1: \documentclass[11pt,a4paper,leqno]{amsart}
2:
3: \usepackage{amsmath,amsfonts,amssymb}
4: \usepackage{enumerate}
5: \usepackage{amsthm}
6: %\usepackage{draftcopy}
7: %\usepackage{showkeys}
8: \usepackage{hyperref}
9:
10:
11: \def\dis
12: {\displaystyle}
13:
14: \def\ii{\int\!\!\!\int}
15:
16: \def\C{{\mathbb C}}
17: \def\R{{\mathbb R}}
18: \def\N{{\mathbb N}}
19: \def\Z{{\mathbb Z}}
20: \def\T{{\mathbb T}}
21: \def\Q{{\mathbb Q}}
22: \def\S{{\mathcal S}}
23: \def\virgp{\raise 2pt\hbox{,}}
24: \def\bv{{\bf v}}
25: \def\bu{{\bf u}}
26: \def\bw{{\bf w}}
27: \def\({\left(}
28: \def\){\right)}
29: \def\<{\left\langle}
30: \def\>{\right\rangle}
31:
32:
33:
34: \def\longformule#1#2{
35: \displaylines{
36: \qquad{#1}
37: \hfill\cr
38: \hfill {#2}
39: \qquad\cr
40: }
41: }
42:
43:
44: \def\Eq#1#2{\mathop{\sim}\limits_{#1\rightarrow#2}}
45: \def\Tend#1#2{\mathop{\longrightarrow}\limits_{#1\rightarrow#2}}
46:
47: \def\d{{\partial}}
48: \def\a{{\tt a}}
49: \def\b{\beta}
50: \def\g{\gamma}
51: \def\e{\varepsilon}
52: \def\l{\lambda}
53: \def\om{\omega}
54: \def\si{{\sigma}}
55: \def\U{{\mathcal U}^\e}
56: \def\V{{\tt V}^\e}
57: \def\W{{\tt W}}
58: \def\u{{\tt u}}
59: \def\v{{\tt v}}
60: \def\w{{\tt w}}
61: \def\H{{\tt H}}
62: \def\F{\mathcal F}
63: \def\M{{\mathcal M}^\e}
64: \def\D{{\tt D}^\e}
65: \def\O{\mathcal O}
66: \def\eik{\phi_{\rm eik}}
67:
68:
69: \newenvironment{idee}{\par
70: \noindent{\bf Develop: }\normalfont\itshape}
71:
72: \theoremstyle{plain}
73: \newtheorem{theo}{Theorem}[section]
74: \newtheorem{lem}[theo]{Lemma}
75: \newtheorem{cor}[theo]{Corollary}
76: \newtheorem{prop}[theo]{Proposition}
77: \newtheorem{hyp}{Assumption}
78:
79: \theoremstyle{definition}
80: \newtheorem{defin}[theo]{Definition}
81: \newtheorem*{nota}{Notation}
82:
83: \theoremstyle{remark}
84: \newtheorem{rema}[theo]{Remark}
85: \newtheorem*{rema*}{Remark}
86: \newtheorem*{remas}{Remarks}
87: \newtheorem{ex}{Example}
88:
89: \numberwithin{equation}{section}
90:
91:
92:
93:
94:
95: \begin{document}
96: \author[R. Carles]{R{\'e}mi Carles}
97: \address{MAB, Universit\'e de Bordeaux 1\\ 351 cours de la
98: Lib{\'e}ration\\ 33405 Talence cedex\\ France}
99: \email{Remi.Carles@math.cnrs.fr}
100: %\date{\today}
101: \thanks{Supports by European network HYKE,
102: funded by the EC as contract HPRN-CT-2002-00282, by Centro de
103: Matem\'atica e Aplica\c c\~oes Fundamentais (Lisbon), funded by
104: FCT as contract POCTI-ISFL-1-209, and by the ANR project SCASEN,
105: are acknowledged.}
106: \title[WKB for NLS]{WKB analysis for nonlinear Schr\"odinger equations
107: with potential}
108: \begin{abstract}
109: We justify the WKB analysis for the semiclassical nonlinear
110: Schr\"o\-din\-ger equation with a subquadratic potential. This concerns
111: subcritical, critical, and supercritical cases as far as the
112: geometrical optics method is concerned. In the supercritical case,
113: this extends a previous result by E.~Grenier; we also have to
114: restrict to nonlinearities which are defocusing and cubic at the
115: origin, but besides subquadratic potentials, we consider
116: initial phases which may be unbounded. For this, we construct solutions
117: for some compressible Euler equations with unbounded
118: source term and unbounded initial velocity.
119: \end{abstract}
120: \subjclass[2000]{35B30, 35B33, 35B40, 35C20, 35Q55, 81Q20}
121: \keywords{Nonlinear Schr\"odinger equation,
122: semiclassical analysis,
123: critical indices, supercritical WKB analysis, Euler equation}
124: \maketitle
125:
126:
127: \numberwithin{equation}{section}
128:
129:
130:
131:
132:
133: \section{Introduction}
134: \label{sec:intro}
135: Consider the initial value problem, for $x\in \R^n$ and $\kappa \ge 0$:
136: \begin{align}
137: i\e \d_t u^\e +\frac{\e^2}{2}\Delta u^\e &= V(t,x)u^\e + \e^\kappa
138: f\(|u^\e|^2\)
139: u^\e\label{eq:nlssemi}\\
140: u^\e_{\mid t=0} &= a_0^\e(x)e^{i\phi_0(x)/\e} .\label{eq:CI}
141: \end{align}
142: The aim of WKB methods is to describe $u^\e$ in the limit $\e \to 0$,
143: when $\phi_0$ does not depend on $\e$, and
144: $a_0^\e$ has an asymptotic expansion of the form:
145: \begin{equation}\label{eq:CIbkw}
146: a_0^\e (x)\sim a_0(x) +\e a_1(x)+\e^2 a_2(x) +\ldots
147: \end{equation}
148: The parameter $\kappa \ge 0$ describes the strength of a coupling
149: constant, which makes nonlinear effects more or less important in the
150: limit $\e \to 0$; the larger the $\kappa$, the weaker the nonlinear
151: interactions. In this paper, we describe the asymptotic behavior of
152: $u^\e$ at leading order, when the potential $V$ and the initial phase
153: $\phi_0$ are smooth, and subquadratic in the space variable.
154:
155: \smallbreak
156:
157: Such equations as \eqref{eq:nlssemi} appear in physics: see
158: e.g. \cite{Sulem} for a general overview. For instance, they are used to
159: model Bose-Einstein condensation when $V$ is an
160: harmonic potential (isotropic or anisotropic) and the nonlinearity is
161: cubic or quintic (see e.g. \cite{DGPS,KNSQ,PiSt}). In most of this
162: paper, the initial data
163: we consider are in Sobolev spaces $H^s$.
164: %We outline a WKB analysis for
165: %the Gross-Pitaevskii equation in Appendix~\ref{sec:gross}.
166: We refer to \cite{BJM} for numerics on the semi-classical limit of
167: \eqref{eq:nlssemi}.
168:
169: \smallbreak
170:
171: We shall not recall the results concerning the Cauchy
172: problem for \eqref{eq:nlssemi}-\eqref{eq:CI}, and refer to
173: \cite{CazCourant}
174: for an overview on the semilinear Schr\"odinger equation.
175:
176: \smallbreak
177:
178: In the one-dimensional case $n=1$, the cubic nonlinear
179: Schr\"odinger equation is completely integrable, in the absence of an
180: external potential, or when $V$ is a quadratic polynomial
181: (\cite[p.~375]{AblowitzClarkson}). Several tough papers
182: analyze the semi-classical limit in the case $V\equiv 0$, for $\kappa =0$: see
183: e.g. \cite{JLM,KMM,TianYe,TVZ}. We shall not use this approach, but
184: rather work in the spirit of \cite{Grenier98}.
185:
186: \smallbreak
187:
188: An interesting feature of \eqref{eq:nlssemi} is that one does not
189: expect the creation of harmonics, provided that only one phase is
190: present initially, like in \eqref{eq:CI}. The WKB methods consist in
191: seeking an approximate solution to \eqref{eq:nlssemi} of the form:
192: \begin{equation}\label{eq:defBKW}
193: u^\e(t,x)\sim \(\a_0(t,x) + \e \a_1(t,x) + \e^2
194: \a_2(t,x)+\ldots \) e^{i\Phi(t,x)/\e}\, .
195: \end{equation}
196: One must not expect this approach to be valid when caustics are
197: formed: near a caustic, all the terms
198: $\Phi$, $\a_0$, $\a_1$, \ldots become singular. Past the
199: caustic, several phases are necessary in general to describe the
200: asymptotic behavior of the solution (see e.g. \cite{Du} for a general
201: theory in the linear case). In this paper,
202: we restrict our attention to times preceding this break-up.
203: \smallbreak
204:
205: For such an expansion to be available with profiles $\a_j$
206: independent of $\e$, it is reasonable to assume that $\kappa$ is an
207: integer. However, we do not assume that $\kappa$ is an integer:
208: we study the asymptotic behavior of $u^\e$ \emph{at leading order}
209: (strong limits in $L^2\cap
210: L^\infty$ for instance), including cases where other powers of $\e$
211: would come into play.
212: \smallbreak
213:
214: We distinguish two families of assumptions: ``geometrical''
215: assumptions, on the potential $V$ and the initial phase $\phi_0$, and
216: ``analytical'' assumptions, on $f$ and the initial amplitude
217: $a_0^\e$. In all the cases, we shall not try to seek the optimal
218: regularity; we focus our interest on the limit $\e\to 0$.
219:
220: \begin{hyp}[Geometrical]\label{hyp:geom}
221: We assume that the potential and the initial phase are smooth and
222: subquadratic:
223: \begin{itemize}
224: \item $V\in C^\infty(\R_t\times \R^n_x)$, and $\d_x^\alpha V\in
225: L^\infty_{\rm loc}(\R_t ;L^\infty(\R^n_x))$ as soon as $|\alpha|\ge 2$.
226: \item $\phi_0\in C^\infty( \R^n)$, and $\d_x^\alpha \phi_0\in
227: L^\infty(\R^n)$ as soon as $|\alpha|\ge 2$.
228: \end{itemize}
229: \end{hyp}
230: The assumption of $V$ being subquadratic is classical in other
231: contexts; for instance, locally in time, the dispersion for
232: $e^{-i\frac{t}{\e}(-\e^2\Delta +V)}$ is the same as without potential
233: (see \cite{Fujiwara79,Fujiwara}),
234: \begin{equation*}
235: \left\| e^{-i\frac{t}{\e}(-\e^2\Delta +V)}\right\|_{L^1\to L^\infty}
236: \le \frac{C}{|\e t|^{n/2}},\quad \forall |t|\le \delta,
237: \end{equation*}
238: hence yielding the same local Strichartz
239: estimates as in the free case. Global in time Strichartz
240: estimates must not be expected in general, as shown by the example
241: of the harmonic oscillator, which has eigenvalues. For
242: positive superquadratic potentials, the
243: smoothing effects and Strichartz estimates are different (see
244: \cite{YajZha01,YajZha04}). This is related to the
245: properties of the Hamilton flow, which also imply:
246: \begin{lem}\label{lem:hj}
247: Under Assumption~\ref{hyp:geom}, there exist $T>0$ and a unique
248: solution $\eik\in C^\infty([0,T]\times\R^n)$ to:
249: \begin{equation}
250: \label{eq:eik}
251: \d_t \eik +\frac{1}{2}|\nabla_x \eik|^2 +V(t,x)=0\quad ;\quad
252: \phi_{{\rm eik} \mid t=0}=\phi_0\, .
253: \end{equation}
254: This solution is subquadratic: $\d_x^\alpha \eik \in
255: L^\infty([0,T]\times\R^n)$ as soon as $|\alpha|\ge 2$.
256: \end{lem}
257: This result is proved in Section~\ref{sec:HJ}, where other remarks on
258: Assumption~\ref{hyp:geom} are made.
259:
260: \begin{hyp}[Analytical]\label{hyp:ana}
261: We assume that the nonlinearity is smooth, and
262: that the initial amplitude converges in
263: Sobolev spaces:
264: \begin{itemize}
265: \item $f\in C^\infty(\R; \R)$.
266: \item There exists $ a_0\in H^\infty := \cap_{s\ge
267: 0}H^s(\R^n)$, such that
268: $ a_0^\e$ converges to $a_0$ in $H^s$ for any $s$, as
269: $\e\to 0$.
270: \end{itemize}
271: \end{hyp}
272: \begin{rema*}
273: Some of the results we shall prove remain valid when $f$ is
274: complex-valued. In that case, the conservation of mass associated to
275: the Schr\"odinger equation,
276: $\|u^\e(t)\|_{L^2}= \|a_0^\e\|_{L^2}$, no longer holds. On the other
277: hand, when $0\le \kappa <1$, this assumption is necessary in
278: our approach, and we even assume $f'>0$.
279: \end{rema*}
280: \subsection{Subcritical and critical cases: $\kappa \ge 1$}
281: \label{sec:sub0}
282:
283: When the initial data is of the form \eqref{eq:CIbkw}, the usual
284: approach consists in plugging a formal expansion of the form
285: \eqref{eq:defBKW} into \eqref{eq:nlssemi}. Ordering the terms in
286: powers of $\e$, and canceling the cascade of equations thus obtained
287: yields $\Phi$, $\a_0$, $\a_1$, \ldots
288: \smallbreak
289:
290: Assume in this section
291: that $\kappa \ge 1$, and apply the above procedure. To cancel the term
292: of order $\O(\e^0)$, we find that $\Phi$ must solve
293: \eqref{eq:eik}: $\Phi=\eik$. Canceling the term of order $\O(\e^1)$,
294: we get:
295: \begin{equation*}
296: \d_t \a_0 +\nabla \eik\cdot \nabla \a_0 +
297: \frac{1}{2}\a_0\Delta \eik =
298: \left\{
299: \begin{aligned}
300: &0 & \text{ if }\kappa >1,\\
301: &-if\(|\a_0|^2\)\a_0& \text{ if }\kappa =1.
302: \end{aligned}
303: \right.
304: \end{equation*}
305: We see that the value $\kappa =1$ is critical as far as
306: nonlinear effects are concerned: if $\kappa>1$, no nonlinear effect is
307: expected at leading order, since formally, $u^\e \sim \a_0 e^{i\eik
308: /\e}$, and $\eik$ and $\a_0$ do not depend on the nonlinearity
309: $f$. If $\kappa =1$, then $\a_0$ solves a nonlinear equation
310: involving $f$.
311: \smallbreak
312:
313: We will see in Section~\ref{sec:sub} that $\a_0$ solves a transport
314: equation that turns out to be a ordinary differential equation along
315: the rays of geometrical optics, as is usual in the hyperbolic case
316: (see e.g. \cite{RauchUtah}). More typical of Schr\"odinger equation is
317: the fact that this ordinary differential equation can be solved
318: explicitly: the nonlinear effect is measured by a nonlinear phase
319: shift (see the example of \cite{LannesRauch} for a similar result
320: in the hyperbolic setting). We prove the following result in
321: Section~\ref{sec:sub}:
322: \begin{prop}\label{prop:sub}
323: Let Assumptions~\ref{hyp:geom} and \ref{hyp:ana} be satisfied. Let
324: $\kappa \ge 1$. Then for any $\e \in ]0,1]$,
325: \eqref{eq:nlssemi}-\eqref{eq:CI} has a unique solution $u^\e \in
326: C^\infty([0,T]\times \R^n)\cap C([0,T];H^s)$ for any $s>n/2$ ($T$ is
327: given by
328: Lemma~\ref{lem:hj}). Moreover,
329: there exist $ a, G\in C^\infty([0,T]\times \R^n)$,
330: independent of $\e \in ]0,1]$, where
331: $ a \in C([0,T]; L^2\cap L^\infty)$, and $G$ is
332: real-valued with
333: $G \in C([0,T];
334: L^\infty)$, such that:
335: \begin{equation*}
336: \left\| u^\e - a e^{i\e^{\kappa -1}G}e^{i\eik
337: /\e}\right\|_{L^\infty([0,T]; L^2\cap L^\infty) } \to 0\quad
338: \text{as }\e \to 0.
339: \end{equation*}
340: The profile $ a$ solves the initial value problem:
341: \begin{equation}\label{eq:alibre}
342: \d_t a +\nabla \eik\cdot \nabla a +
343: \frac{1}{2} a\Delta \eik =0\quad ;\quad a_{\mid
344: t=0}=a_0,
345: \end{equation}
346: and $G$ depends nonlinearly on $ a$ through $f$.
347: In particular, if $\kappa>1$, then
348: \begin{equation*}
349: \left\| u^\e - a e^{i\eik
350: /\e}\right\|_{L^\infty([0,T]; L^2\cap L^\infty) } \to 0\quad
351: \text{as }\e \to 0,
352: \end{equation*}
353: and no nonlinear effect is present in the leading order behavior of
354: $u^\e$. If $\kappa =1$, nonlinear effects are present at leading
355: order, measured by $G$.
356: \end{prop}
357: The dependence of $G$ upon $ a$ and $f$ is made more explicit in
358: Section~\ref{sec:sub}, in terms of the Hamilton flow determining
359: $\eik$ (see \eqref{eq:exprG}). Note that in the above result, we do
360: not assume that $\kappa$ is an integer.
361:
362: \subsection{Supercritical case: $\kappa=0$}
363: \label{sec:surcrit0}
364:
365: It follows from the above analysis that the case $0\le \kappa <1$ is
366: supercritical. We restrict our attention to the case $\kappa =0$. We
367: present an analysis of the range $0<\kappa<1$ in
368: Section~\ref{sec:kappaint}. Plugging an
369: asymptotic expansion of the form \eqref{eq:defBKW} into
370: \eqref{eq:nlssemi} yields a shifted cascade of equations:
371: \begin{align*}
372: \O\(\e^0\):&\quad \d_t \Phi +\frac{1}{2}|\nabla\Phi|^2 + V +
373: f\(|\a_0|^2\)=0,\\
374: \O\(\e^1\):&\quad \d_t \a_0 +\nabla\Phi \cdot \nabla \a_0
375: +\frac{1}{2}\a_0\Delta \Phi =
376: 2if'\(|\a_0|^2\) {\rm Re}\(\a_0\overline{\a_1}\).
377: \end{align*}
378: Two comments are in order. First, we see that there is a strong
379: coupling between the phase and the main amplitude: $\a_0$ is present
380: in the equation for $\Phi$. Second, the above system is not closed:
381: $\Phi$ is determined in function of $\a_0$, and $\a_0$ is determined in
382: function of $\a_1$. Even if we pursued the cascade of equations, this
383: phenomenon would remain: no matter how many terms are computed, the
384: system is never closed (see \cite{PGX93}). This is a typical feature
385: of supercritical
386: cases in nonlinear geometrical optics (see
387: \cite{CheverryBullSMF,CG05}).
388: \smallbreak
389:
390: In the case when $V\equiv 0$ and $\phi_0\in H^s$, this problem was
391: resolved by E.~Grenier \cite{Grenier98}, by modifying the usual WKB
392: methods; this approach is recalled in Section~\ref{sec:grenier}. Note
393: that even though $\a_1$ is not determined by the above system, the
394: pair $(\rho,v):= (|\a_0|^2,\nabla \Phi)$ solves a compressible Euler
395: equation:
396: \begin{equation}
397: \label{eq:eulerV}
398: \begin{aligned}
399: \d_t v +v\cdot \nabla v + \nabla V + \nabla f(\rho)&=0\ ;\quad
400: v\big|_{ t=0}=\nabla \phi_0\\
401: \d_t \rho+ \nabla\cdot (\rho v) &=0\ ;\quad \rho\big|_{t=0}=|a_0|^2.
402: \end{aligned}
403: \end{equation}
404: Using techniques introduced in the study of quasilinear hyperbolic
405: equations, E.~Grenier justified a WKB expansion for nonlinearities
406: which are defocusing, and cubic at the origin ($f'>0$). We shall not
407: change this assumption, but show how to treat the case of a
408: subquadratic potential with a subquadratic initial phase. Note that
409: even the construction of solution to \eqref{eq:eulerV} under
410: Assumption~\ref{hyp:geom} is not standard: the source term $ \nabla V$
411: may be unbounded, as well as the initial velocity $\nabla \phi_0$.
412: \begin{hyp}\label{hyp:surcrit}
413: In addition to Assumption~\ref{hyp:ana}, we assume:
414: \begin{itemize}
415: \item $f'>0$.
416: \item There exists $a_0, a_1\in H^\infty$, with $xa_0, xa_1\in H^\infty$,
417: such that:
418: \begin{equation*}
419: \left\|a_0^\e -a_0 -\e a_1\right\|_{H^s} +
420: \left\|xa_0^\e -xa_0 -\e xa_1\right\|_{H^s}
421: =o(\e),\quad \forall s\ge 0.
422: \end{equation*}
423: \end{itemize}
424: \end{hyp}
425: We can then describe the asymptotic behavior of the solution to
426: \eqref{eq:nlssemi}-\eqref{eq:CI} for small times:
427: \begin{theo}\label{theo:BKWV}
428: Let Assumptions~\ref{hyp:geom}, \ref{hyp:ana} and \ref{hyp:surcrit}
429: be satisfied. Let $\kappa =0$. There exists $T_*>0$ independent
430: of $\e \in ]0,1]$ and a unique solution $u^\e \in
431: C^\infty([0,T_*]\times \R^n)\cap C([0,T_*];H^s)$ for any $s>n/2$ to
432: \eqref{eq:nlssemi}-\eqref{eq:CI}. Moreover, there exist $a,\phi \in
433: C([0,T_*];H^s)$ for every $s\ge 0$, such that:
434: \begin{equation}\label{eq:BKWVtpetit}
435: \limsup_{\e \to 0}\left\| u^\e - a
436: e^{i(\phi+\eik)/\e}\right\|_{L^2\cap L^\infty}=\O(t)\quad
437: \text{as }t\to 0.
438: \end{equation}
439: Here, $a$ and $\phi$ are nonlinear functions of $\eik$ and
440: $a_0$, given by \eqref{eq:11h09}. Finally, there exists $\phi^{(1)}\in
441: C([0,T_*];H^s)$ for every $s\ge 0$, real-valued, such that:
442: \begin{equation}\label{eq:BKWVtgrand}
443: \limsup_{\e \to 0}\sup_{0\le t\le T_*}\left\| u^\e - ae^{i\phi^{(1)}}
444: e^{i(\phi+\eik)/\e}\right\|_{L^2\cap L^\infty}=0.
445: \end{equation}
446: The phase shift $\phi^{(1)}$ is a nonlinear function of $\eik, a_0$
447: and $a_1$.
448: \end{theo}
449: This result can be understood as follows. At leading order, the
450: amplitude of $u^\e$ is given by $a e^{i\phi^{(1)}}$, which can be
451: approximated for small times by $a$, because
452: $\phi^{(1)}\big|_{t=0}=0$. The rapid oscillations are described by the
453: phase $\phi +\eik$. The
454: function $\phi$ is constructed as a perturbation of $\eik$, but must not
455: be considered as negligible: its $H^s$-norms are not small in
456: general, see \eqref{eq:11h09} (at time $t=0$ for instance). As a
457: consequence of our analysis, the pair
458: \begin{equation*}
459: (\rho,v) = \( |a|^2,\nabla (\phi +\eik)\) = \(
460: \left|a e^{i\phi^{(1)}}\right|^2,\nabla (\phi +\eik)\)
461: \end{equation*}
462: solves the system \eqref{eq:eulerV}.
463: \begin{rema}
464: With this result, we could deduce instability phenomena for
465: \eqref{eq:nlssemi}-\eqref{eq:CI} in the same fashion as in
466: \cite{CaInstab}. Note that because of Assumption~\ref{hyp:geom}, it
467: seems that the
468: approaches of \cite{BGTENS,BZ,CCT2} cannot be adapted to the
469: present case: the Laplacian can never be neglected, and apparently,
470: WKB approach is really needed.
471: \end{rema}
472:
473:
474: The rest of this paper is organized as follows. In
475: Section~\ref{sec:HJ}, we prove Lemma~\ref{lem:hj}. In
476: Section~\ref{sec:sub}, we prove Proposition~\ref{prop:sub}, and
477: explain how $G$ is obtained. We recall the approach of
478: \cite{Grenier98} in Section~\ref{sec:grenier}, and show how to adapt
479: it to prove Theorem~\ref{theo:BKWV}
480: in Section~\ref{sec:surcrit}.
481: We present an
482: analysis for the case $0<\kappa<1$ in Section~\ref{sec:kappaint}.
483: %We sketch the
484: %proof of the analogue of Theorem~\ref{theo:BKWV} for the
485: %Gross-Pitaevskii equation in Appendix~\ref{sec:gross}.
486:
487: %%%%%%%%%%%%%%%%%
488:
489:
490:
491:
492:
493:
494: \section{Global in space Hamilton-Jacobi theory}
495: \label{sec:HJ}
496: In this section, we prove Lemma~\ref{lem:hj}.
497: Consider the classical Hamiltonian:
498: \begin{equation*}
499: H(t,x,\tau,\xi)= \tau +\frac{1}{2}|\xi|^2 +V(t,x),\quad
500: (t,x,\tau,\xi)\in \R_+\times\R^n \times\R\times\R^n.
501: \end{equation*}
502: It is smooth by Assumption~\ref{hyp:geom}. Therefore, it is classical
503: (see e.g. \cite{DG}) that in the neighborhood of each
504: point $x\in \R^n$, one can construct a smooth solution to the eikonal
505: equation \eqref{eq:eik}, on some time interval $[-t(x),t(x)]$, for
506: some $t(x)>0$ depending on $x$. The fact
507: that in Lemma~\ref{lem:hj}, we can find some $T>0$ uniform in $x\in
508: \R^n$ is due to the fact that the potential and the initial phase are
509: subquadratic.
510: \smallbreak
511:
512: Recall that if there
513: exist some constants $a,b>0$ such that a potential $\tt V$ satisfies
514: ${\tt V}(x)\geq -a |x|^2-b$, then $-\Delta +{\tt V}$ is
515: essentially self-adjoint on $C_0^\infty({\mathbb R}^n)$ (see
516: \cite[p.~199]{ReedSimon2}). If $-{\tt V}$ has
517: superquadratic growth,
518: then it is not possible to define $e^{-it(-\Delta +{\tt V})}$
519: (see \cite[Chap.~13, Sect.~6, Cor.~22]{Dunford} for the case ${\tt
520: V}(x)=-x^4$ in space dimension one). This is due to the fact that
521: classical trajectories can reach an infinite speed. We will see below
522: that if the initial phase $\phi_0$ is superquadratic, then focusing at
523: the origin may occur ``instantly'' (see Example~\ref{ex:phase}).
524: \smallbreak
525:
526: To construct the solution of \eqref{eq:eik}, introduce the flow
527: associated to $H$: let $x(t,y)$ and $\xi(t,y)$ solve
528: \begin{equation}
529: \label{eq:hamilton}
530: \left\{
531: \begin{aligned}
532: &\d_t x(t,y) = \xi \(t,y\)\quad ;\quad x(0,y)=y,\\
533: &\d_t \xi(t,y) = -\nabla_x V\(t,x(t,y)\)\quad ;\quad
534: \xi(0,y)=\nabla \phi_0(y).
535: \end{aligned}
536: \right.
537: \end{equation}
538: Recall the result (valid under weaker conditions than
539: Assumption~\ref{hyp:geom}):
540: \begin{theo}[\cite{DG}, Th.~A.3.2]\label{theo:hj}
541: Suppose that Assumption~\ref{hyp:geom} is satisfied. Let $t\in
542: [0,T]$ and $\theta_0$ an open set of $\R^n$. Denote
543: \begin{equation*}
544: \theta_t := \{ x(t,y)\in \R^n, y\in\theta_0\}\quad ;\quad \theta
545: := \{ (t,x)\in [0,T]\times\R^n, x\in \theta_t\}.
546: \end{equation*}
547: Suppose that for $t\in [0,T]$, the mapping
548: \begin{equation*}
549: \theta_0\ni y\mapsto x(t,y)\in \theta_t
550: \end{equation*}
551: is bijective, and denote by $y(t,x)$ its inverse. Assume also that
552: \begin{equation*}
553: \nabla_x y \in L^\infty_{\rm loc}(\theta).
554: \end{equation*}
555: Then there exists a unique function $\theta\ni (t,x)\mapsto \eik(t,x)\in
556: \R$ that solves \eqref{eq:eik}, and satisfies $\nabla_x^2\eik \in
557: L^\infty_{\rm loc}(\theta)$. Moreover,
558: \begin{equation}\label{eq:gradeik}
559: \nabla_x \eik(t,x) =\xi(t,y(t,x)).
560: \end{equation}
561: \end{theo}
562:
563: \begin{prop}[\cite{SchwartzBook}, Th.~1.22 and \cite{DG},
564: Prop.~A.7.1]\label{prop:invglobale}
565: Suppose that the function $\R^n\ni y\mapsto x(y)\in \R^n$ satisfies:
566: \begin{equation*}
567: |\operatorname{det}\nabla_y x|\ge C_0>0\quad \text{and}\quad \left|
568: \d_y^\alpha x\right|\le C, \ |\alpha|=1,2.
569: \end{equation*}
570: Then $x$ is bijective.
571: \end{prop}
572: \begin{proof}[Proof of Lemma~\ref{lem:hj}]
573: Lemma~\ref{lem:hj} follows from the above two
574: results. From Assumption~\ref{hyp:geom}, we know that we can solve
575: \eqref{eq:hamilton} locally in time in the neighborhood of any
576: $y\in\R^n$. Differentiate \eqref{eq:hamilton} with respect to $y$:
577: \begin{equation}
578: \label{eq:dyhamilton}
579: \left\{
580: \begin{aligned}
581: &\d_t \d_y x(t,y) = \d_y\xi \(t,y\)\quad ;\quad \d_y
582: x(0,y)={\rm Id},\\
583: &\d_t \d_y \xi(t,y) = -\nabla_x^2 V\(t,x(t,y)\)\d_y x(t,y) \quad ;\quad
584: \d_y\xi(0,y)=\nabla^2 \phi_0(y).
585: \end{aligned}
586: \right.
587: \end{equation}
588: Integrating \eqref{eq:dyhamilton} in time, we infer from
589: Assumption~\ref{hyp:geom} that for any $T>0$, there exists $C_T$ such
590: that for $(t,y)\in [0,T]\times \R^n$:
591: \begin{equation*}
592: \left| \d_y x(t,y) \right| +\left| \d_y \xi(t,y)\right|
593: \le C_T + C_T\int_0^t \( \left| \d_y x(s,y) \right| +\left| \d_y
594: \xi(s,y)\right| \) ds.
595: \end{equation*}
596: Gronwall lemma yields:
597: \begin{equation}\label{eq:10:04}
598: \left\| \d_y x(t) \right\|_{L^\infty_y} +\left\| \d_y
599: \xi(t)\right\|_{L^\infty_y}
600: \le C'(T).
601: \end{equation}
602: Similarly,
603: \begin{equation}\label{eq:10:05}
604: \left\| \d_y^\alpha x(t) \right\|_{L^\infty_y} +\left\|
605: \d_y^\alpha \xi(t)\right\|_{L^\infty_y}
606: \le C(\alpha,T),\quad \forall \alpha \in \N^n,\ |\alpha|\ge 1.
607: \end{equation}
608: Integrating the first line of \eqref{eq:dyhamilton} in time, we
609: have:
610: \begin{equation*}
611: \operatorname{det}\nabla_y x(t,y) = \operatorname{det}\({\rm
612: Id}+\int_0^t \nabla_y\xi \(s,y\)ds\).
613: \end{equation*}
614: We infer from \eqref{eq:10:04} that for $t\in [0,T]$, provided that
615: $T>0$ is sufficiently small, we can find $C_0>0$ such that:
616: \begin{equation}\label{eq:det}
617: \left| \operatorname{det}\nabla_y x(t,y)\right|\ge C_0,\quad \forall
618: (t,y)\in [0,T]\times \R^n.
619: \end{equation}
620: Applying Proposition~\ref{prop:invglobale}, we deduce that we can
621: invert $y\mapsto x(t,y)$ for $t\in [0,T]$.
622: \smallbreak
623:
624: To apply Theorem~\ref{theo:hj} with $\theta_0= \theta= \theta_t=\R^n$,
625: we must check that $\nabla_x y\in L^\infty_{\rm loc}
626: (\R^n)$. Differentiate the relation
627: \begin{equation*}
628: x\(t,y(t,x)\) = x
629: \end{equation*}
630: with respect to $x$:
631: \begin{equation*}
632: \nabla_x y(t,x)\nabla_y x\(t,y(t,x)\) = {\rm Id}.
633: \end{equation*}
634: Therefore, $\nabla_x y(t,x) = \nabla_y x\(t,y(t,x)\)^{-1}$ as
635: matrices, and
636: \begin{equation}\label{eq:adj}
637: \nabla_x y(t,x) =\frac{1}{\operatorname{det}\nabla_y x(t,y)}{\rm
638: adj}\( \nabla_y x\(t,y(t,x)\)\),
639: \end{equation}
640: where ${\rm adj}\( \nabla_y x\)$ denotes the adjugate of $\nabla_y
641: x$. We infer from \eqref{eq:10:04} and \eqref{eq:det} that $\nabla_x y
642: \in L^\infty(\R^n)$ for $t\in [0,T]$. Therefore, Theorem~\ref{theo:hj}
643: yields a smooth solution $\eik$ to \eqref{eq:eik}, local in time and
644: global in space: $\eik \in C^\infty([0,T]\times \R^n)$.
645: \smallbreak
646:
647: The fact that $\eik$ is subquadratic as stated in Lemma~\ref{lem:hj}
648: then stems from \eqref{eq:gradeik}, \eqref{eq:10:05}, \eqref{eq:det}
649: and \eqref{eq:adj}.
650: \end{proof}
651:
652: We now give some two examples showing that
653: Assumption~\ref{hyp:geom} is essentially sharp to solve \eqref{eq:eik}
654: globally in space, at least when no assumption is made on the sign of
655: $V$ nor on the geometry of $\nabla \phi_0$. We already recalled that
656: if $-V$ has a
657: superquadratic growth, then $-\Delta +V$ is not essentially
658: self-adjoint on $C_0^\infty(\R^n)$, so we shall rather study the
659: dependence of $\eik$ on the initial phase $\phi_0$.
660: \begin{ex}\label{ex:phase}
661: Assume that $V\equiv 0$ and
662: \begin{equation*}
663: \phi_0(x) = -\frac{1}{(2+2\delta)T}\(|x|^2+1 \)^{1+\delta}, \quad
664: T>0,\ \delta
665: \not = -1.
666: \end{equation*}
667: Then Assumption~\ref{hyp:geom} is satisfied if and only if $\delta\le
668: 0$. When $\delta =0$, then \eqref{eq:eik} is solved explicitly:
669: \begin{equation*}
670: \eik(t,x) = \frac{|x|^2}{2(t-T)}-\frac{1}{2T}.
671: \end{equation*}
672: This shows that we can solve \eqref{eq:eik} globally in space, but only
673: locally in time: as $t\to T$, a caustic reduced to a single point (the
674: origin) is
675: formed. Note that with $T<0$, \eqref{eq:eik} can be solved globally in
676: time \emph{for positive times}.\\
677: When $\delta>0$, then integrating \eqref{eq:hamilton} yields:
678: \begin{align*}
679: x(t,y) &=y +\int_0^t \xi(s,y)ds = y +\int_0^t \xi(0,y)ds= y -
680: \frac{t}{T}\(|y|^2+1 \)^{\delta} y\\
681: &= y\(1 - \frac{t}{T}\(|y|^2+1
682: \)^{\delta}\).
683: \end{align*}
684: For $R>0$, we see that the rays starting from the
685: ball $\{|y|=R\}$ meet at the origin at time
686: \begin{align*}
687: T_c(R) = \frac{T}{(R^2+1)^\delta}.
688: \end{align*}
689: Since $R$ is arbitrary, this shows that several rays can meet
690: arbitrarily fast, thus showing that Theorem~\ref{theo:hj} cannot be
691: applied uniformly in space.
692: \end{ex}
693: \begin{ex}
694: When $V(t,x) = \frac{1}{2}\sum_{j=1}^n \om_j^2 x_j^2$ is an harmonic
695: potential ($\om_j \not= 0$), and $\phi_0\equiv 0$, we have:
696: \begin{equation*}
697: \phi(t,x) = -\sum_{j=1}^n \frac{\om_j}{2}x_j^2\tan (\om_j t).
698: \end{equation*}
699: This also shows that we can solve \eqref{eq:eik} globally in space,
700: but locally in time only. Note that if we replace formally $\om_j$ by
701: $i\om_j$, then $V$ is turned into $-V$, and the trigonometric
702: functions become hyperbolic functions: we can then solve \eqref{eq:eik}
703: globally in space \emph{and} time.
704: \end{ex}
705:
706: Instead of invoking Theorem~\ref{theo:hj} and
707: Proposition~\ref{prop:invglobale}, one might try to differentiate
708: \eqref{eq:eik} in order to prove that $\eik$ is subquadratic, in the
709: same fashion as in \cite{BahouriCheminAJM,BahouriCheminIMRN}. For
710: $1\le j,k\le n$, differentiate \eqref{eq:eik} with respect to $x_j$
711: and $x_k$:
712: \begin{align*}
713: &\d_t \d^2_{jk}\eik +\nabla_x \eik\cdot \nabla_x\(\d^2_{jk}\eik \)+
714: \sum_{l=1}^n \d^2_{jl}\eik\d^2_{lk}\eik
715: +\d^2_{jk}V(t,x)=0\ ;\\
716: &\d^2_{jk}\phi_{{\rm eik} \mid t=0}=\d^2_{jk}\phi_0\, .
717: \end{align*}
718: We see that we obtain a system of the form
719: \begin{equation*}
720: D_t y = Q(y) + R\quad ;\quad y_{\mid t=0} = y(0),
721: \end{equation*}
722: where $y$ stands for the family $(\d^2_{jk}\eik)_{1\le j,k\le n}$, $Q$
723: is quadratic, and
724: $R$ and $y(0)$ are bounded. The operator $D_t$ is a well-defined
725: transport operator provided that the characteristics given by:
726: \begin{equation*}
727: \d_t x(t,y) = \nabla_x \eik \(t,x(t,y)\)\quad ; \quad x(0,y)=y,
728: \end{equation*}
729: define a global diffeomorphism. Proving this amounts to using
730: Proposition~\ref{prop:invglobale}, for the rather general initial data
731: we consider. So it seems that this approach does not allow to shorten
732: the proof of Lemma~\ref{lem:hj}.
733:
734:
735: %%%%%%%%%%%%%%%%%%%%%%
736:
737:
738:
739: \section{Subcritical and critical cases}
740: \label{sec:sub}
741:
742: To establish Proposition~\ref{prop:sub}, define
743: \begin{equation*}
744: a^\e(t,x) := u^\e (t,x) e^{-i\eik(t,x)/\e}.
745: \end{equation*}
746: Then $u^\e$ solves \eqref{eq:nlssemi}-\eqref{eq:CI} if and only if
747: $a^\e$ solves:
748: \begin{equation}
749: \label{eq:asub}
750: \begin{aligned}
751: \d_t a^\e +\nabla \eik\cdot \nabla a^\e +
752: \frac{1}{2}a^\e\Delta \eik &=i\frac{\e}{2}\Delta a^\e
753: -i\e^{\kappa -1}f\(|a^\e|^2\)a^\e ,\\
754: a^\e_{\mid t=0}&=a_0^\e.
755: \end{aligned}
756: \end{equation}
757: \begin{prop}\label{prop:asub}
758: Let Assumptions~\ref{hyp:geom} and \ref{hyp:ana} be satisfied. Let
759: $\kappa \ge 1$. For any $\e \in ]0,1]$, \eqref{eq:asub} has a
760: unique solution $a^\e \in
761: C^\infty([0,T]\times \R^n)\cap C([0,T];H^s)$ for any
762: $s>n/2$. Moreover, $a^\e$ is bounded in
763: $L^\infty([0,T]; H^s)$ uniformly in $\e \in ]0,1]$, for any $s\ge
764: 0$.
765: \end{prop}
766: \begin{proof}
767: Using a mollification procedure, we see that it is enough to
768: establish energy estimates for \eqref{eq:asub} in $H^s$, for any
769: $s\ge 0$. Let $s>n/2$, and $\alpha\in \N^n$, with $s=|\alpha|$. Applying
770: $\d^\alpha_x$ to \eqref{eq:asub}, we find:
771: \begin{equation}
772: \label{eq:asubder}
773: \d_t \d^\alpha_x a^\e +\nabla \eik\cdot \nabla \d^\alpha_x a^\e
774: =i\frac{\e}{2}\Delta \d^\alpha_x a^\e
775: -i\e^{\kappa -1}\d^\alpha_x\(f\(|a^\e|^2\)a^\e\)
776: + R_\alpha^\e,
777: \end{equation}
778: where
779: \begin{equation*}
780: R_\alpha^\e = \left[ \nabla \eik\cdot \nabla ,\d^\alpha_x \right]a^\e -
781: \frac{1}{2}\d^\alpha_x \(a^\e\Delta \eik\).
782: \end{equation*}
783: Take the inner product of \eqref{eq:asubder} with $\d^\alpha a^\e$,
784: and consider the real part:
785: the first term of the right hand side of \eqref{eq:asubder} vanishes,
786: and we have:
787: \begin{equation*}
788: \begin{aligned}
789: \frac{1}{2}\frac{d}{dt}\|\d^\alpha_x a^\e\|_{L^2}^2+ {\rm Re}\int_{\R^n}
790: \overline{\d^\alpha_x a^\e}\nabla
791: \eik\cdot \nabla \d^\alpha_x a^\e \le &\e^{\kappa -1}
792: \left\|f\(|a^\e|^2\)a^\e\right\|_{H^s} \left\|a^\e\right\|_{H^s}\\
793: & +\left\|R_\alpha^\e\right\|_{L^2}\left\|a^\e\right\|_{H^s} .
794: \end{aligned}
795: \end{equation*}
796: Notice that we have
797: \begin{align*}
798: \left| {\rm Re}\int_{\R^n}
799: \overline{\d^\alpha_x a^\e}\nabla
800: \eik\cdot \nabla \d^\alpha_x a^\e \right| &= \frac{1}{2}\left| \int_{\R^n}
801: \nabla
802: \eik\cdot \nabla \left|\d^\alpha_x a^\e \right|^2\right|\\
803: & = \frac{1}{2}\left| \int_{\R^n}
804: \left|\d^\alpha_x a^\e \right|^2 \Delta \eik \right|\le C
805: \left\|a^\e\right\|_{H^s}^2,
806: \end{align*}
807: since $\Delta \eik \in L^\infty([0,T]\times\R^n)$ from
808: Lemma~\ref{lem:hj}. Moser's inequality yields:
809: \begin{equation*}
810: \left\|f\(|a^\e|^2\)a^\e\right\|_{H^s} \le C \(
811: \left\|a^\e\right\|_{L^\infty}\) \left\|a^\e\right\|_{H^s}.
812: \end{equation*}
813: Summing over $\alpha$ such that $|\alpha|=s$, we infer:
814: \begin{equation*}
815: \frac{d}{dt}\| a^\e\|_{H^s} \le C \(
816: \left\|a^\e\right\|_{L^\infty}\) \left\|a^\e\right\|_{H^s} +
817: \left\|R_\alpha^\e\right\|_{H^s}.
818: \end{equation*}
819: Note that the above locally bounded map $C(\cdot)$ is independent of
820: $\e$ if and only if $\kappa \ge 1$.
821: To apply Gronwall lemma, we need to estimate the last term: we use the
822: fact that the derivatives of order at least two of $\eik$ are bounded,
823: from Lemma~\ref{lem:hj}, to have:
824: \begin{equation*}
825: \left\|R_\alpha^\e\right\|_{L^2} \le C \left\|a^\e\right\|_{H^s}.
826: \end{equation*}
827: We can then conclude by a continuity argument and Gronwall lemma:
828: \begin{equation*}
829: \| a^\e\|_{L^\infty([0,T];H^s)}\le C\(s, \left\|a_0^\e\right\|_{H^s} \).
830: \end{equation*}
831: For the mollification procedure, this yields boundedness in the high
832: norm. Contraction in the small norm then follows easily with classical
833: arguments (see e.g. \cite{AlinhacGerard,Majda}), completing the proof
834: of Proposition~\ref{prop:asub}.
835: \end{proof}
836: \begin{cor}\label{cor:simplifsub}
837: Let Assumptions~\ref{hyp:geom} and \ref{hyp:ana} be satisfied. Let
838: $\kappa \ge 1$. Then
839: \begin{equation*}
840: \left\| a^\e - \widetilde a^\e
841: \right\|_{L^\infty([0,T];H^s)}\to 0 \quad \text{as }\e \to 0,\quad
842: \forall s\ge 0,
843: \end{equation*}
844: where $\widetilde a^\e$ solves:
845: \begin{equation}\label{eq:atildeeps}
846: \d_t \widetilde a^\e +\nabla \eik\cdot \nabla \widetilde a^\e +
847: \frac{1}{2}\widetilde a^\e\Delta \eik =
848: -i\e^{\kappa -1}f\(|\widetilde a^\e|^2\)\widetilde a^\e \quad ; \quad
849: \widetilde a^\e_{\mid t=0}=a_0.
850: \end{equation}
851: \end{cor}
852: \begin{proof}
853: The proof of Proposition~\ref{prop:asub} shows that $\widetilde a^\e \in
854: C^\infty([0,T]\times \R^n)$, and that $\widetilde a^\e$ is bounded in
855: $L^\infty([0,T]; H^s)$ uniformly in $\e \in ]0,1]$, for any $s\ge
856: 0$. Let $w^\e = a^\e -\widetilde a^\e$: it solves
857: \begin{align*}
858: \d_t w^\e +\nabla \eik\cdot \nabla w^\e +
859: \frac{1}{2}w^\e\Delta \eik &=i\frac{\e}{2}\Delta a^\e
860: -i\e^{\kappa -1}\(F(a^\e) -F(\widetilde a^\e)\),\\
861: w^\e_{\mid t=0}&=a_0^\e -a_0,
862: \end{align*}
863: where we have denoted $F(z)=f(|z|^2)z$. Proceeding as in the proof of
864: Proposition~\ref{prop:asub}, we have the following energy estimate:
865: \begin{equation*}
866: \frac{d}{dt}\| w^\e\|_{H^s} \le C \left\|w^\e\right\|_{H^s} +
867: \e \left\|\Delta a^\e\right\|_{H^s} + \left\|F(a^\e) -F(\widetilde
868: a^\e)\right\|_{H^s}.
869: \end{equation*}
870: Since $H^s$ is an algebra and $F$ is $C^1$, the Fundamental Theorem of
871: Calculus yields:
872: \begin{equation*}
873: \left\|F(a^\e) -F(\widetilde
874: a^\e)\right\|_{H^s}\le C\(\left\| a^\e\right\|_{H^s},
875: \left\|\widetilde a^\e\right\|_{H^s}\) \| w^\e\|_{H^s}.
876: \end{equation*}
877: Now since $a^\e$ and $\widetilde a^\e$ are bounded in
878: $L^\infty([0,T]; H^{s+2})$ uniformly in $\e \in ]0,1]$, we have an
879: estimate of the form:
880: \begin{equation*}
881: \frac{d}{dt}\| w^\e\|_{H^s} \le C(s) \left\|w^\e\right\|_{H^s} +
882: C(s) \e .
883: \end{equation*}
884: We conclude by Gronwall lemma, since $\|a_0^\e -a_0\|_{H^s} \to 0$
885: from Assumption~\ref{hyp:ana}.
886: \end{proof}
887: \begin{rema}
888: If we assume moreover that like in \eqref{eq:CIbkw},
889: \begin{equation*}
890: a_0^\e = a_0 + \O(\e) \quad \text{in }H^s, \ \forall s\ge 0,
891: \end{equation*}
892: then the above estimate can be improved to:
893: \begin{equation*}
894: \left\| a^\e - \widetilde a^\e
895: \right\|_{L^\infty([0,T];H^s)}=\O(\e),\quad \forall s\ge 0.
896: \end{equation*}
897: \end{rema}
898: We have reduced the study of the asymptotic behavior of $u^\e$ to the
899: understanding of $\widetilde a^\e$. To complete the proof of
900: Proposition~\ref{prop:sub}, we resume the framework of
901: Section~\ref{sec:HJ}. With $x(t,y)$ given by the Hamilton flow
902: \eqref{eq:hamilton}, introduce the Jacobi determinant
903: \begin{equation*}
904: J_t(y) ={\rm det}\nabla_y x(t,y).
905: \end{equation*}
906: Denote
907: \begin{equation*}
908: A^\e(t,y) := \widetilde a^\e \(t, x(t,y) \)\sqrt{J_t(y)}.
909: \end{equation*}
910: We see that so long as $y\mapsto x(t,y)$ defines a global
911: diffeomorphism (which is guaranteed for $t\in [0,T]$ by construction),
912: \eqref{eq:atildeeps} is equivalent to:
913: \begin{equation*}
914: \d_t A^\e = -i\e^{\kappa -1}f\(J_t(y)^{-1}\left|A^\e \right|^2\)
915: A^\e\quad ; \quad A^\e(0,y)=a_0(y).
916: \end{equation*}
917: This ordinary differential equation along the rays of geometrical
918: optics can be solved explicitly: since $f$ is real-valued, we see that
919: $\d_t |A^\e|^2 =0$, hence
920: \begin{equation*}
921: A^\e(t,y) = a_0(y) \exp\(-i\e^{\kappa -1} \int_0^t
922: f\(J_s(y)^{-1}\left|a_0(y) \right|^2\)ds\).
923: \end{equation*}
924: Back to the function $\widetilde a^\e$, Proposition~\ref{prop:sub}
925: follows, with:
926: \begin{equation}\label{eq:exprG}
927: \begin{aligned}
928: a(t,x) &= \frac{1}{\sqrt{J_t(y(t,x))}}a_0\(y(t,x)\),\\
929: G(t,x) &= -\int_0^t
930: f\(J_s(y(t,x))^{-1}\left|a_0(y(t,x)) \right|^2\)ds.
931: \end{aligned}
932: \end{equation}
933: One may wonder if this approach could be extended to some values
934: $\kappa <1$. Seek a solution of \eqref{eq:asub}. To have a simple ansatz
935: as in Proposition~\ref{prop:sub}, we would like to remove the
936: Laplacian in the limit $\e \to 0$ in \eqref{eq:asub}, and obtain the
937: analogue of Corollary~\ref{cor:simplifsub}. Following the same lines
938: as above, we find:
939: \begin{equation*}
940: \widetilde a^\e (t,x) = a(t,x) e^{i\e^{\kappa-1}G(t,x)},
941: \end{equation*}
942: which is exactly the first formula in Proposition~\ref{prop:sub}. Now
943: recall that in Proposition~\ref{prop:asub}, we prove that $a^\e$ is
944: bounded in $H^s$, uniformly for $\e \in ]0,1]$; this property is used
945: to approximate $a^\e$ by $\widetilde a^\e$. But when $\kappa <1$,
946: $\widetilde a^\e$ is no longer uniformly bounded in $H^s$, because
947: what was a phase modulation for $\kappa \ge 1$ is now a rapid
948: oscillation.
949: \smallbreak
950:
951: This remark in a particular case reveals a much more general
952: phenomenon. When studying geometric optics in a supercritical r\'egime
953: (when $\kappa <1$ in the present context), distinguishing phase and
954: amplitude becomes a much more delicate issue. Suppose for instance
955: that we seek $u^\e = a^\e e^{i\phi^\e/\e}$, where $a^\e$ and $\phi^\e$
956: have asymptotic expansions as $\e \to 0$. All the terms
957: for $\phi^\e$ which are not $o(\e)$ are relevant, since $\phi^\e$ is
958: divided by $\e$. To determine these terms, it is not sufficient to
959: determine the leading order amplitude $\lim a^\e$ in general: because of
960: supercritical interactions, initial perturbations of the amplitude
961: may develop non-negligible phase terms. To illustrate this discussion
962: in the above case, let $\kappa <1$ and $\widetilde a^\e$ solve
963: \eqref{eq:atildeeps} with initial data
964: \begin{equation*}
965: \widetilde a^\e_{\mid t=0} =a_0 +\e^\gamma a_1,\quad \gamma >0,
966: \end{equation*}
967: where $a_0$ and $a_1$ are smooth and independent of $\e$. Integrating
968: \eqref{eq:atildeeps}, we find
969: \begin{equation*}
970: \widetilde a^\e (t,x) = \frac{1}{\sqrt{J_t(y(t,x))}}\(
971: a_0\(y(t,x)\)+ \e^\gamma a_1\(y(t,x)\)\) e^{i\e^{\kappa -1}G^\e(t,x)},
972: \end{equation*}
973: where
974: \begin{equation*}
975: G^\e (t,x) = -\int_0^t
976: f\(J_s(y(t,x))^{-1}\left|a_0(y(t,x))+ \e^\gamma a_1\(y(t,x)\) \right|^2\)ds.
977: \end{equation*}
978: We have the identity $G^\e = G + \e^\gamma G_1 +\e^{2\gamma}G_2$, for
979: the same $G$ as before, and $G_1,G_2$ independent of $\e$, but
980: depending on $a_1$. We infer:
981: \begin{equation*}
982: \widetilde a^\e (t,x) \Eq \e 0 \frac{1}{\sqrt{J_t(y(t,x))}}
983: a_0\(y(t,x)\) e^{i\e^{\kappa -1}(G+\e^\gamma G_1 +\e^{2\gamma}G_2) (t,x)}.
984: \end{equation*}
985: In particular, if $\kappa +\gamma \le 1$, we see that $G_1$ has to be
986: incorporated to describe the leading order behavior of $\widetilde
987: a^\e$. Since the three requirements $\kappa <1$, $\gamma>0$ and
988: $\kappa +\gamma \le 1$ can be met, we see that the small initial
989: perturbation $\e^\gamma a_1$ may produce relevant phase
990: perturbations. This example explains why in
991: Assumption~\ref{hyp:surcrit}, we require the asymptotic behavior of
992: the initial amplitude up to order $o(\e)$ and not only $o(1)$ (take
993: $\kappa=0$ and $\gamma =1$).
994: We refer to
995: \cite{CheverryBullSMF} for a
996: general explanation of this phenomenon, and to the next three sections
997: as far as Schr\"odinger equations are concerned.
998:
999:
1000:
1001: %%%%%%%%%%%%%
1002:
1003:
1004: \section{A hyperbolic point of view}
1005: \label{sec:grenier}
1006:
1007: In this section, we study \eqref{eq:nlssemi} in the case
1008: $\kappa =0$, with no potential:
1009: \begin{equation}\label{eq:hs}
1010: i\e \d_t u^\e +\frac{\e^2}{2}\Delta u^\e = f\(|u^\e|^2\)
1011: u^\e\quad ; \quad u^\e_{\mid t=0} = a_0^\e(x)e^{i\varphi_0(x)/\e}\, .
1012: \end{equation}
1013: We recall the method introduced by E.~Grenier \cite{Grenier98},
1014: which is valid for smooth nonlinearities which are defocusing and
1015: cubic at the origin. Throughout this section, we assume the
1016: following:
1017: \begin{hyp}[Study of \eqref{eq:hs}]\label{hyp:hyper}
1018: We have $f'>0$. In addition,
1019: $\varphi_0\in H^\infty$, and there exists $a_0,a_1 \in H^\infty$
1020: such that
1021: \begin{equation*}
1022: a_0^\e = a_0 +\e a_1 +o(\e)\quad \text{in }H^s, \ \forall s\ge 0.
1023: \end{equation*}
1024: \end{hyp}
1025: Note that this assumption is closely akin to
1026: Assumption~\ref{hyp:surcrit}: nevertheless, we do not make any
1027: assumption on the
1028: momenta of $a_0$ and $a_1$, and the initial phase is
1029: bounded. We will see in Section~\ref{sec:surcrit} how to weaken this
1030: assumption.
1031: \subsection{Grenier's approach}
1032: \label{sec:hyper}
1033: The principle is somehow to perform the usual WKB analysis ``the other
1034: way round''. First, write the \emph{exact} solution as
1035: \begin{equation}\label{eq:ecrexacte}
1036: u^\e(t,x)= a^\e(t,x)e^{i\Phi^\e(t,x)/\e}\, ,
1037: \end{equation}
1038: where $\Phi^\e$ is real-valued. Then show that the ``amplitude'' $a^\e$
1039: and the ``phase'' $\Phi^\e$ have asymptotic expansions as $\e\to 0$:
1040: \begin{equation*}
1041: a^\e \sim a +\e a^{(1)}+\e^2 a^{(2)}+\ldots \quad ;\quad
1042: \Phi^\e \sim \phi +\e \phi^{(1)}+\e^2 \phi^{(2)}+\ldots
1043: \end{equation*}
1044: Introducing two unknown functions to solve one
1045: equation yields a degree of freedom. The historical approach
1046: \cite[Chap.~III]{LandauQ} consisted
1047: in writing
1048: \begin{align*}
1049: \d_t \Phi^\e +\frac{1}{2}\left|\nabla \Phi^\e\right|^2 + f\(
1050: |a^\e|^2\)= \e^2 \frac{\Delta a^\e}{2 a^\e}\quad &; \quad
1051: \Phi^\e\big|_{t=0}=\varphi_0\, ,\\
1052: \d_t a^\e +\nabla \Phi^\e \cdot \nabla a^\e +\frac{1}{2}a^\e
1053: \Delta \Phi^\e = 0\quad &;\quad
1054: a^\e\big|_{t=0}= a^\e_0\, .
1055: \end{align*}
1056: Of course, this choice is not adapted when the amplitude $a^\e$
1057: vanishes (see \cite{PGX93}), so it must be left out when $a^\e_0 \in
1058: L^2(\R^n)$ in general.
1059: The approach introduced by
1060: E.~Grenier consists in imposing:
1061: \begin{equation}\label{eq:systexact0}
1062: \begin{aligned}
1063: \d_t \Phi^\e +\frac{1}{2}\left|\nabla \Phi^\e\right|^2 + f\(
1064: |a^\e|^2\)= 0\quad &; \quad
1065: \Phi^\e\big|_{t=0}=\varphi_0\, ,\\
1066: \d_t a^\e +\nabla \Phi^\e \cdot \nabla a^\e +\frac{1}{2}a^\e
1067: \Delta \Phi^\e = i\frac{\e}{2}\Delta a^\e\quad & ;\quad
1068: a^\e\big|_{t=0}= a^\e_0\, .
1069: \end{aligned}
1070: \end{equation}
1071: Before recalling the results of \cite{Grenier98}, observe that if
1072: $a^\e$ and $\Phi^\e$ are bounded in some sufficiently small Sobolev
1073: spaces uniformly in $\e$, passing to the limit formally in
1074: \eqref{eq:systexact0} yields:
1075: \begin{equation}\label{eq:systlim}
1076: \begin{aligned}
1077: \d_t \phi +\frac{1}{2}\left|\nabla \phi\right|^2 + f\(
1078: |a|^2\)= 0\quad &; \quad
1079: \phi\big|_{t=0}=\varphi_0\, ,\\
1080: \d_t a +\nabla \phi \cdot \nabla a +\frac{1}{2}a
1081: \Delta \phi = 0\quad & ;\quad
1082: a\big|_{t=0}= a_0\, .
1083: \end{aligned}
1084: \end{equation}
1085: We see that when the nonlinearity is
1086: exactly cubic ($f(y)\equiv y$), $(\rho,v):=\( |a|^2,\nabla
1087: \phi\)$ solves the compressible, isentropic Euler equation
1088: \begin{equation}
1089: \label{eq:euler}
1090: \begin{aligned}
1091: \d_t v +v\cdot \nabla v + \nabla \rho= 0\quad &; \quad
1092: v\big|_{t=0}=\nabla \varphi_0\, ,\\
1093: \d_t \rho +\nabla\cdot (\rho v) = 0\quad & ;\quad
1094: \rho\big|_{t=0}= |a_0|^2\, .
1095: \end{aligned}
1096: \end{equation}
1097: From this point of view, the formulation
1098: \eqref{eq:systlim} is closely akin to the change of unknown function $\rho
1099: \to \sqrt\rho$ introduced in \cite{MUK86} (see also
1100: \cite{JYC90}) to study \eqref{eq:euler} when the initial density is
1101: compactly supported, a situation more or less similar to the present
1102: one. Note however that here, $a$ is complex-valued in general.
1103: \smallbreak
1104:
1105:
1106: Introducing the ``velocity'' $v^\e = \nabla \Phi^\e$,
1107: \eqref{eq:systexact0} yields
1108: \begin{equation}\label{eq:systexact}
1109: \begin{aligned}
1110: \d_t v^\e +v^\e \cdot \nabla v^\e + 2 f'\(
1111: |a^\e|^2\) \operatorname{Re}\(\overline{a^\e}\nabla a^\e\)=
1112: 0\quad &; \quad
1113: v^\e\big|_{t=0}=\nabla \phi_0\, ,\\
1114: \d_t a^\e +v^\e\cdot \nabla a^\e +\frac{1}{2}a^\e
1115: \nabla\cdot v^\e = i\frac{\e}{2}\Delta a^\e\quad & ;\quad
1116: a^\e\big|_{t=0}= a^\e_0\, .
1117: \end{aligned}
1118: \end{equation}
1119: Separate real and
1120: imaginary parts of $a^\e$, $a^\e =
1121: a_1^\e + ia_2^\e$. Then we have
1122: \begin{equation}
1123: \label{eq:systhyp}
1124: \d_t \bu^\e +\sum_{j=1}^n A_j(\bu^\e)\d_j \bu^\e
1125: = \frac{\e}{2} L
1126: \bu^\e\, ,
1127: \end{equation}
1128: \begin{equation*}
1129: \text{with}\quad \bu^\e = \left(
1130: \begin{array}[l]{c}
1131: a_1^\e \\
1132: a_2^\e \\
1133: v^\e_1 \\
1134: \vdots \\
1135: v^\e_n
1136: \end{array}
1137: \right)\quad , \quad L = \left(
1138: \begin{array}[l]{ccccc}
1139: 0 &-\Delta &0& \dots & 0 \\
1140: \Delta & 0 &0& \dots & 0 \\
1141: 0& 0 &&0_{n\times n}& \\
1142: \end{array}
1143: \right),
1144: \end{equation*}
1145: \begin{equation*}
1146: %\begin{aligned}
1147: \text{and}\quad A(\bu,\xi)=\sum_{j=1}^n A_j(\bu)\xi_j
1148: = \left(
1149: \begin{array}[l]{ccc}
1150: v\cdot \xi & 0& \frac{a_1 }{2}\,^{t}\xi \\
1151: 0 & v\cdot \xi & \frac{a_2}{2}\,^{t}\xi \\
1152: 2f' a_1 \, \xi
1153: &2f' a_2\, \xi & v\cdot \xi I_n
1154: \end{array}
1155: \right),
1156: %\end{aligned}
1157: \end{equation*}
1158: where $f'$ stands for $f'(|a_1|^2+|a_2|^2)$.
1159: The matrix $A(\bu,\xi)$ can be symmetrized by
1160: \begin{equation}\label{eq:symetriseur}
1161: S=\left(
1162: \begin{array}[l]{cc}
1163: I_2 & 0\\
1164: 0& \frac{1}{4f'}I_n
1165: \end{array}
1166: \right),
1167: \end{equation}
1168: which is symmetric and positive since $f'>0$. For an integer
1169: $s>2+n/2$, we bound $(S
1170: \d_{x}^{\alpha } \bu^\e
1171: , \d_{x}^{\alpha }\bu^\e )$
1172: where $\alpha $ is a multi index of length $\le s$, and
1173: $(\cdot ,\cdot)$ is the usual $L^{2}$ scalar product. We have
1174: \begin{equation*}
1175: \frac{d}{dt}\(S \d_{x}^{\alpha } \bu^\e ,
1176: \d_{x}^{\alpha } \bu^\e\)
1177: = \(\d_{t} S \d_{x}^{\alpha } \bu^\e , \partial
1178: _{x}^{\alpha } \bu^\e\)
1179: + 2 \( S \d_{t} \d_{x}^{\alpha } \bu^\e , \partial
1180: _{x}^{\alpha } \bu^\e\)
1181: \end{equation*}
1182: since $S$ is symmetric. For the first term, we consider the lower
1183: $n\times n$ block:
1184: \begin{equation*}
1185: \(\d_{t} S \d_{x}^{\alpha } \bu^\e ,
1186: \d_{x}^{\alpha } \bu^\e\)
1187: \le \left\|\frac{1}{f'}\d_t\(f'\( | a_1^\e|^2 +
1188: | a_2^\e|^2\)\)\right\|_{L^\infty}\(S \d_{x}^{\alpha } \bu^\e ,
1189: \d_{x}^{\alpha } \bu^\e\)\, .
1190: \end{equation*}
1191: So long as $\|\bu^\e\|_{L^\infty}\le 2\|a_0^\e\|_{L^\infty}$, we have:
1192: \begin{equation*}
1193: f'\( |a_1^\e|^2 +
1194: | a_2^\e|^2\) \ge \inf\left\{ f'(y)\ ;\ 0\le y\le
1195: 4\sup_{0<\e\le 1}\|a_0^\e\|_{L^\infty}^2\right\} =\delta_n>0\, ,
1196: \end{equation*}
1197: where $\delta_n$ is now fixed, since $f'$ is continuous with
1198: $f'>0$. We infer,
1199: \begin{equation*}
1200: \left\|\frac{1}{f'}\d_t\(f'\( |a_1^\e|^2 +
1201: | a_2^\e|^2\)\)\right\|_{L^\infty}\le C \|\bu^\e\|_{H^s}\, ,
1202: \end{equation*}
1203: where we used Sobolev embeddings and
1204: \eqref{eq:systhyp}.
1205: For the second term we use
1206: \begin{equation*}
1207: \( S \d_{t} \d_{x}^{\alpha } \bu^\e , \d_{x}^{\alpha } \bu^\e\)
1208: = \frac{\e}{2}\(S L( \d_{x}^{\alpha } \bu^\e ) , \d_{x}^{\alpha }
1209: \bu^\e\)
1210: -
1211: \Big( S \d_{x}^{\alpha } \Big(\sum _{j=1}^{n} A_j(\bu^\e) \d_{j}
1212: \bu^\e \Big) ,
1213: \d_{x}^{\alpha } \bu^\e\Big).
1214: \end{equation*}
1215: We notice that $SL$
1216: is a skew-symmetric second order operator, so the first term is zero.
1217: For the second term, use the symmetry of $SA_j(\bu^\e)$ and usual
1218: estimates on commutators to get finally:
1219: \begin{equation*}
1220: \frac{d}{dt} \sum _{|\alpha | \le s} \(S \d_{x}^{\alpha } \bu^\e,
1221: \d_{x}^{\alpha } \bu^\e\) \le C\(\left\|\bu^\e\right\|_{H^s}\)
1222: \sum _{|\alpha | \le s} \(S \d_{x}^{\alpha } \bu^\e,
1223: \d_{x}^{\alpha } \bu^\e\)\, ,
1224: \end{equation*}
1225: for $s > 2+d/2 $.
1226: Gronwall lemma along with a
1227: continuity argument yield:
1228: \begin{prop}[\cite{Grenier98}, Th.~1.1]\label{prop:p}
1229: Let Assumption~\ref{hyp:hyper} be satisfied. Let
1230: $s>2+n/2$. There exist $T_s>0$ independent of $\e\in ]0,1]$ and
1231: $u^\e = a^\e e^{i\Phi^\e/\e}$ solution to
1232: \eqref{eq:hs} on
1233: $[0,T_s]$. Moreover, $a^\e$ and
1234: $\Phi^\e$ are bounded
1235: in $L^\infty([0,T_s];H^s)$,
1236: uniformly in $\e\in ]0,1]$.
1237: \end{prop}
1238: The solution to \eqref{eq:systexact0} formally converges to the
1239: solution of \eqref{eq:systlim}. Under Assumption~\ref{hyp:hyper},
1240: \eqref{eq:systlim} has a unique solution $(a,\phi)\in
1241: L^\infty([0,T_*];H^m)^2$ for any $m>0$ for some $T_*>0$
1242: independent of $m$ (see e.g. \cite{AlinhacGerard,Majda}). We
1243: infer:
1244: \begin{prop}\label{prop:estprec}
1245: Let $s\in \N$. Then $T_s\ge T_*$, and there exists $C_s$ independent of
1246: $\e$ such that for
1247: every $0\le t\le T_*$,
1248: \begin{equation}\label{eq:tpetit}
1249: \| a^\e (t)- a(t)\|_{H^s}\le C_s \e\quad ;\quad
1250: \| \Phi^\e(t) - \phi(t)\|_{H^s} \le C_s \e t .
1251: \end{equation}
1252: \end{prop}
1253: \begin{proof}
1254: We keep the same notations as above,
1255: \eqref{eq:systhyp}. Denote by $\bv$
1256: the analog of $\bu^\e$
1257: corresponding to $(a,\phi)$. We have
1258: \begin{equation*}
1259: \d_t \(\bu^\e-\bv\) +\sum_{j=1}^n A_j(\bu^\e)\d_j \(\bu^\e-\bv\) +
1260: \sum_{j=1}^n \(A_j(\bu^\e)-A_j(\bv)\)\d_j \bv
1261: = \frac{\e}{2} L
1262: \bu^\e \, .
1263: \end{equation*}
1264: Keeping the symmetrizer $S$ corresponding to $\bu^\e$, we can do
1265: similar computations to the previous ones. Note that we
1266: know that $\bu^\e$ and $\bv$ are bounded in
1267: $L^\infty([0,\min(T_s,T_*)];H^s)$. Denote ${\bw}^\e
1268: =\bu^\e-\bv$. Writing $L
1269: \bu^\e = L {\bw}^\e + L\bv$, the term $L {\bw}^\e$ disappears from
1270: the energy estimate, and we get, for $s>2+n/2$:
1271: \begin{align*}
1272: \frac{d}{dt} \sum _{|\alpha | \le s} \(S \d_{x}^{\alpha } \bw^\e,
1273: \d_{x}^{\alpha } \bw^\e\) \le & \, C\( \|\bu^\e\|_{H^s},
1274: \|\bv\|_{H^{s+2}}\)
1275: \sum _{|\alpha | \le s} \(S \d_{x}^{\alpha } \bw^\e,
1276: \d_{x}^{\alpha } \bw^\e\)\\
1277: &+ \e \|\bv\|_{H^{s+2}} \| \bw^\e(t)\|_{H^s}\, .
1278: \end{align*}
1279: Gronwall lemma and a continuity argument show that
1280: $\bw^\e$ (hence $\bu^\e$) is defined on $[0,T_*]$. By
1281: Assumption~\ref{hyp:hyper}, $\|\bw^\e_{\mid t=0}\|_{H^s} = \O(\e)$,
1282: and we get:
1283: \begin{equation*}
1284: \left\| \bw^\e\right\|_{L^\infty([0,T_*];H^s)}=\O(\e ).
1285: \end{equation*}
1286: The estimate for the phase (and not only its gradient) then follows
1287: from the above estimate and the integration in time of
1288: \eqref{eq:systexact0}-\eqref{eq:systlim}.
1289: \end{proof}
1290: Proposition~\ref{prop:estprec} yields an approximation of $u^\e$
1291: \emph{for small times only}:
1292: \begin{align*}
1293: \big\| u^\e(t) - & a(t)e^{i\phi(t)/\e}\big\|_{L^2}=\left\|
1294: a^\e(t)e^{i\Phi^\e(t)/\e}- a(t)e^{i\phi(t)/\e}\right\|_{L^2}\\
1295: &=\O\(\left\|
1296: a^\e(t)- a(t)\right\|_{L^2}+\left\|
1297: e^{i\Phi^\e(t)/\e}- e^{i\phi(t)/\e}\right\|_{L^\infty }\left\|
1298: a(t)\right\|_{L^2} \)\\
1299: & = \O(\e)+\O(t).
1300: \end{align*}
1301: For times of order $\O(1)$, the corrector $a_1$ must be taken into
1302: account:
1303: \begin{prop}\label{prop:correc}
1304: Let Assumption~\ref{hyp:hyper} be satisfied. Define
1305: $(a^{(1)},\phi^{(1)})$ by
1306: \begin{equation*}
1307: \begin{aligned}
1308: \d_t \phi^{(1)} +\nabla \phi \cdot \nabla \phi^{(1)} +
1309: 2\operatorname{Re}\(\overline a a^{(1)}\)f'\( |a|^2\)&=0,\\
1310: \d_t a^{(1)} +\nabla\phi\cdot \nabla a^{(1)} + \nabla
1311: \phi^{(1)}\cdot \nabla a + \frac{1}{2} a^{(1)}\Delta \phi
1312: +\frac{1}{2}a\Delta \phi^{(1)} &= \frac{i}{2}\Delta a,\\
1313: \phi^{(1)}\big|_{t=0}=0\quad ; \quad a^{(1)}\big|_{t=0}=a_1.
1314: \end{aligned}
1315: \end{equation*}
1316: Then $a^{(1)},\phi^{(1)}\in
1317: L^\infty([0,T_*];H^s)$ for every $s\ge 0$, and
1318: \begin{equation*}
1319: \|a^\e - a - \e a^{(1)}\|_{L^\infty([0,T_*];H^s)}+ \|\Phi^\e - \phi - \e
1320: \phi^{(1)}\|_{L^\infty([0,T_*];H^s)} \le C_s\e^2,\quad \forall s\ge 0\, .
1321: \end{equation*}
1322: \end{prop}
1323: The proof is a straightforward consequence of the above analysis, and
1324: is given in \cite{Grenier98}. Despite
1325: the notations, it seems unadapted to consider $\phi^{(1)}$ as being
1326: part of the phase. Indeed, we infer from Proposition~\ref{prop:correc}
1327: that
1328: \begin{equation*}
1329: \left\|u^\e - a e^{i\phi^{(1)}}
1330: e^{i\phi/\e}\right\|_{L^\infty([0,T_*];L^2\cap L^\infty)}= \O(\e).
1331: \end{equation*}
1332: Relating this information to the WKB methods presented in
1333: the introduction, we would have:
1334: \begin{equation*}
1335: \a_0 = a e^{i\phi^{(1)}}.
1336: \end{equation*}
1337: Since $\phi^{(1)}$ depends on $a_1$ while $a$ does not, we retrieve
1338: the fact that in super-critical r\'egimes, the leading order amplitude
1339: in WKB methods depends on the initial first corrector $a_1$.
1340: \begin{rema}
1341: The term $e^{i\phi^{(1)}}$ does not appear in the Wigner measure
1342: of $a e^{i\phi^{(1)}} e^{i\phi/\e}$. Thus, from the point of view of
1343: Wigner measures, the asymptotic behavior of the exact solution is
1344: described by the Euler-type system \eqref{eq:systlim}.
1345: \end{rema}
1346:
1347: \begin{rema}[Introducing an isotropic harmonic potential]
1348: The above method makes it possible to consider the semi-classical of
1349: \eqref{eq:nlssemi} when $V(t,x)=\frac{1}{2}|x|^2$ is an isotropic harmonic
1350: potential, and Assumption~\ref{hyp:hyper} is satisfied. Let
1351: \begin{equation*}
1352: U^\e(t,x) =
1353: \frac{1}{(1+t^2)^{n/4}}e^{i\frac{t}{1+t^2}\frac{|x|^2}{2\e}}u^\e \(
1354: \arctan t , \frac{x}{\sqrt{1+t^2}}\)\, .
1355: \end{equation*}
1356: Then $U^\e$ solves:
1357: \begin{equation*}\left\{
1358: \begin{aligned}
1359: i\e \d_t U^\e + \frac{\e^2}{2}\Delta U^\e& =
1360: \frac{1}{1+t^2}f\(\(1+t^2\)^{n/2}|U^\e|^2\)U^\e\, , \\
1361: U^\e(0,x)&= a_0^\e(x)e^{i\varphi_0(x)/\e} .
1362: \end{aligned}\right.
1363: \end{equation*}
1364: We can then proceed as above. The only difference is the presence of
1365: time in the nonlinearity, which changes very little the analysis.
1366: \end{rema}
1367:
1368: \begin{rema}[Momenta]\label{rema:moment}
1369: If in Assumption~\ref{hyp:hyper}, we replace $H^s$ with
1370: \begin{equation*}
1371: \Sigma^s =H^s \cap \F (H^s) =\left\{ w \in L^2\ ; \
1372: (1-\Delta)^{k/2}\< x\>^{s-k}w\in L^2, \ 0\le k\le s\right\},
1373: \end{equation*}
1374: then the above analysis can be repeated in $\Sigma^s$. The main
1375: difference is due to the commutations of the
1376: powers of $x$ with the differential operators; it is easy to check
1377: that they introduce semilinear terms, which can be treated as
1378: source terms when applying Gronwall lemma.
1379: \end{rema}
1380:
1381:
1382:
1383:
1384:
1385:
1386:
1387:
1388:
1389:
1390:
1391:
1392:
1393: \subsection{Remarks about some conserved quantities}
1394: \label{sec:conserved}
1395: Consider the case of the cubic, defocusing Schr\"odinger equation:
1396: $f(y)\equiv y$.
1397: Recall three important evolution laws for \eqref{eq:nlssemi}:
1398: \begin{align*}
1399: \text{Mass: }& \frac{d}{dt}\|u^\e(t)\|_{L^2}=0\, .\\
1400: \text{Energy: }& \frac{d}{dt}\(\|\e\nabla_x u^\e\|_{L^2}^2
1401: + \|u^\e\|_{L^4}^4 \)=0\, .\\
1402: \text{Momentum: }&\frac{d}{dt}\operatorname{Im}\int \overline {u^\e}(t,x)
1403: \e \nabla_x u^\e(t,x)dx =0\, .\\
1404: \text{Pseudo-conformal law: }& \frac{d}{dt}\(\|J^\e(t)
1405: u^\e\|_{L^2}^2
1406: + t^2\|u^\e\|_{L^{4}}^{4}
1407: \)=t (2-n)\|u^\e\|_{L^{4}}^{4} ,
1408: \end{align*}
1409: where $J^\e(t) = x + i\e t\nabla_x$. These evolutions are deduced
1410: from the usual ones ($\e =1$, see e.g. \cite{CazCourant,Sulem}) via
1411: the scaling
1412: $\psi(t,x)= u(\e t,\e x)$. Using \eqref{eq:ecrexacte} and passing to
1413: the limit formally in the above formulae yields:
1414: \begin{align*}
1415: \frac{d}{dt}\|a(t)\|_{L^2}&=0\, .\\
1416: \frac{d}{dt}\int \( |a(t,x)|^2 |\nabla \phi(t,x)|^2 +
1417: |a(t,x)|^4\)dx &=0\, .\\
1418: \frac{d}{dt}\int |a(t,x)|^2 \nabla \phi(t,x)dx &=0\, .\\
1419: \frac{d}{dt}\int \( \left| \(x -t \nabla
1420: \phi(t,x)\)a(t,x)\right|^2 +t^2 |a(t,x)|^4
1421: \)&dx=\\
1422: =(2-n)t &\int |a(t,x)|^4dx\, .
1423: \end{align*}
1424: Note that we also
1425: have the conservation (\cite{CN}):
1426: \begin{equation*}
1427: \frac{d}{dt}\operatorname{Re}\int \overline{u^\e}(t,x)
1428: J^\e(t)u^\e(t,x) dx=0 \, ,
1429: \end{equation*}
1430: which yields:
1431: \begin{equation*}
1432: \frac{d}{dt}\int \(\(x -t \nabla
1433: \phi(t,x)\)|a(t,x)|^2 \)dx =0\, .
1434: \end{equation*}
1435: All these expressions involve only $(|a|^2, \nabla \phi)$, that
1436: is, the solution of \eqref{eq:euler}. We thus
1437: retrieve formally some evolution laws
1438: for the Euler equation.
1439:
1440:
1441:
1442:
1443:
1444: %%%%%%%%%%%%%%%%
1445:
1446:
1447:
1448:
1449:
1450:
1451: \section{Introducing subquadratic potential and initial phase}
1452: \label{sec:surcrit}
1453: In this section, we prove Theorem~\ref{theo:BKWV}.
1454: First, we point out that the uniqueness for $u^\e$ in $C([0,T_*];H^s)$
1455: is straightforward for $s>n/2$. We thus have to prove that there
1456: exists such a solution, and that it is smooth.
1457: \smallbreak
1458:
1459: As suggested by
1460: the statements of Theorem~\ref{theo:BKWV}, the idea consists in
1461: resuming Grenier's method, and in writing the phase $\Phi^\e$ as
1462: \begin{equation*}
1463: \Phi^\e = \eik +\phi^\e.
1464: \end{equation*}
1465: We take $\phi^\e$ as a new unknown function. Recall that the
1466: system~\eqref{eq:systexact0} reads, with the present notations:
1467: \begin{align*}
1468: \d_t \Phi^\e +\frac{1}{2}\left|\nabla \Phi^\e\right|^2 + V+ f\(
1469: |a^\e|^2\)= 0\quad &; \quad
1470: \Phi^\e\big|_{t=0}=\phi_0\, ,\\
1471: \d_t a^\e +\nabla \Phi^\e \cdot \nabla a^\e +\frac{1}{2}a^\e
1472: \Delta \Phi^\e = i\frac{\e}{2}\Delta a^\e\quad & ;\quad
1473: a^\e\big|_{t=0}= a^\e_0\, .
1474: \end{align*}
1475: This system becomes, in terms of $\phi^\e$, and given \eqref{eq:eik}:
1476: \begin{equation}\label{eq:modif0}
1477: \begin{aligned}
1478: \d_t \phi^\e +\frac{1}{2}\left|\nabla \phi^\e\right|^2 +\nabla
1479: \eik\cdot \nabla \phi^\e+ f\(
1480: |a^\e|^2\)&= 0,\\
1481: \d_t a^\e +\nabla \phi^\e \cdot \nabla a^\e +\nabla \eik \cdot \nabla
1482: a^\e +\frac{1}{2}a^\e
1483: \Delta \phi^\e +\frac{1}{2}a^\e
1484: \Delta \eik &= i\frac{\e}{2}\Delta a^\e,\\
1485: \phi^\e\big|_{t=0}=0\quad ;\quad a^\e\big|_{t=0}&= a^\e_0\, .
1486: \end{aligned}
1487: \end{equation}
1488: Like in Section~\ref{sec:hyper}, we
1489: work with $v^\e =\nabla \phi^\e$ instead of $\phi^\e$, to begin
1490: with. The new terms are the factors where $\eik$ is present. The point
1491: is to check that they are semilinear perturbations, which can be treated as
1492: source terms in view of Gronwall lemma. Again, separate real and
1493: imaginary parts of $a^\e$, $a^\e =
1494: a_1^\e + ia_2^\e$, and introduce:
1495: \begin{equation*}
1496: \bu^\e = \left(
1497: \begin{array}[l]{c}
1498: a_1^\e \\
1499: a_2^\e \\
1500: v^\e_1 \\
1501: \vdots \\
1502: v^\e_n
1503: \end{array}
1504: \right)\quad , \quad L = \left(
1505: \begin{array}[l]{ccccc}
1506: 0 &-\Delta &0& \dots & 0 \\
1507: \Delta & 0 &0& \dots & 0 \\
1508: 0& 0 &&0_{n\times n}& \\
1509: \end{array}
1510: \right),
1511: \end{equation*}
1512: \begin{equation*}
1513: %\begin{aligned}
1514: \text{and}\quad A(\bu,\xi)=\sum_{j=1}^n A_j(\bu)\xi_j
1515: = \left(
1516: \begin{array}[l]{ccc}
1517: v\cdot \xi & 0& \frac{a_1 }{2}\,^{t}\xi \\
1518: 0 & v\cdot \xi & \frac{a_2}{2}\,^{t}\xi \\
1519: 2f' a_1 \, \xi
1520: &2f' a_2\, \xi & v\cdot \xi I_n
1521: \end{array}
1522: \right),
1523: %\end{aligned}
1524: \end{equation*}
1525: where $f'$ stands for $f'(|a_1|^2+|a_2|^2)$. Instead of
1526: \eqref{eq:systhyp}, we now have a system of the form
1527: \begin{equation}
1528: \label{eq:systhypV}
1529: \d_t \bu^\e +\sum_{j=1}^n A_j(\bu^\e)\d_j \bu^\e
1530: +\sum_{j=1}^n B_j(\nabla \eik)\d_j \bu^\e + M\(\nabla^2
1531: \eik\)\bu^\e= \frac{\e}{2} L
1532: \bu^\e\, ,
1533: \end{equation}
1534: where the matrices $B_j$ depend linearly on their argument, and the matrix
1535: $M$ is smooth, locally bounded. The quasilinear part of
1536: \eqref{eq:systhypV} is the same as in
1537: Section~\ref{sec:hyper}, and involves the matrices $A_j$. In
1538: particular, we keep the same symmetrizer $S$ given by
1539: \eqref{eq:symetriseur}. The matrices
1540: $B_j$ have a semilinear contribution, as we see below. The term
1541: corresponding to the matrix $M$ can obviously be considered as a
1542: source term, since $\eik$ is subquadratic.
1543: \smallbreak
1544:
1545: Let $s$ be an integer,
1546: $s>2+n/2$, and let $\alpha $ be a multi index of length $\le s$. We
1547: have:
1548: \begin{equation*}
1549: \frac{d}{dt}\(S \d_{x}^{\alpha } \bu^\e ,
1550: \d_{x}^{\alpha } \bu^\e\)
1551: = \(\d_{t} S \d_{x}^{\alpha } \bu^\e , \partial
1552: _{x}^{\alpha } \bu^\e\)
1553: + 2 \( S \d_{t} \d_{x}^{\alpha } \bu^\e , \partial
1554: _{x}^{\alpha } \bu^\e\)
1555: \end{equation*}
1556: since $S$ is symmetric. For the first term, we consider the lower
1557: $n\times n$ block:
1558: \begin{equation}\label{eq:estdtS}
1559: \(\d_{t} S \d_{x}^{\alpha } \bu^\e ,
1560: \d_{x}^{\alpha } \bu^\e\)
1561: \le \left\|\frac{1}{f'}\d_t\(f'\( | a_1^\e|^2 +
1562: | a_2^\e|^2\)\)\right\|_{L^\infty}\(S \d_{x}^{\alpha } \bu^\e ,
1563: \d_{x}^{\alpha } \bu^\e\)\, .
1564: \end{equation}
1565: We consider times not larger than $T$ given by Lemma~\ref{lem:hj}, so
1566: that the function $\eik$ remains smooth and subquadratic.
1567: So long as $\|\bu^\e\|_{L^\infty}\le 2\|a_0^\e\|_{L^\infty}$, we have:
1568: \begin{equation*}
1569: f'\( |a_1^\e|^2 +
1570: | a_2^\e|^2\) \ge \inf\left\{ f'(y)\ ;\ 0\le y\le
1571: 4\sup_{0<\e\le 1}\|a_0^\e\|_{L^\infty}^2\right\} =\delta_n>0\, .
1572: \end{equation*}
1573: We infer,
1574: \begin{equation}\label{eq:dtS}
1575: \left\|\frac{1}{f'}\d_t\(f'\( |a_1^\e|^2 +
1576: | a_2^\e|^2\)\)\right\|_{L^\infty}\le C \(\|\bu^\e\|_{H^s}+
1577: \|x\bu^\e\|_{H^{s-1}}\)\, ,
1578: \end{equation}
1579: for some locally bounded map $C(\cdot)$. We used Sobolev embeddings,
1580: \eqref{eq:systhypV} and Lemma~\ref{lem:hj}: the terms $B_j$ are
1581: sublinear in $x$, hence the norm $\|x\bu^\e\|_{H^{s-1}}$ which we did
1582: not consider in Section~\ref{sec:hyper}. We emphasize that this
1583: estimate explains why we assume $s>2+n/2$, and not only $s>1+n/2$:
1584: we control $\d_t \bu^\e$ in $L^\infty$ using \eqref{eq:systhypV},
1585: so we need to estimate $L \bu^\e$ in $L^\infty$. For all the other
1586: terms, $s>1+n/2$ would suffice. This also explains why we wrote
1587: $\|x\bu^\e\|_{H^{s-1}}$ and not $\|x\bu^\e\|_{H^{s}}$.
1588: For the second term we use
1589: \begin{align*}
1590: \( S \d_{t} \d_{x}^{\alpha } \bu^\e , \d_{x}^{\alpha } \bu^\e\)
1591: =& \frac{\e}{2}\(S L( \d_{x}^{\alpha } \bu^\e ) , \d_{x}^{\alpha }
1592: \bu^\e\)
1593: -
1594: \Big( S \d_{x}^{\alpha } \Big(\sum _{j=1}^{n} A_j(\bu^\e) \d_{j}
1595: \bu^\e \Big) ,
1596: \d_{x}^{\alpha } \bu^\e\Big)\\
1597: - \Big( S \d_{x}^{\alpha } \Big(\sum _{j=1}^{n} B_j&(\nabla\eik) \d_{j}
1598: \bu^\e \Big) ,
1599: \d_{x}^{\alpha } \bu^\e\Big)- \Big( S \d_{x}^{\alpha }
1600: \Big(M(\nabla^2 \eik)
1601: \bu^\e \Big) ,
1602: \d_{x}^{\alpha } \bu^\e\Big).
1603: \end{align*}
1604: The first two terms of the right hand side are handled in the same way
1605: as in Section~\ref{sec:hyper}: the first one is zero, and the second
1606: can be estimated by:
1607: \begin{equation}\label{eq:estQL}
1608: \Big( S \d_{x}^{\alpha } \Big(\sum _{j=1}^{n} A_j(\bu^\e) \d_{j}
1609: \bu^\e \Big) ,
1610: \d_{x}^{\alpha } \bu^\e\Big)\le C\(\left\|\bu^\e\right\|_{H^s}\)
1611: \sum _{|\alpha | \le s} \(S \d_{x}^{\alpha } \bu^\e,
1612: \d_{x}^{\alpha } \bu^\e\),
1613: \end{equation}
1614: where we keep the convention that $C(\cdot)$ is a locally bounded
1615: map. Let us briefly explain this quasilinear estimate. First, write
1616: \begin{equation*}
1617: \begin{aligned}
1618: \( S \partial _{x}^{\alpha } \( A_j(\bu^\e) \d_j
1619: \bu^\e \) ,
1620: \partial _{x}^{\alpha } \bu^\e\)&
1621: =
1622: \(S A_j(\bu^\e) \d_j \partial _{x}^{\alpha
1623: } \bu^\e,
1624: \partial _{x}^{\alpha } \bu^\e\) \\
1625: +& \(S \( \partial _{x}^{\alpha } (
1626: A_j(\bu^\e) \d_j \bu^\e ) -
1627: A_j(\bu^\e) \d_j \partial _{x}^{\alpha }
1628: \bu^\e \),
1629: \partial _{x}^{\alpha } \bu^\e\).
1630: \end{aligned}
1631: \end{equation*}
1632: By symmetry of $SA_j(\bu^\e)$,
1633: \begin{equation*}
1634: \begin{aligned}
1635: \(S A_j(\bu^\e) \d_j \partial_{x}^{\alpha } \bu^\e,
1636: \partial _{x}^{\alpha } \bu^\e\)
1637: = & - \(\d_j (S A_j(\bu^\e)) \d_{x}^{\alpha } \bu^\e,
1638: \d_{x}^{\alpha } \bu^\e\)\\
1639: &- \(S A_j(\bu^\e) \d_j \partial _{x}^{\alpha } \bu^\e,
1640: \partial _{x}^{\alpha } \bu^\e\) .
1641: \end{aligned}
1642: \end{equation*}
1643: We infer:
1644: \begin{align*}
1645: \left| \(S A_j(\bu^\e) \d_j \partial _{x}^{\alpha } \bu^\e,
1646: \d_{x}^{\alpha } \bu^\e\) \right|&
1647: \le
1648: \left\| \d_j\(S A_j (\bu^\e)\)
1649: \right\|_{L^{\infty }}
1650: \left\|\d_{x}^{\alpha } \bu^\e \right\|_{L^{2}}^{2}\\
1651: &\le
1652: C\( \left\| \bu^\e \right\|_{L^{\infty }}\) \left\| \nabla _{x} \bu^\e
1653: \right\|_{L^{\infty }}
1654: \left\|\d_{x}^{\alpha } \bu^\e \right\|_{L^{2}}^{2} .
1655: \end{align*}
1656: The usual estimates on commutators (see e.g. \cite{Majda}) lead to
1657: \begin{equation*}
1658: \left|\(S \( \partial_{x}^{\alpha } \(
1659: A_j(\bu^\e) \d_j \bu^\e \) -
1660: A_j(\bu^\e) \d_j \partial_{x}^{\alpha } \bu^\e \),
1661: \partial _{x}^{\alpha } \bu^\e\) \right|
1662: \le C\(\left\|\bu^\e\right\|_{H^s}\) \left\|\bu^\e\right\|_{H^s}^{2},
1663: \end{equation*}
1664: and \eqref{eq:estQL} follows, since we consider times where
1665: $S^{-1}$ is bounded.
1666: \smallbreak
1667:
1668:
1669: For the third term of $\( S \d_{t} \d_{x}^{\alpha } \bu^\e ,
1670: \d_{x}^{\alpha } \bu^\e\)$, write:
1671: \begin{align*}
1672: \( S \d_{x}^{\alpha } \(B_j(\nabla\eik) \d_{j}
1673: \bu^\e \) ,
1674: \d_{x}^{\alpha } \bu^\e\) =& \int S B_j(\nabla\eik) \d_{j}\d_{x}^{\alpha }
1675: \bu^\e\d_{x}^{\alpha } \bu^\e dx\\
1676: &+\int S \left[\d_{x}^{\alpha } ,B_j(\nabla\eik) \d_{j}\right]
1677: \bu^\e\d_{x}^{\alpha } \bu^\e dx.
1678: \end{align*}
1679: For the first term of the right hand side, an integration by parts
1680: yields:
1681: \begin{align}
1682: \left|\int S B_j(\nabla\eik) \d_{j}\d_{x}^{\alpha }
1683: \bu^\e\d_{x}^{\alpha } \bu^\e dx\right| &\le \left\|
1684: \d_j\( S B_j(\nabla\eik)
1685: \)\right\|_{L^\infty}\|\bu^\e\|_{H^s}^2 \label{eq:16h47} \\
1686: &\le C(\|a^\e \|_{L^\infty})\left\|\<x\> a^\e \nabla a^\e
1687: \right\|_{L^\infty}\|\bu^\e\|_{H^s}^2\notag \\
1688: \le C(\|\bu^\e\|_{L^\infty})
1689: \Big(\left\|\bu^\e\right\|_{L^\infty}&+\left\|x\bu^\e\right\|_{L^\infty}
1690: + \left\|\nabla\bu^\e\right\|_{L^\infty}\Big)^2\|\bu^\e\|_{H^s}^2,\notag
1691: \end{align}
1692: where we have used Lemma~\ref{lem:hj}. Again
1693: from Lemma~\ref{lem:hj}, the commutator
1694: \begin{equation*}
1695: \left[\d_{x}^{\alpha } ,B_j(\nabla\eik) \d_{j}\right]
1696: \end{equation*}
1697: is a differential operator of degree $\le s$, with bounded
1698: coefficients. We infer:
1699: \begin{align*}
1700: \Big| \Big( S \d_{x}^{\alpha } \Big(\sum _{j=1}^{n} B_j(\nabla\eik) \d_{j}
1701: \bu^\e \Big) ,
1702: \d_{x}^{\alpha } \bu^\e\Big)\Big|
1703: \le C(\|\bu^\e \|_{H^s} +\|x\bu^\e \|_{H^{s-1}} )
1704: \|\bu^\e\|_{H^s}^2.
1705: \end{align*}
1706: We have obviously
1707: \begin{align*}
1708: \Big| \Big( S \d_{x}^{\alpha } \Big(M(\nabla^2\eik)
1709: \bu^\e \Big) ,
1710: \d_{x}^{\alpha } \bu^\e\Big)\Big| \le C(\|\bu^\e \|_{L^\infty})
1711: \|\bu^\e\|_{H^s}^2.
1712: \end{align*}
1713: This yields:
1714: \begin{equation}\label{eq:10h45}
1715: \frac{d}{dt}\(S \d_{x}^{\alpha } \bu^\e ,
1716: \d_{x}^{\alpha } \bu^\e\) \le C
1717: \(\left\|\bu^\e\right\|_{H^s}+\left\|x\bu^\e\right\|_{H^{s-1}}
1718: \)\|\bu^\e\|_{H^s}^2,
1719: \end{equation}
1720: where the map $C(\cdot)$ is locally bounded.
1721: We now have to
1722: bound $x\bu^\e$ in $H^{s-1}$ to close our family of
1723: estimates: we consider
1724: \begin{equation*}
1725: \frac{d}{dt}\(S \d_{x}^{\beta } (x_k\bu^\e) ,
1726: \d_{x}^{\beta } (x_k\bu^\e)\);\quad 1\le k\le n, \ |\beta|\le s-1.
1727: \end{equation*}
1728: We can proceed as above, replacing $\bu^\e$ with $x_k\bu^\e$:
1729: $x_k\bu^\e$ solves almost the same equation as $\bu^\e$, and we must
1730: control some commutators.
1731: \begin{align*}
1732: \d_t (x_k\bu^\e) +\sum_{j=1}^n A_j(\bu^\e)\d_j &(x_k \bu^\e)
1733: +\sum_{j=1}^n B_j(\nabla \eik)\d_j (x_k \bu^\e)+ \\
1734: + M\(\nabla^2
1735: \eik\)x_k\bu^\e
1736: = \frac{\e}{2} L
1737: (x_k \bu^\e)
1738: &+ A_k (\bu^\e) \bu^\e + B_k(\nabla \eik)\bu^\e +
1739: \frac{\e}{2}[x_k,L]\bu^\e .
1740: \end{align*}
1741: The term $A_k (\bu^\e) \bu^\e$ is harmless. The term $B_k(\nabla
1742: \eik)\bu^\e$ is controlled by $\<x\> \bu^\e$, since $\eik$ is
1743: subquadratic: this is a (semi)linear perturbation. Finally,
1744: \begin{equation*}
1745: [x_k,L] = \left(
1746: \begin{array}[l]{ccccc}
1747: 0 &2\d_k &0& \dots & 0 \\
1748: -2\d_k & 0 &0& \dots & 0 \\
1749: 0& 0 &&0_{n\times n}& \\
1750: \end{array}
1751: \right).
1752: \end{equation*}
1753: Now we only have to notice that estimating $x_k \bu^\e$ does not
1754: involve extra regularity or extra momenta. In the above computations,
1755: the first time we needed to consider momenta was for \eqref{eq:dtS}:
1756: we need exactly the same estimate now, since it is due to the
1757: symmetrizer, which remains the same. The same remark is valid for
1758: \eqref{eq:16h47}. For $\beta$ a multi index of length $\le s-1$, we
1759: find:
1760: \begin{equation}\label{eq:10h46}
1761: \begin{aligned}
1762: \frac{d}{dt}&\(S \d_{x}^{\beta } (x_k\bu^\e ),
1763: \d_{x}^{\beta } (x_k \bu^\e)\) \le \\
1764: &\le C \(\left\|\bu^\e\right\|_{H^s}+\left\|x\bu^\e\right\|_{H^{s-1}}
1765: \)\(\|\bu^\e\|_{H^s}^2+ \|x \bu^\e\|_{H^{s-1}}^2\).
1766: \end{aligned}
1767: \end{equation}
1768: The term $\|\bu^\e\|_{H^s}^2$ is due to the commutator $[x_k,L]$:
1769: \begin{align*}
1770: \left|\( S \d^\beta_x \( [x_k,L]\bu^\e\),\d^\beta_x\(x_k
1771: \bu^\e\)\)\right| & =
1772: \left|\( \d^\beta_x \( [x_k,L]\bu^\e\),\d^\beta_x\(x_k
1773: \bu^\e\)\)\right|\\
1774: &\le \|\bu^\e\|_{H^s}\|x \bu^\e\|_{H^{s-1}}.
1775: \end{align*}
1776: Summing over the inequalities \eqref{eq:10h45} and \eqref{eq:10h46}
1777: yields a closed set of estimates, from which we infer the analogue of
1778: Proposition~\ref{prop:p}; note that the time $T_s$ is not larger
1779: than $T$ by construction, and may be strictly smaller than $T$, due to
1780: possible shocks for \eqref{eq:systhypV}. We also
1781: mention the fact that
1782: the above analysis gives $v^\e = \nabla \phi^\e \in C([0,T_s];H^s)$,
1783: with $x \nabla \phi^\e \in C([0,T_s];H^{s-1})$: back to \eqref{eq:modif0},
1784: this shows that $ \phi^\e \in C([0,T_*];H^{s-1})$, but that we cannot
1785: claim that $ x\phi^\e \in C([0,T_*];H^{s-1})$.
1786: \smallbreak
1787:
1788: Passing to the limit $\e \to 0$ in \eqref{eq:modif0}, it is natural to
1789: introduce the system:
1790: \begin{equation}\label{eq:11h09}
1791: \begin{aligned}
1792: \d_t \phi +\frac{1}{2}\left|\nabla \phi\right|^2 +\nabla
1793: \eik\cdot \nabla \phi+ f\(
1794: |a|^2\)&= 0,\\
1795: \d_t a +\nabla \phi \cdot \nabla a +\nabla \eik \cdot \nabla
1796: a +\frac{1}{2}a
1797: \Delta \phi +\frac{1}{2}a
1798: \Delta \eik &= 0,\\
1799: \phi\big|_{t=0}=0\quad ;\quad a\big|_{t=0}&= a_0\, .
1800: \end{aligned}
1801: \end{equation}
1802: The above analysis shows that this system has a unique solution in
1803: $H^s$ with the first momentum in $H^{s-1}$,
1804: locally in time for $t\in [0,T_*]$, for some $T_*\in ]0,T]$; $T_*$ is
1805: independent of $s$, from the usual continuation
1806: principle, explained for instance in \cite[Section~2.2]{Majda}. We
1807: easily obtain the analogue of
1808: Proposition~\ref{prop:estprec}: mimicking the above computations, we
1809: can estimate the error $(a^\e -a,\phi^\e-\phi)$ in $H^s$, by first
1810: estimating $(a^\e -a,\nabla\phi^\e-\nabla\phi)$ \emph{and its first
1811: momentum} in $H^s$. We deduce that $T_s\ge T_*$, and
1812: \eqref{eq:BKWVtpetit} follows.
1813: \smallbreak
1814:
1815: The end of Theorem~\ref{theo:BKWV} can then be proved as in
1816: \cite{Grenier98}: from the above analysis, the functions
1817: \begin{equation*}
1818: \frac{a^\e -a}{\e}\virgp\ \frac{\nabla \phi^\e -\nabla\phi}{\e}
1819: \end{equation*}
1820: and their first momentum, are bounded in $H^s$ for every $s\ge 0$. A
1821: subsequence converges to the linearization of
1822: \eqref{eq:systhypV}, yielding a pair $(a^{(1)},\phi^{(1)})$. By
1823: uniqueness for the limit system, the whole sequence is convergent, and
1824: the analogue of Proposition~\ref{prop:correc} follows. This completes
1825: the proof of Theorem~\ref{theo:BKWV}.
1826:
1827:
1828:
1829: %%%%%%%%%%%%%%%
1830:
1831:
1832: \section{Extension to the case $0<\kappa <1$}
1833: \label{sec:kappaint}
1834:
1835: When $0<\kappa <1$, we propose an analysis which can be considered as
1836: a generalization of the study led in
1837: Section~\ref{sec:surcrit}. Throughout this paragraph, we suppose that
1838: Assumptions~\ref{hyp:geom}-\ref{hyp:surcrit} are satisfied. Again,
1839: we write the exact solution as
1840: \begin{equation*}
1841: u^\e = a^\e e^{i\Phi^\e/\e},\quad \text{with }\Phi^\e=\eik +\phi^\e.
1842: \end{equation*}
1843: The unknown function is the pair $(a^\e,\phi^\e)$. We have two unknown
1844: functions to solve a single equation, \eqref{eq:nlssemi}. We can
1845: choose how to balance the terms: we resume the approach followed when
1846: $\kappa=0$. Note that this approach would also be efficient for the
1847: case $\kappa \ge 1$, with the serious drawback that we still assume
1848: $f'>0$, an assumption proven to be unnecessary when $\kappa\ge 1$ (see
1849: Section~\ref{sec:sub}). We impose:
1850: \begin{equation}\label{eq:modifint}
1851: \begin{aligned}
1852: \d_t \phi^\e +\frac{1}{2}\left|\nabla \phi^\e\right|^2 +\nabla
1853: \eik\cdot \nabla \phi^\e+ \e^{\kappa}f\(
1854: |a^\e|^2\)&= 0,\\
1855: \d_t a^\e +\nabla \phi^\e \cdot \nabla a^\e +\nabla \eik \cdot \nabla
1856: a^\e +\frac{1}{2}a^\e
1857: \Delta \phi^\e +\frac{1}{2}a^\e
1858: \Delta \eik &= i\frac{\e}{2}\Delta a^\e,\\
1859: \phi^\e\big|_{t=0}=0\quad ;\quad a^\e\big|_{t=0}&= a^\e_0\, .
1860: \end{aligned}
1861: \end{equation}
1862: This is the same system as \eqref{eq:modif0}, with only $f$ replaced
1863: by $\e^{\kappa}f$. Mimicking the analysis of
1864: Section~\ref{sec:surcrit}, we work with the unknown $\bu^\e$ given by
1865: the same definition: it solves the system \eqref{eq:systhypV}, where
1866: only the matrices $A_j$ have changed, and now depend on $\e$. The
1867: symmetrizer is the same as before, with $f'$ replaced by $\e^\kappa
1868: f'$: the matrix $S=S^\e$ is not bounded as $\e \to 0$, but its inverse
1869: is. We see that \eqref{eq:estdtS} and \eqref{eq:dtS}
1870: still hold, independent of $\kappa$. We claim that inequalities
1871: similar to \eqref{eq:10h45} and \eqref{eq:10h46} hold:
1872: \begin{align*}
1873: \frac{d}{dt}\(S^\e \d_{x}^{\alpha } \bu^\e ,
1874: \d_{x}^{\alpha } \bu^\e\) \le C
1875: \(\left\|\bu^\e\right\|_{H^s}+\left\|x\bu^\e\right\|_{H^{s-1}}
1876: \)&\sum_{|\gamma|\le s}\(S^\e \d_{x}^{\gamma } \bu^\e ,
1877: \d_{x}^{\gamma } \bu^\e\),\\
1878: \frac{d}{dt}\(S^\e \d_{x}^{\beta } (x_k\bu^\e ),
1879: \d_{x}^{\beta } (x_k \bu^\e)\)
1880: \le C \(\left\|\bu^\e\right\|_{H^s}+\left\|x\bu^\e\right\|_{H^{s-1}}
1881: \)&\Big( \sum_{|\gamma|\le s}\(S^\e \d_{x}^{\gamma } \bu^\e ,
1882: \d_{x}^{\gamma } \bu^\e\)\\
1883: + \sum_{{|\gamma|\le s-1}\atop{1\le j\le
1884: n}}(S^\e \d_{x}^{\gamma
1885: }& (x_j\bu^\e) ,
1886: \d_{x}^{\gamma } (x_j\bu^\e))\Big),
1887: \end{align*}
1888: where the map $C(\cdot)$ is locally bounded, \emph{independent of
1889: }$\e\in]0,1]$. The fact that such estimates remain valid, with this
1890: dependence upon $\e$, stems essentially from the following reasons:
1891: \begin{itemize}
1892: \item The matrices $S^\e$ and $B_j$ are diagonal.
1893: \item The matrix $M$ is block diagonal: the blocks correspond to the
1894: presence/absence of $\e$ in $S^\e$.
1895: \item The matrices $S^\e A_j^\e$ are independent of $\e\in]0,1]$.
1896: \item The inverse of $S^\e$ is uniformly bounded on compacts, as $\e
1897: \to 0$.
1898: \end{itemize}
1899: A continuity argument and Gronwall lemma then imply the analogue of
1900: Proposition~\ref{prop:p}: $\bu^\e$ exists locally in time, with
1901: $H^s$-norm uniformly bounded as $\e\to 0$. Note that since
1902: $\phi^\e\big|_{t=0}=0$, we have:
1903: \begin{equation*}
1904: \(S^\e \d_{x}^{\alpha } \bu^\e ,
1905: \d_{x}^{\alpha } \bu^\e\)\big|_{t=0} = \O(1),
1906: \end{equation*}
1907: and we infer more precisely:
1908: \begin{equation*}
1909: \| a^\e \|_{L^\infty ([0,T_*];H^s)} =\O(1) \quad ; \quad
1910: \left\| \nabla \phi^\e \right\|_{L^\infty ([0,T_*];H^s)} =\O\(\e^\kappa\).
1911: \end{equation*}
1912: It seems natural to change unknown functions, and work with
1913: $\widetilde \phi^\e = \e^{-\kappa}\phi^\e$ instead of $\phi^\e$. With
1914: this, we somehow correct the shift in the cascade of equations caused by
1915: the factor $\e^\kappa$ in front of the nonlinearity. Then
1916: \eqref{eq:modifint} becomes:
1917: \begin{equation}\label{eq:modifinttilde}
1918: \begin{aligned}
1919: \d_t \widetilde \phi^\e +\frac{\e^\kappa}{2}\left|\nabla \widetilde
1920: \phi^\e\right|^2 +\nabla
1921: \eik\cdot \nabla \widetilde \phi^\e+ f\(
1922: |a^\e|^2\)&= 0,\\
1923: \d_t a^\e +\e^\kappa\nabla \widetilde \phi^\e \cdot \nabla a^\e +\nabla
1924: \eik \cdot \nabla
1925: a^\e +\frac{\e^\kappa}{2}a^\e
1926: \Delta \widetilde \phi^\e +\frac{1}{2}a^\e
1927: \Delta \eik &= i\frac{\e}{2}\Delta a^\e,\\
1928: \widetilde \phi^\e\big|_{t=0}=0\quad ;\quad a^\e\big|_{t=0}&= a^\e_0\, .
1929: \end{aligned}
1930: \end{equation}
1931: The pair $(\widetilde \phi^\e, a^\e$) is bounded in
1932: $C([0,T];H^s)$. Therefore, a subsequence is convergent, and the limit
1933: is given by:
1934: \begin{equation*}
1935: \begin{aligned}
1936: \d_t \widetilde \phi +\nabla
1937: \eik\cdot \nabla \widetilde \phi+ f\(
1938: |a|^2\)&= 0\ ; \quad \widetilde \phi\big|_{t=0}=0, \\
1939: \d_t a +\nabla \eik \cdot \nabla a +\frac{1}{2}a
1940: \Delta \eik &= 0\ ; \quad a\big|_{t=0}= a_0\, .
1941: \end{aligned}
1942: \end{equation*}
1943: We see that $a$ solves \eqref{eq:alibre}; $\widetilde \phi $ is
1944: given by an
1945: ordinary differential equation along the rays associated to $\eik$,
1946: with a source term showing nonlinear effect: $f\( |a|^2\)$. By
1947: uniqueness, the whole sequence is convergent. Roughly speaking, we see that if
1948: \begin{equation*}
1949: \bw^\e = \,^t \(\nabla \( \widetilde \phi^\e-\widetilde \phi\),a^\e -a\),
1950: \end{equation*}
1951: then Gronwall lemma yields:
1952: \begin{equation*}
1953: \(S^\e \d^\alpha_x \bw^\e , \d^\alpha_x \bw^\e\) \le C \( \e +
1954: \e^\kappa\)\le 2 C\e^\kappa.
1955: \end{equation*}
1956: We infer:
1957: \begin{prop}\label{prop:estprecint}
1958: Let $s>2+n/2$. Then \eqref{eq:modifint} has a unique solution
1959: $(a^\e,\phi^\e)\in C([0,T];H^s)^2$, such that $x_k a^\e, x_k\d_j \phi^\e\in
1960: C([0,T];H^s)$, for every $1\le j,k\le n$ ($T$ is given by
1961: Lemma~\ref{lem:hj}). Moreover, there exists
1962: $C_s$ independent of $\e$ such that for every $0\le t\le T$,
1963: \begin{equation}\label{eq:tpetitint}
1964: \| a^\e (t)- a(t)\|_{H^s}\le C_s \e^\kappa \quad ;\quad
1965: \| \phi^\e(t) -\e^\kappa \widetilde \phi\|_{H^s} \le C_s \e^{2\kappa} t ,
1966: \end{equation}
1967: where $a$ is given by \eqref{eq:alibre}.
1968: \end{prop}
1969:
1970: Three cases must be distinguished:
1971: \begin{itemize}
1972: \item If $1/2< \kappa<1$, then we can infer the analogue of
1973: \eqref{eq:BKWVtgrand}.
1974: \item If $\kappa =1/2$, then we can infer the analogue of
1975: \eqref{eq:BKWVtpetit} (but not yet of \eqref{eq:BKWVtgrand}).
1976: \item If $0<\kappa <1/2$, then we must pursue the analysis, and
1977: compute a corrector of order $\e^{2\kappa}$.
1978: \end{itemize}
1979: We shall not go further into detailed computations, but instead, discuss
1980: the whole analysis in a rather loose fashion. However, we note that
1981: all the ingredients have been given for a complete justification.
1982: \smallbreak
1983:
1984: Let $N= [1/\kappa]$, where $[r]$ is the largest integer not larger than
1985: $r>0$. We construct $a^{(1)},\ldots,a^{(N)}$ and
1986: $\widetilde\phi^{(1)},\ldots,\widetilde\phi^{(N)}$ such that:
1987: \begin{align*}
1988: & \left\| a^\e -a-\e^\kappa a^{(1)}-\ldots -\e^{N\kappa}a^{(N)}
1989: \right\|_{L^\infty([0,T];H^s)} +\\
1990: &+\left\| \widetilde\phi^\e -\widetilde \phi -\e^{\kappa}
1991: \widetilde \phi^{(1)}- \ldots -\e^{N\kappa}\widetilde \phi^{(N)}
1992: \right\|_{L^\infty([0,T];H^s)}
1993: =o\( \e^{N\kappa}\).
1994: \end{align*}
1995: But since $N +1 >1/\kappa$, we have:
1996: \begin{equation*}
1997: \left\| \phi^\e -\e^{\kappa} \widetilde \phi -\e^{2\kappa}
1998: \widetilde \phi^{(1)}- \ldots -\e^{N\kappa}\widetilde \phi^{(N-1)}
1999: \right\|_{L^\infty([0,T];H^s)}
2000: =\O\( \e^{(N+1)\kappa}\)=o(\e).
2001: \end{equation*}
2002: The analogue of \eqref{eq:BKWVtgrand} follows:
2003: \begin{equation*}
2004: \left\| u^\e - a e^{ i \eik/\e + i\phi_{\rm app}^\e}
2005: \right\|_{L^\infty([0,T];L^2\cap L^\infty)} =o(1),
2006: \end{equation*}
2007: where
2008: \begin{equation*}
2009: \phi_{\rm app}^\e = \frac{\widetilde \phi}{\e^{1-\kappa}} +\frac{\widetilde
2010: \phi^{(1)}}{\e^{1-2\kappa}}+\ldots +
2011: \frac{\widetilde \phi^{(N-1)}}{\e^{1-N\kappa}}.
2012: \end{equation*}
2013: \begin{rema}
2014: In the case $\kappa =1$, $N=1$, and the above analysis shows that one phase
2015: shift factor appears: we retrieve the function $G$ of
2016: Proposition~\ref{prop:sub} (under the unnecessary assumption $f'>0$). If
2017: $\kappa >1$, then $N=0$, and we see that $a e^{i\eik/\e}$ is a good
2018: approximation for $u^\e$.
2019: \end{rema}
2020:
2021: %%%%%%%%%
2022:
2023: %\appendix
2024:
2025: %\input{gross}
2026: \bibliographystyle{amsplain}
2027: \bibliography{../../carles}
2028:
2029: \end{document}
2030:
2031:
2032:
2033:
2034:
2035: