eps_m\)_k- \d_j v_k \RE \( \eps \d_k \overline 
1109:   u^\eps_m \eps \d_j u^\eps_m \).
1110: \end{align*}
1111: Factoring out the term $\d_j v_k$, and recalling that 
1112: \begin{equation*}
1113:  J^\eps_m =\IM\(\overline u_m^\eps \eps \nabla u_m^\eps\), 
1114: \end{equation*}
1115: the above sum can be simplified to:
1116: \begin{equation*}
1117:    -\sum_{j,k}\int \d_j v_k \RE\( \eps\d_j u_m^\eps -iv_j
1118:     u_m^\eps\)\(\overline{\eps\d_k u_m^\eps -iv_k
1119:     u_m^\eps}\)= \O\( K^\eps_m\), 
1120: \end{equation*}
1121: from Cauchy--Schwarz inequality, since $\nabla v\in
1122: L^\infty([0,T]\times\R^n)$. 
1123: We are now left with:
1124: \begin{equation*}
1125:    \frac{d}{dt}K^\eps_m =  \O\( K^\eps_m +\eps^2\) -\frac{d}{dt} \int
1126:   F_m(\rho^\eps_m)+  \int \nabla 
1127:   f_m(\rho^\eps_m) \cdot(\rho^\eps_m v) 
1128: - \int \nabla f(\rho)\cdot \(
1129:   \rho^\eps_m v - 
1130:   J^\eps_m\).  
1131:   \end{equation*}
1132: Since $G'_m(y)=yf'_m(y)$, we infer:
1133: \begin{equation*}
1134:    \frac{d}{dt}K^\eps_m = \O\( K^\eps_m +\eps^2\) -\frac{d}{dt} \int
1135:   F_m(\rho^\eps_m)-  \int 
1136:   G_m(\rho^\eps_m) \DIV v 
1137: - \int \nabla f(\rho)\cdot \(
1138:   \rho^\eps_m v - 
1139:   J^\eps_m\).  
1140:   \end{equation*}
1141: Direct computations yield
1142: \begin{align*}
1143:   \frac{d}{dt} \(H^\eps_m-K^\eps_m\) =&\frac{d}{dt} \int
1144:   F_m\(\rho^\eps_m\) -\int \nabla
1145:   f_m(\rho)\cdot J_m^\eps + \int \( \rho^\eps_m
1146:   - \rho\) v\cdot \nabla f_m\(\rho\)\\
1147: & +\int\( \rho^\eps_m
1148:   - \rho\) G'_m \(\rho\)\DIV v. 
1149: \end{align*}
1150: We therefore come up with:
1151: \begin{equation*}
1152:   \begin{aligned}
1153:     \frac{d}{dt}H^\eps_m=& \O\( K^\eps_m +\eps^2\)  -\int \(
1154:     G_m(\rho^\eps_m)-  G_m(\rho)
1155:     -(\rho^\eps_m -\rho)G_m'(\rho)\)\DIV v\\
1156: &+ \int \nabla\( f(\rho)-f_m(\rho)\)\cdot (J^\eps_m - \rho^\eps_m v).  
1157:   \end{aligned}
1158: \end{equation*}
1159: Note that $f(\rho)-f_m(\rho)\to 0$ in $L^\infty([0,T];W^{1,\infty})$
1160: as $m\to \infty$. We can thus write:
1161: \begin{equation}\label{eq:modul19}
1162:   \begin{aligned}
1163:         \frac{d}{dt}H^\eps_m= &\O\( K^\eps_m +\eps^2\)+o_{m\to
1164:  \infty}(1)\\
1165:  & -\int \(
1166:     G_m(\rho^\eps_m)-  G_m(\rho)
1167:     -(\rho^\eps_m -\rho)G_m'(\rho)\)\DIV v.
1168:   \end{aligned}
1169: \end{equation}
1170: We conclude thanks to the following lemma, whose proof is postponed to the
1171:  end of this section:
1172: \begin{lemma}\label{lem:convex}
1173: There exists $K>0$ independent of $m$ such that  $\forall \rho',\rho \ge 0$,
1174: \begin{equation*}
1175: \left\lvert G_m(\rho')-  G_m(\rho)
1176:     -(\rho' -\rho)G_m'(\rho)\right\rvert \le K \left\lvert F_m(\rho')-
1177:     F_m(\rho) -(\rho' -\rho)F_m'(\rho)\right\rvert .
1178: \end{equation*}
1179: \end {lemma}
1180: Using this lemma and \eqref{eq:modul19}, we infer:
1181: \begin{equation*}
1182:  \frac{d}{dt}H^\eps_m\le  C \(H^\eps_m +\eps^2\)+o_{m\to
1183:  \infty}(1),
1184: \end{equation*}
1185: for some $C$ independent of $m$.
1186: Gronwall lemma yields
1187: \begin{equation*}
1188:   \sup_{t\in [0,T]} H^\eps_m(t) \le C'\eps^2 +o_{m\to \infty }(1),
1189: \end{equation*}
1190: for some $C'$ independent of $m$.
1191: Letting $m\to \infty$, Fatou's lemma yields 
1192: \begin{equation*}
1193:   \sup_{t\in [0,T]} H^\eps(t)\le C' \eps^2.
1194: \end{equation*}
1195: Theorem~\ref{theo:og} then follows from \eqref{eq:convexfinal}.  
1196: \end{proof}
1197: 
1198: \begin{proof}[Proof of Lemma~\ref{lem:convex}] Taylor's formula yields
1199: \begin{align*}
1200: G_m(\rho')-  G_m(\rho)
1201:     -(\rho' -\rho)G_m'(\rho) &= \( \rho' -\rho\)^2 \int_0^1
1202:     (1-\theta)G_m''\(\rho + \theta(\rho' -\rho)\)d\theta.\\ 
1203:     F_m(\rho')-  F_m(\rho)
1204:     -(\rho' -\rho)F_m'(\rho) &= \( \rho' -\rho\)^2 \int_0^1
1205:     (1-\theta)F_m''\(\rho + \theta(\rho' -\rho)\)d\theta. 
1206: \end{align*}
1207: By definition,
1208: \begin{equation*}
1209: F''_m(y)= f'_m(y)\quad ; \quad G''_m(y) = f'_m (y)+ y f''_m(y).
1210: \end{equation*}
1211: Set, for $y\ge 0$, $h(y)=y^\si /(1+y^\si)$:
1212: \begin{equation*}
1213:  f'_m(y)=\delta_m^{1-\si} h'(\delta_m y)\quad ; \quad
1214:  f''_m(y)=\delta_m^{2-\si} h''(\delta_m y). 
1215: \end{equation*}
1216: Moreover,
1217: \begin{equation*}
1218: h'(y)= \frac{\si y^{\si-1}}{(1+y^\si)^2}\ge 0.
1219: \end{equation*}
1220: Therefore, to prove the lemma, it suffices to show that for all $y\ge 0$,
1221: \begin{equation}\label{eq:23h25}
1222: \left\lvert y h''(y)\right\rvert \le C  h'(y).
1223: \end{equation}
1224: We check the identity
1225: \begin{equation*}
1226: y h''(y) = h'(y )\times\frac{\si-1 -(\si+1)y^\si}{1+y^\si}.
1227: \end{equation*}
1228: The estimate \eqref{eq:23h25} is then straightforward, and
1229: the lemma follows. 
1230: \end{proof}
1231: 
1232: 
1233: 
1234: 
1235: \section{End of the proof of Theorem~\ref{theo:new}}
1236: \label{sec:concl}
1237: 
1238: To conclude, the heuristic argument is as follows. From
1239: Theorem~\ref{theo:og}, we expect
1240: \begin{equation*}
1241:   \left\| \eps \nabla u^\eps(t)\right\|_{L^2} \thickapprox \left\|
1242:   v(t) u^\eps(t)\right\|_{L^2} \thickapprox \left\|
1243:   v(t) a(t)\right\|_{L^2}.
1244: \end{equation*}
1245: This follows easily from H\"older's inequality. For the values
1246: $k\in]0,1[$ in Theorem~\ref{theo:og}, we morally use an estimate of
1247: the form
1248: \begin{equation*}
1249:   \left\| |v(t)|^k u^\eps(t)\right\|_{L^2}\lesssim 
1250: \left\| |\eps D_x|^k u^\eps(t)\right\|_{L^2}+ 
1251: \left\| |\eps D_x -v(t)|^k u^\eps(t)\right\|_{L^2},
1252: \end{equation*}
1253: where the first term of the right-hand side goes to zero by
1254: interpolation between  $k=0$ and $k=1$. The aim of the
1255: following lemma is to justify such a  statement. 
1256: \begin{lemma}\label{prop:micro}There exists a constant $K$ such that, for all 
1257: $\eps\in ]0,1]$, for all $s\in [0,1]$, 
1258: for all $u\in H^{1}(\R^n)$ and for all $v\in W^{1,\infty}(\R^n)$,
1259: \begin{equation*}
1260: \| |v|^s u\|_{L^2}\le \| |\eps D_{x}|^s u\|_{L^2}+\| (\eps\nabla -i
1261: v)u\|_{L^2}^{s} 
1262: \| u\|_{L^2}^{1-s}+
1263: \eps^{s/2} K \(1+\| \nabla v\|_{L^\infty}\)\| u\|_{L^2}.
1264: \end{equation*}
1265: \end{lemma}
1266: \begin{proof} 
1267: We begin with the following elementary inequality: 
1268: For all $(x,y)\in \R^{n}\times\R^{n}$ and all $s\in [0,1]$, there holds
1269: \begin{equation}\label{elementary}
1270: | x |^s \le | y|^s + |x-y|^s.
1271: \end{equation}
1272: To see this, note that the result is obvious if $|x|\le |y|$. 
1273: Else, write $|y|=\lambda |x|$ with $\lambda\in [0,1]$ and use the
1274: inequalities $\lambda\le \lambda^s$ and $(1-\lambda)\le (1-\lambda)^s$.  
1275: 
1276: With this preliminary established, 
1277: introduce the wave-packets operator (see e.g. \cite{CF,Delort,Martinez})
1278: $$
1279: W^\varepsilon v (x,\xi) = c_n \varepsilon^{-3n/4} 
1280: \int _{\R^n} e^{i(x-y)\cdot\xi/\eps  - (x-y)^2/2\eps} v (y) \, dy,
1281: $$
1282: with $c_n=2^{-n/2} \pi^{-3n/4}$. 
1283: The mapping $v\mapsto W^\eps v$ is continuous from the Schwartz class
1284: $\mathcal{S}(\R^n)$  
1285: to $\mathcal{S}(\R^{2n})$, 
1286: and $W^\eps$ extends as an isometry from 
1287: $L^2(\R^n)$ to $L^2(\R^{2n})$:
1288: $$
1289: \| W^\eps v\|_{L^2(\R^{2n})}=\| v\|_{L^2(\R^{n})}.
1290: $$
1291: By applying \eqref{elementary}, we have
1292: $$
1293: \big\| |v(x)|^s W^\eps u\big\|_{L^2(\R^{2n})}\le \big\| |\xi |^s
1294: W^\eps u \big\|_{L^2(\R^{2n})}  
1295: + \big\| | \xi - v(x)|^s W^\eps u\big\|_{L^2(\R^{2n})}.
1296: $$
1297: Therefore, since
1298: \begin{align*}
1299:  \| | \xi - v(x)|^s W^\eps u\|_{L^2(\R^{2n})} &\le 
1300:  \|  W^\eps u\|_{L^2(\R^{2n})}^{1-s} 
1301:  \big\| | \xi - v(x)| W^\eps u\big\|_{L^2(\R^{2n})}^{s}\\
1302: & \le  \|  u\|_{L^2(\R^{n})}^{1-s} 
1303:  \big\| ( \xi - v(x)) W^\eps u\big\|_{L^2(\R^{2n})}^{s},
1304: \end{align*}
1305: to obtain the desired estimate, we need only prove:
1306: \begin{align}
1307: &\big\| |v(x)|^s W^\eps u- W^\eps(|v|^su) \big\|_{L^2(\R^{2n})} \le
1308: K\eps^{s/2} \| \nabla v\|_{L^\infty}^{s}\| u\|_{L^2},\label{last1}\\ 
1309: &\big\| |\xi|^s W^\eps u - W^\eps(|\eps D_x|^s u)
1310: \big\|_{L^2(\R^{2n})} \le K \eps^{s/2} \| u\|_{L^2},\label{last2}\\ 
1311: & \big\| (i\xi - iv) W^\eps u - W^\eps \bigl(( \eps\nabla -
1312: iv)u\bigr) \big\|_{L^2(\R^{2n})} 
1313: \le K\eps^{1/2} (1+ \| \nabla v\|_{L^\infty})\| u\|_{L^2}.\label{last3}
1314: \end{align}
1315: These properties follows from the fact that the wave packets operator
1316: conjugates the action of  
1317: pseudo-differential operators, approximately, to multiplication by
1318: symbols. For smooth symbols, one has  
1319: sharp results (see \cite{CF,Delort,Martinez}). For the rough symbols
1320: $|v(x)|^s$ and $|\eps\xi|^s$, one can proceed as follows.
1321: 
1322: To prove~\eqref{last1}, directly from the definition, we compute
1323: \begin{align*}
1324: &\left\| |v|^s W ^\varepsilon u - W ^ \varepsilon (|v|^s u)
1325: \right\|^2_{L^2 (\R^{2n})} \\
1326: &\qquad
1327: =c_{n}^2(2\pi)^n \varepsilon ^{-n/2} \iint  e^{-(x-y)^2
1328:   /\varepsilon } \left\lvert |v(x)|^s  - |v(y)|^s\right\rvert ^2 \left| 
1329: u(y) \right| ^2 \, dy dx.
1330: \end{align*}
1331: Consequently, since $v\in W^{1,\infty}(\R^n)$, the inequality
1332: \eqref{elementary} implies
1333: \begin{align*}
1334: &\left\| |v|^s W ^\varepsilon u - W ^ \varepsilon (|v|^s u)
1335: \right\|^2_{L^2 (\R^{2n})} \\
1336: &\qquad\le K\| \nabla v\|_{L^\infty}^{2s}
1337: \iint \varepsilon ^{-n/2} e^{-(x-y)^2 /\varepsilon } \left|
1338:   x-y\right|^{2s} 
1339: \left| u(y)\right|^2 \, dy dx\\
1340: &\qquad\le K \| \nabla v\|_{L^\infty}^{2s} \iint e^{-z^2} \left|
1341:   \sqrt{\eps}z\right|^{2s} 
1342: \left| u(x-\sqrt{\eps}z) \right|^2 \,dz dx,
1343: \end{align*}
1344: which proves~\eqref{last1}.
1345: We next compute 
1346: $W^\eps(|\eps D_x|^s u) (x,\xi)$: it is given by
1347: \begin{align*}
1348: c_n (2\pi)^{-n/2} \eps^{-7n/4}\iint e^{i(x-y)\cdot
1349:   (\xi-\theta)/\eps-(x-y)^2/2\eps}e^{ix\cdot\theta/\eps} 
1350: |\theta|^s\widehat{u}\(\frac{\theta}{\eps}\)\,d\theta dy,
1351: \end{align*}
1352: where $\widehat{u}$ is the Fourier transform of $u$. Hence, by using 
1353: $$ 
1354: (2\pi)^{-n/2}\int e^{i(x-y)\cdot (\xi-\theta)/\eps-(x-y)^2/2\eps}\, dy =
1355: \eps^{n/2}e^{-(\xi-\theta)^2/2\eps}, 
1356: $$ 
1357: we find
1358: $$
1359: W^\eps\bigl( |\eps D_x|^s u\bigr) (x,\xi)\defn e^{i x\cdot\xi /\eps} 
1360: W^\eps w^\eps (\xi,-x),
1361: $$
1362: with $w^\eps (\tau)\defn | \tau|^s \eps^{-n/2}\widehat{u}(\tau/\eps)$. 
1363: This leads us back to the situation of the previous step (with
1364: $|v(x)|^s$ replaced with $|x|^s$),  and hence \eqref{last2} is proved.  
1365: \smallbreak
1366: 
1367: Finally, the arguments establishing \eqref{last1} and
1368: \eqref{last2} also yield the usual estimates 
1369: \begin{align*}
1370: &\big\| v W^\eps u - W^\eps(v u ) \big\|_{L^2(\R^{2n})} \le
1371: K\eps^{1/2}\| \nabla v\|_{L^\infty}\| u\|_{L^2},\\ 
1372: &\big\| i\xi W^\eps u - W^\eps(\eps \nabla u)
1373: \big\|_{L^2(\R^{2n})} \le K \eps^{1/2} \| u\|_{L^2}, 
1374: \end{align*}
1375: which proves \eqref{last3}. This completes the proof of the lemma.
1376: \end{proof}
1377: We infer that the heuristic argument of the beginning of this section
1378: is justified:
1379: \begin{corollary}For all $t\in [0,T]$ and all $k\in]0,1]$, we have:
1380: \begin{equation}\label{claimEP}
1381: \liminf_{\eps\to 0} \left\| |\eps D_x|^k u^\eps (t)\right\|_{L^2} \ge 
1382: \left\| |v(t)|^k a(t) \right\|_{L^2}.
1383: \end{equation}
1384: \end{corollary}
1385: \begin{proof} 
1386: Let $t\in [0,T]$. It follows from the previous lemma that
1387: $$
1388: \bigl\| |\eps D_x|^k u^\eps (t)\bigr\|_{L^2} = \| |v(t)|^k u^\eps(t)
1389: \|_{L^2}+o(1). 
1390: $$
1391: Write
1392: $$
1393: \left\| |v(t)|^{k}a(t)\right\|_{L^2} \le \left\| |v(t)|^{k}
1394:   u^\eps(t)\right\|_{L^2}+ 
1395: \left\|  |v(t)|^{2k} \(|u^\eps(t)|^2- |a(t)|^2\)\right\|_{L^1}. 
1396: $$
1397: From H\"older's inequality, the last term is bounded by
1398: \begin{equation}\label{eq:14h27}
1399: \left\| |v(t)|^{2k}\right\|_{L^{1+1/\si}} \left\| |u^\eps(t)|^2-
1400:   |a(t)|^2\right\|_{L^{\si+1}}.   
1401: \end{equation}
1402: When $k\ge \si/(\si+1)$, Lemma~\ref{lem:muk} and Sobolev embedding show that
1403: the first term is bounded on $[0,T]$. 
1404: When $0<k< \si/(\si+1)$, H\"older's inequality yields:
1405: \begin{equation*}
1406:  \left\| |v(t)|^{2k}\right\|_{L^{1+1/\si}}\le  C_N \left\|
1407:  \<x\>^{N}v(t)\right\|_{L^2}^{2k\si/(\si +1)}\quad \text{for
1408:  }N>\frac{n}{2k}\(\frac{\si}{\si+1}-k\). 
1409: \end{equation*}
1410: Lemma~\ref{lem:muk} and Theorem~\ref{theo:og} show
1411: that \eqref{eq:14h27} goes to zero as $\eps$ tends to $0$. 
1412: \end{proof}
1413: To complete the proof of Theorem~\ref{theo:reduc}, 
1414: it remains only to prove that the right-hand side in~\eqref{claimEP}
1415: is non trivial. To see this, we note that, 
1416: from \eqref{eq:limitev},
1417: \begin{equation*}
1418:  a_{\mid t=0}=a_0\quad ;\quad v_{\mid t=0} =0\quad ;\quad \d_t v_{\mid
1419:  t=0} = -\nabla \( |a_0|^{2\si}\). 
1420: \end{equation*}
1421: Therefore, by continuity (see Lemma~\ref{lem:muk}), we obtain the
1422: following result. 
1423: \begin{lemma}
1424: There exists $\tau>0$ such that
1425: \begin{equation}\label{eq:decolle}
1426:   \int |v(\tau,x)|^{2k} |a(\tau,x)|^2dx>0, \quad \forall k\in [0,1]. 
1427: \end{equation}
1428: \end{lemma}
1429: %\begin{proof}
1430: %From \eqref{eq:limitev}, we know that
1431: %\begin{equation*}
1432:  %a_{\mid t=0}=a_0\quad ;\quad v_{\mid t=0} =0\quad ;\quad \d_t v_{\mid
1433: % t=0} = -\nabla \( |a_0|^{2\si}\). 
1434: %\end{equation*}
1435: %Therefore, by continuity (see Lemma~\ref{lem:muk}), there exists
1436: %$\tau>0$ such that 
1437: %\eqref{eq:decolle} holds. This implies Theorem~\ref{theo:reduc}, hence
1438: %Theorem~\ref{theo:new}. 
1439: %\end{proof}
1440: 
1441: This implies Theorem~\ref{theo:reduc}, hence
1442: Theorem~\ref{theo:new}. 
1443: 
1444: \begin{remark}\label{rem:semiquasi}
1445:   We can compare the results of this paper with the analysis in
1446:   \cite{CCT2}. The approximate solution used in \cite{CCT2} consists
1447:   in neglecting the Laplacian in \eqref{eq:nlssemi}:
1448:   \begin{equation*}
1449:     i\eps\d_t w^\eps = |w^\eps|^{2\si}w^\eps \quad ;\quad w^\eps_{\mid
1450:     t=0}=a_0,\quad \text{hence}\quad w^\eps(t,x)=a_0(x)e^{-it
1451:     |a_0(x)|^{2\si}/\eps}.   
1452:   \end{equation*}
1453: A direct application of Gronwall lemma shows that $w^\eps$
1454:   is  a suitable 
1455:   approximation of $u^\eps$ up to time of order $c\eps |\log
1456:   \eps|^\theta$,  for some $c,\theta>0$. The Taylor expansion in time for $v$
1457:   shows that
1458:   \begin{equation*}
1459:     v(t,x) = -t \nabla \(|a_0(x)|^{2\si}\) +\O\(t^3\).
1460:   \end{equation*}
1461: The formal analysis of \cite[\S 3.1]{CaARMA} is thus
1462: justified also in this case: $w^\eps(t)$ is a good approximation of
1463: $u^\eps(t)$ for $t\ll \eps^{1/3}$:
1464:   \begin{equation*}
1465:     \||\eps D_x|^s u^\eps(t)\|_{L^2}\thickapprox \||v(t)|^s
1466:     a(t)\|_{L^2}\thickapprox  \||\eps D_x|^s w^\eps(t)\|_{L^2}\quad
1467:     \text{for }t\ll \eps^{1/3}. 
1468:   \end{equation*}
1469:  To prove this point, it seems
1470:   necessary to perform a quasilinear analysis (see~\S\ref{sec:muk}),
1471:   and the semilinear approach based on Gronwall lemma is not enough.  
1472: \end{remark}
1473: \section{Final remarks}\label{sec:geometrie}
1474: 
1475: To conclude this paper, we note that the approach presented here
1476: remains efficient in  the case of non-trivial geometries. Indeed, the
1477: scaling argument that we have used in \S\ref{sec:reduc} is merely
1478: helpful for the intuition, to guess a suitable approximate
1479: solution. This argument meets the strategy adopted in the appendix of
1480: \cite{BGTENS}. Introduce $\om^h$ and $A^h$ given by
1481: \begin{equation*}
1482:   v(t,x) = \om^h\(h^2\eps t,hx\)\quad ;\quad a(t,x) =
1483:   h^{\frac{n}{2}-s}A^h\(h^2\eps t,hx\). 
1484: \end{equation*}
1485: The key approximation that we have used,
1486: \begin{equation*}
1487:   \left\lvert \eps\nabla u^\eps(t,x)\right\rvert^2 \approx \left\lvert
1488:   v(t,x)a(t,x) \right\rvert^2,\quad 0\le t\le T,
1489: \end{equation*}
1490: then reads
1491: \begin{equation*}
1492:   \left\lvert \eps h \nabla \psi^h (t,x)\right\rvert^2 \approx \left\lvert
1493:   \om^h(t,x)A^h(t,x) \right\rvert^2, \quad 0\le t\le h^2 \eps T.
1494: \end{equation*}
1495: More precisely, in terms of the initial problem \eqref{eq:nls},
1496: Theorem~\ref{theo:og} reads:
1497: \begin{equation*}
1498:   \begin{aligned}
1499:    & h^{-s} \left\lVert \(\eps h \nabla
1500:     -i\om^h\)\psi^h\right\rVert_{L^\infty([0,h^2\eps T];L^2)}^2 \\
1501: +&h^{\(\frac{n}{2}-s\)2(\si+1) -\frac{n}{2}}\left\lVert \(\lvert
1502:     \psi^h\rvert^2 - \lvert A^h\rvert^2\)^2\(\lvert
1503:     \psi^h\rvert^{2\si-2} +\lvert
1504:     A^h\rvert^{2\si-2}\)\right\rVert_{L^\infty([0,h^2\eps
1505:     T];L^1)}\\
1506: &=\O\(\eps^2\). 
1507:   \end{aligned}
1508: \end{equation*}
1509: It is essentially this estimate that we have used to prove
1510: Theorem~\ref{theo:new} (and Corollary~\ref{cor:energy} stems exactly
1511: from this estimate). 
1512: \smallbreak
1513: 
1514: Suppose for instance that $x$ belongs to a bounded
1515: domain $M$, and not to all of $\R^n$, and that
1516: we consider 
1517: \eqref{eq:nls}  on $M$, with Dirichlet or Neumann boundary
1518: condition. If $a_0$ is compactly supported in a ball, 
1519: contained in the interior
1520: of $M$, 
1521: \begin{equation*}
1522:   \operatorname{supp} a_0\subset B \Subset M,
1523: \end{equation*}
1524: then we can still
1525: consider \eqref{eq:limite}, viewed as a 
1526: system on $\R^n$. The key remark is that for $a_0$ compactly
1527: supported, the smooth solutions to \eqref{eq:limite} have a finite
1528: speed of propagation, which is \emph{zero}. This is an important step
1529: in proving that smooth solutions develop singularities in finite time;
1530: see \cite{MUK86,Xin98}. Therefore, $(\phi,a)$  remains supported in
1531: $B$ for $t\in [0,T]$; up to changing the origin to the center of $B$,
1532: so does $(\om^h,A^h)$ on the time interval $[0,h^2\eps T]$. In
1533: particular, $(\om^h,A^h)$ is supported away from the boundary of $M$
1534: for $t\in [0,h^2\eps T]$.
1535: \smallbreak
1536: 
1537: In the proof of Theorem~\ref{theo:og}, 
1538: the integrations by parts affect $v$ or $a$, that is, $\om^h$ or
1539: $A^h$. Only two terms 
1540: in the differentiation of the modulated energy functional do not
1541: contain $\om^h$ or $A^h$: these two terms correspond to the global
1542: energy of $\psi^h$, which is a non-increasing function of time for
1543: strong solutions in the case of $\R^n$ (since it is in fact
1544: constant). As a matter of fact, this property is needed for 
1545: strong solutions only, since it remains for weak solutions, by Fatou's
1546: lemma. Therefore, Theorem~\ref{theo:new} remains valid
1547: on $M$, provided that we can construct strong solutions with a
1548: non-increasing energy. This is the case of compact surfaces  when
1549: $\si\ge 1$, and of compact three dimensional manifolds 
1550: when $\si=1$, see \cite{BGT}. This is also the case of bounded domains
1551: in $\R^2$ for $\si\ge 1$, see \cite{RamonaBSMF}, of the ball in $\R^3$
1552: for $\si= 1$ and radial data \cite{RamonaCPDE}, and of exterior
1553: domains in $\R^3$, 
1554: for $\si=1$ \cite{RamonaJMPA}. 
1555: \smallbreak
1556: 
1557: Note however that the notion of criticality
1558: may differ on a curved space (see e.g. \cite{BGTMRL,ThomannSurf}): the
1559: curvature of a manifold may create more ill-posedness phenomena, but
1560: since in the proof of ill-posedness in \cite{CCT2} (see also
1561: \cite{CaARMA}), the Laplacian is 
1562: neglected, the critical Sobolev exponent for local well-posedness cannot
1563: be less than in the case of $\R^n$. 
1564: \smallbreak
1565: 
1566: This approach suggests that we can consider a
1567: more general manifold, up to working on a local chart, and provided
1568: that the energy associated to strong solutions of \eqref{eq:nls} is
1569: a non-increasing function of time, a question which we leave out at
1570: this stage. 
1571: \smallbreak
1572: 
1573: Finally, we go back to the whole space case, $x\in \R^n$. As recalled
1574: above, if $a_0$ is compactly  supported, then the ansatz that we
1575: consider remains supported in the same compact so long as the solution
1576: to \eqref{eq:limiteS} remains smooth; see \cite{MUK86}, and also
1577: \cite{Xin98}. The justification of WKB analysis for short time shows
1578: that at least when $\si=1$ (\cite{Grenier98}), or $\si \in \N$ and $n\le 3$
1579: (\cite{AC-BKW}), we have, thanks to Borel lemma,
1580: \begin{equation*}
1581:   u^\eps(t,x) = u^\eps_{\rm app}(t,x) + \O\(\eps^\infty\), \quad
1582:   \text{in }C\([0,T];L^2\cap L^\infty\), 
1583: \end{equation*}
1584: where $u^\eps_{\rm app}$ is supported in the same compact as
1585: $a_0$. This seems to be an encouraging remark, in view of considering
1586: \emph{fixed} initial data as in \cite{Lebeau05}, instead of a sequence
1587: of initial data like here. However, the information that we do not
1588: have for the Schr\"odinger equation, and which is available for the
1589: wave equation, is a notion of finite speed of propagation for
1590: \emph{weak} solutions to the nonlinear equation. This seems to be the
1591: only obstacle to consider fixed initial data in the Schr\"odinger
1592: case.
1593: 
1594: \subsection*{Acknowledgments} The authors are grateful to
1595: Patrick G\'erard for stimulating comments on this work. 
1596: 
1597: \bibliographystyle{amsplain}
1598: \bibliography{../MathAnnalen/loss}
1599: 
1600: 
1601: 
1602: \end{document}
1603: 
1604: 
1605: 
1606: 
1607: 
1608: 
1609: 
1610: 
1611: 
1612: 
1613: 
1614: 
1615: 
1616: 
1617: 
1618: 
1619: 
1620: 
1621: