math0702778/cg.tex
1: \documentclass[leqno]{amsart}
2: \usepackage{amsmath,amsfonts,amssymb}
3: \usepackage{amsthm}
4: \usepackage[dvips]{epsfig}
5:       
6: 
7: 
8: \def\dis
9: {\displaystyle}
10: \def\e
11: {\varepsilon}
12: 
13: \def\C{{\mathbb C}}
14: \def\R{{\mathbb R}}
15: \def\N{{\mathbb N}}
16: \def\Z{{\mathbb Z}}
17: \def\T{{\mathbb T}}
18: \def\Q{{\mathbb Q}}
19: \def\O{{\mathcal O}}
20: \def\si{\sigma}
21: \def\u{{\tt u}}
22: \def\<{\langle}
23: \def\>{\rangle}
24: \def\DT{{\Delta t}}
25: \def\DX{{\Delta x}}
26: \def\({\left(}
27: \def\){\right)}
28: 
29: 
30: \def\Eq#1#2{\mathop{\sim}\limits_{#1\rightarrow#2}}
31: \def\Tend#1#2{\mathop{\longrightarrow}\limits_{#1\rightarrow#2}}
32: 
33: 
34: \def\d{{\partial}}
35: \numberwithin{equation}{section}
36: 
37: 
38: 
39: 
40: 
41: \def\dem
42: {\noindent {\bf Proof.} }
43: 
44: \theoremstyle{plain}
45: \newtheorem{theorem}{Theorem}[section]
46: \newtheorem{lemma}[theorem]{Lemma}
47: \newtheorem{corollary}[theorem]{Corollary}
48: \newtheorem{proposition}[theorem]{Proposition}
49: 
50: 
51: \theoremstyle{definition}
52: \newtheorem{definition}[theorem]{Definition}
53: 
54: \theoremstyle{remark}
55: \newtheorem{remark}[theorem]{Remark}
56: 
57: \numberwithin{equation}{section}
58: 
59: 
60: 
61: 
62: 
63: 
64: 
65: 
66: 
67: 
68: 
69: 
70: 
71: 
72: \begin{document}
73: \title[Numerical aspects of NLS with caustics]{Numerical aspects of
74:   nonlinear Schr\"odinger equations in the presence of caustics}
75: \author[R. Carles]{R\'emi Carles}
76: \address[R. Carles]{Institut CNRS Pauli\\ Wolfgang
77:     Pauli Institute c/o Fak. f. Mathematik.\\
78:         Univ. Wien, UZA 4\\   
79:         Nordbergstr.~15\\ A-1090 Wien\\ Austria\footnote{Permanent
80:   address: Univ. Montpellier 2, UMR CNRS 5149, Math\'ematiques, CC
81:   051, Place Eug\`ene Bataillon, 34095 Montpellier cedex 5, France.}} 
82: \email{Remi.Carles@math.cnrs.fr}
83: \author[L. Gosse]{Laurent Gosse}
84: \address[L. Gosse]{Istituto
85:  per le Applicazioni del Calcolo (sezione di Bari)\\ Via G. Amendola
86:  122\\ 70126 Bari\\ Italy}
87: \email{l.gosse@ba.iac.cnr.it}
88: \thanks{This work was partially
89:   supported by o Centro de Matem\'atica e Aplica\c c\~oes Fundamentais
90:   (Lisbon), funded by FCT as contract POCTI-ISFL-1-209, and by the
91:   Austrian Ministry of Science via its  
92: grant for the Wolfgang Pauli Institute and by the Austrian Science
93: Foundation (FWF) via the START Project (Y-137-TEC).}
94: \begin{abstract}
95: The aim of this text is to develop on the asymptotics
96:   of some 1-D nonlinear Schr\"odinger equations from both the
97:   theoretical and the numerical perspectives, when a caustic is
98:   formed. We review rigorous 
99:   results in the field and give some heuristics in cases where
100:   justification is still needed. The scattering operator theory is
101:   recalled. Numerical 
102:   experiments are carried out on the focus point singularity for which
103:   several results have been proven rigorously. Furthermore, the scattering
104:   operator is numerically studied. Finally, experiments on the cusp
105:   caustic are displayed, and similarities with the focus point are
106:   discussed. 
107: \end{abstract}
108: \subjclass[2000]{35B33, 35P25, 35Q55, 65T50, 81Q20}
109: \keywords{Nonlinear Schr\"odinger equation, time-splitting scheme, Fourier
110: scheme, WKB expansion, caustics.}
111: \maketitle
112: 
113: \section{Introduction}
114: \vspace*{-0.5pt}
115: \noindent
116: We present a numerical study of the semi-classical solutions
117: to the following nonlinear Schr\"odinger equations with $\e \ll 1$,
118: \begin{equation}
119:   \label{eq:nls0}
120:   i\e \d_t \u^\e +\frac{\e^2}{2}\Delta \u^\e
121:   =|\u^\e|^{2\si}\u^\e,\quad(t,x)\in \R_+\times\R^n\quad  ;\quad
122:   \u^\e_{\mid t=0} = \e^p 
123:   f(x)e^{i\phi_0(x)/\e},
124: \end{equation}
125: when a caustic (a point or a cusp) is formed, that is to say, beyond
126: {\it breakup  
127: time}. Since the nonlinearity is homogeneous, the change of unknown function 
128: $u^\e = \e^{p}\u^\e$ shows that \eqref{eq:nls0} is equivalent to:
129: \begin{equation}
130:   \label{eq:nls1}
131:   i\e \d_t u^\e +\frac{\e^2}{2}\Delta u^\e
132:   =\e^{2\si p}|u^\e|^{2\si}u^\e\quad ;\quad u^\e_{\mid t=0} =
133:   f(x)e^{i\phi_0(x)/\e},
134: \end{equation}
135: so that we can always consider initial data of order $\O(1)$.
136: 
137: There are several motivations to study the behavior of \eqref{eq:nls1}
138: when a caustic is formed. First, on a purely academic level, we
139: recall that the description of the caustic crossing is complete in the
140: case of linear equations; see \cite{Du}. For nonlinear equations,
141: very interesting formal computations were proposed in \cite{HK87} (we
142: recall the main idea in Section~\ref{sec:analyt} below). For
143: \emph{dissipative} nonlinear wave equations, Joly, M\'etivier and Rauch
144: \cite{JMRTAMS95,JMRMemoir} have proved that the amplification of the
145: wave near the caustic can ignite the dissipation phenomenon in such a
146: way that the oscillations (that carry highest energy) are absorbed. 
147: The above nonlinear Schr\"odinger equation is the simplest model of a
148: \emph{conservative, nonlinear equation}. The mass and the energy of
149: the solution are independent of time (see \eqref{eq:conserv}
150: below). Therefore, different nonlinear mechanisms are expected. We
151: recall in Section~\ref{sec:analyt} some results that have been
152: established rigorously, and give heuristic arguments to extend these
153: results. This serves as a guideline for the numerical experiments
154: proposed after. 
155: \smallbreak
156: 
157: Second, \eqref{eq:nls1} may be considered as a
158: simplified model for Bose--Einstein condensation, which may be 
159: modeled (see e.g. \cite{DGPS,PiSt}) by:
160: \begin{equation}
161:   \label{eq:nlsharmo}
162:   i\e \d_t u^\e +\frac{\e^2}{2}\Delta u^\e
163:   =\omega^2 \frac{|x|^2}{2}u^\e +\e^{2}|u^\e|^{2\si}u^\e,
164: \end{equation}
165: with $\si =2$ if $n=1$, and $\si =1$ if $n=2$ or $3$. 
166: The power $\e^2$ in front of the nonlinearity depends on the r\'egime
167: considered, and in particular on the respective scales of different
168: parameters (see e.g. \cite{BurqZworski} and references therein). 
169: The role of the harmonic potential $|x|^2$ is to model a magnetic
170: trap. In the semi-classical limit $\e\to 0$ for the linear equation,
171: this potential causes focusing at the origin for solutions whose data are
172: independent of $\e$. This is to be compared with the case of
173: \eqref{eq:nls1} with initial quadratic oscillations as considered
174: below: the initial quadratic oscillations force the solution to
175: concentrate at one point in the limit $\e \to 0$. The parallel between
176: \eqref{eq:nls1} and \eqref{eq:nlsharmo} 
177: was extended and justified in \cite{CaIHP} for these nonlinear
178: equations.
179: \smallbreak
180: 
181: From both points of view, when a caustic point is formed, the caustic
182: crossing may be described in terms of the scattering operator
183: associated to 
184: \begin{equation*}
185:   i\d_t \psi +\frac{1}{2}\Delta\psi = |\psi|^{2\si}\psi.
186: \end{equation*}
187: This aspect is recalled in Section~\ref{sec:analyt}. For this reason,
188: we also pay a particular attention to this operator, independently of
189: the above semi-classical limit. Note  that besides the existence
190: of this operator, very few of its properties (dynamical, for instance)
191: are known. 
192: \smallbreak
193: 
194: In this paper, we
195: always assume $2\si p\ge 1$: one of the reasons is that when
196: $0\le 2\si p<1$, instability occurs, see
197: \cite{CaARMA,CaBKW}. Suppose for instance that $u^\e$ solves
198: \eqref{eq:nls1}, and that $\widetilde u^\e$ solves \eqref{eq:nls1},
199: where $f$ replaced by $(1+\delta^\e)f$, where $\delta^\e$ is a sequence
200: of real numbers going to zero as $\e \to 0$. Then there are some
201: choices of $\delta^\e$ for which 
202: \begin{equation*}
203:   \liminf_{\e \to 0}\|u^\e(t^\e)- \widetilde u^\e(t^\e)\|_{L^2}>0,
204: \end{equation*}
205: for some sequence of time $t^\e\to 0$ (see \cite{CaARMA}, and
206: \cite{BurqZworski} for a similar phenomenon with different initial
207: data). Therefore, 
208: producing reliable numerical tests in 
209: the case $0\le 2\si p<1$ (which is super-critical as far as WKB
210: analysis is concerned \cite{CaARMA,CaBKW}) seems to be a very delicate
211: issue, that we leave out in the present paper. 
212: \smallbreak
213: 
214: The rest of this paper is structured as follows. In
215: Section~\ref{sec:analyt}, we recall the general approach of WKB
216: analysis for the Schr\"odinger equation, the arguments of \cite{HK87},
217: and the rigorous results available for the semi-classical limit of
218: \eqref{eq:nls1} when a caustic reduced to a  point is formed. We then
219: recall the definition of the scattering operator. We also give
220: heuristic arguments to tackle the case of a ``supercritical focal
221: point'', and to guess what the critical indices are when a cusp
222: caustic is formed, instead of a focal point. 
223: In Section~\ref{sec:numgen}, we present the different strategies that
224: have been followed in the literature to study numerically the
225: semi-classical limit for nonlinear Schr\"odinger equations. 
226: Numerical experiments on the semi-classical limit for \eqref{eq:nls1}
227: in the presence of a focal point appear in Section~\ref{sec:foc}, and
228: the scattering operator is simulated in Section~\ref{sec:scattnum}. 
229: We present the numerical experiments of the semi-classical limit for
230: \eqref{eq:nls1} in the presence of a cusp caustic in
231: Section~\ref{sec:cusp}, and make conclusive remarks in
232: Section~\ref{sec:concl}. 
233: 
234: \section{Analytical approach}\label{sec:analyt}
235: \vspace*{-0.5pt}
236: \noindent
237: \subsection{Semi-classical limit of the free Schr\"odinger equation}
238: Consider the initial value problem, for $(t,x)\in \R_+\times\R^n$:
239: \begin{equation}
240:   \label{eq:schrodlibre}
241:   i\e \d_t v^\e +\frac{\e^2}{2}\Delta v^\e =0\quad ;\quad v^\e_{\mid
242:   t=0} = f(x)e^{i\phi_0(x)/\e}.
243: \end{equation}
244: The aim of WKB methods is to describe the asymptotic behavior of
245: $v^\e$ as $\e \to 0$. For instance, $\e$ can be related to the Planck
246: constant, and the asymptotic behavior of $v^\e$ is expected to yield a
247: good description of $v^\e$ when $\e$ is fixed, but small compared to
248: the other parameters. More precisely, seek
249: $v^\e$ of the form
250: \begin{equation}
251:   \label{eq:bkw}
252:   v^\e(t,x)\sim e^{i\phi(t,x)/\e}\(a_0(t,x)+\e a_1(t,x)+\ldots\)
253:   \quad\text{as }\e \to 0.
254: \end{equation}
255: Plugging this expansion into \eqref{eq:schrodlibre} and canceling the
256: $\O(\e^0)$ term, we see that
257: the phase $\phi$ must solve the eikonal equation:
258: \begin{equation}
259:   \label{eq:eikonale}
260:   \d_t \phi +\frac{1}{2}|\nabla \phi|^2 =0\quad ;\quad \phi_{\mid
261:   t=0}=\phi_0. 
262: \end{equation}
263: To cancel the $\O(\e^1)$ term,  the leading order amplitude
264: solves the transport equation:
265: \begin{equation}
266:   \label{eq:transport}
267:   \d_t a_0 +\nabla\phi \cdot \nabla a_0+\frac{1}{2}a_0 \Delta \phi
268:   =0\quad ;\quad a_{0\mid t=0}=f.
269: \end{equation}
270: The eikonal equation \eqref{eq:eikonale} is solved thanks to
271: Hamilton-Jacobi theory\footnote{but not according to the theory of viscosity
272: solutions! See e.g. \cite{koko}.}: $\phi$ is constructed locally in space and
273: time (see e.g. \cite{CaBKW} for a discussion on this aspect). Even if
274: $\phi_0$ is smooth, $\phi$ develops singularities in finite time in
275: general: the locus where $\phi$ is singular is called \emph{caustic}
276: (see e.g. the second volume of \cite{Hormander}). When $\phi$ becomes
277: singular, all the terms $a_0,a_1,\ldots$ may become singular as well. 
278: One easily observes that (\ref{eq:transport}) admits a ``divergence form":
279: $\d_t |a_0|^2+ \nabla \cdot (|a_0|^2 \nabla \phi)=0$.
280: To illustrate this general discussion, we consider two examples that
281: will organize the rest of this paper.\\
282: 
283: \noindent\emph{Example (Quadratic phase). }
284:   Let $\phi_0(x) = -\frac{|x|^2}{2}$. Then \eqref{eq:eikonale} and
285:   \eqref{eq:transport} can be  solved explicitly:
286:   \begin{equation*}
287:     \phi(t,x)=\frac{|x|^2}{2(t-1)}\quad ;\quad
288:     a(t,x)=\frac{1}{(1-t)^{n/2}}f\( \frac{x}{1-t}\).
289:   \end{equation*}
290: This shows that as $t\to 1$, $\phi$ and $a$ become singular: the wave
291: $u^\e$ focuses at the origin. This example can be viewed as the smooth
292: counterpart of the Cauchy problem
293: \begin{equation*}
294:   i\d_t \psi+\frac{1}{2}\Delta \psi=0\quad ;\quad \psi_{\mid t=0}=
295:   e^{-i\frac{|x|^2}{2}} .
296: \end{equation*}
297: Fourier analysis shows that $\psi_{\mid t=1}=\delta$, the Dirac
298: measure at the origin.\\ 
299: 
300: 
301: Of course, the solution of \eqref{eq:schrodlibre} can be represented
302: as an oscillatory integral:
303: \begin{equation}\label{eq:intosc}
304:   v^\e(t,x) = \frac{1}{(2i\pi t)^{n/2}}\int e^{i\frac{|x-y|^2}{2\e
305:   t} +i\frac{\phi_0(y)}{\e}}f(y)dy. 
306: \end{equation}
307: The caustic set is exactly the locus where the
308: critical points for the phase
309: \begin{equation*}
310:  \Phi_{t,x}(y)= \frac{|x-y|^2}{2 t}+\phi_0(y)
311: \end{equation*}
312: are degenerate. Outside the caustic, an approximation of $v^\e$ is
313: given by the stationary phase theorem (that we recalled as simply as 
314: possible in \cite{GosseWKB}). This leads us to the second
315: example we shall consider numerically:\\
316: \noindent
317: \emph{Example (Cusp). }
318:   Let $n=1$ and $\phi_0(x)=\cos x$. 
319: The set of degenerate critical points for $\Phi_{t,x}(y)$ (caustic) is
320: given implicitly by: 
321: \begin{equation*}
322:   {\mathcal C}=\left\{ (t,x)\in \R_+\times \R; \exists y\in \R ,\
323:   \frac{y-x}{t}=\sin y,\text{ and 
324:   }\frac{1}{t}=\cos y\right\}. 
325: \end{equation*}
326: As soon as $t\ge 1$, a caustic is formed (see Figure 2 in
327: \cite{GosseJinLi}). \\
328: 
329: 
330: When considering the asymptotic behavior of $u^\e$ beyond the caustic,
331: two main features must be considered: the creation of other phases\footnote{in
332: our mind, phases are always associated to oscillations whose
333: period depends on $\e$, and goes to infinity as $\e\to 0$ - rapid
334: oscillations. The wavelength may be proportional to $\e$, or, say, to
335: $\sqrt \e$.}, 
336: and  the Maslov index (see \cite{Du} for more general linear
337: equations). In the case of a focal point, the first aspect does not
338: exist: there is no creation of phase, and one phase is enough to
339: describe $v^\e$ past the focal point $(t,x)=(1,0)$. One can prove
340: easily the following result:
341: \begin{lemma}\label{lem1}
342:   Let $n \ge 1$ and $f\in {\mathcal S}(\R^n;\C)$. If
343:   $\phi_0(x)=-|x|^2/2$, then the asymptotic behavior (in $L^2(\R^n)$)
344:   of the solution $v^\e$ to \eqref{eq:schrodlibre} is given by:
345:   \begin{equation*}
346:     v^\e(t,x)\Eq \e 0
347: \left\{
348:   \begin{aligned}
349:     \frac{e^{i|x|^2/(2\e (t-1))}}{(1-t)^{n/2}}f\(\frac{x}{1-t}\)&
350:     \ \text{ if }t<1,\\
351: e^{-in\frac{\pi}{2}}\frac{e^{i|x|^2/(2\e
352:     (t-1))}}{(t-1)^{n/2}}f\(\frac{x}{1-t}\)& 
353:     \ \text{ if }t>1.
354:   \end{aligned}
355: \right.
356:   \end{equation*}
357: \end{lemma}
358: In this example, the Maslov index is $-n\pi/2$. In the case of the
359: cusp, three phases must be considered to describe the asymptotic
360: behavior of $v^\e$ beyond the caustic (see
361: e.g. \cite{GosseJinLi,GosseWKB}). \\
362: %% ajout 1
363: For future discussion on the numerical results, we state the following
364: more precise result, which follows from the stationary phase theorem:
365: \begin{lemma}\label{lem2}
366:   Let $n \ge 1$ and $f\in {\mathcal S}(\R^n;\C)$. If
367:   $\phi_0(x)=-|x|^2/2$, then the asymptotic behavior of the solution
368:   $v^\e$ to \eqref{eq:schrodlibre} at time $t=2$ is given by: 
369:   \begin{equation*}
370:     v^\e(2,x)=e^{-in\frac{\pi}{2}}e^{i|x|^2/(2\e)}f\(-x\)+ O(\e)\quad
371:     \text{in }L^2\cap  L^\infty(\R^n).
372:   \end{equation*}
373: \end{lemma}
374: %% fin ajout 1 
375: \subsection{Caustics in the nonlinear case: heuristics}
376: \label{sec:heur}
377: 
378: Consider now the perturbation of \eqref{eq:schrodlibre} with a
379: nonlinear term:
380: \begin{equation}
381:   \label{eq:nls}
382:   i\e \d_t u^\e +\frac{\e^2}{2}\Delta u^\e
383:   =\e^{\alpha}|u^\e|^{2\si}u^\e\quad ;\quad u^\e_{\mid 
384:   t=0} = f(x)e^{i\phi_0(x)/\e}.
385: \end{equation}
386: The sign of the nonlinearity is chosen so that no finite time blow-up
387: occurs. The following two important
388: quantities are formally independent of time:
389: \begin{equation}
390:   \label{eq:conserv}
391:   \begin{aligned}
392:   \text{Mass: }& \|u^\e(t)\|_{L^2}=\text{Const.}= \|f\|_{L^2}.\\
393: \text{Energy: }& E^\e(t):= \frac{1}{2}\|\e \nabla u^\e(t)\|^2_{L^2}
394: +\frac{\e^{\alpha}}{\si +
395:   1}\|u^\e(t)\|_{L^{2\si+2}}^{2\si+2}=E^\e(0).
396: \end{aligned}
397: \end{equation}
398: We refer to \cite{CazCourant} for a justification. Fix the power
399: $2\si>0$ of the nonlinearity, and consider different values for
400: $\alpha$. Two notions of criticality arise: for the WKB methods on the
401: one hand, and for the caustic crossing on the other hand. This
402: discussion is presented in \cite{HK87} for conservation laws, and we
403: summarize it  in the case of \eqref{eq:nls}. Plugging an expansion of
404: the form \eqref{eq:bkw} into \eqref{eq:nls}, we see that the value
405: $\alpha =1$ is critical for the WKB methods: if $\alpha >1$, then the
406: nonlinearity does 
407: not affect the transport equation \eqref{eq:transport} (``linear
408: propagation''), while if 
409: $\alpha =1$, then the nonlinearity appears in the right hand side of
410: \eqref{eq:transport} (``nonlinear propagation''). Recall that in this
411: paper, we always assume $\alpha \ge 1$. Therefore, the eikonal equation
412: \eqref{eq:eikonale} is not altered: 
413: the geometry of the propagation remains the same as in the linear 
414: WKB approach,
415: and we have to face the same caustic sets. The idea presented in
416: \cite{HK87} consists in saying that according to the geometry of the
417: caustic, different notions of criticality exist, as far as $\alpha$ is
418: concerned, near the
419: caustic. In the linear setting \eqref{eq:schrodlibre}, the influence
420: of the caustic is relevant only in a neighborhood of this set
421: (essentially, in a boundary layer whose size depends on $\e$ and the
422: geometry of $\mathcal C$). View the nonlinearity in \eqref{eq:nls}
423: as a potential, and assume that the nonlinear effects are negligible
424: near the caustic: then $u^\e\sim v^\e$ near $\mathcal C$. 
425: View the term $\e^\alpha |u^\e|^{2\si}$ as a (nonlinear)
426: potential. 
427: The average nonlinear effect near $\mathcal C$ is expected to be:
428: \begin{equation*}
429:  \e^{-1}\int_{{\mathcal C}^\e}\e^\alpha |u^\e|^{2\si}\sim
430:  \e^{-1}\int_{{\mathcal C}^\e}\e^\alpha |v^\e|^{2\si},  
431: \end{equation*}
432: where ${\mathcal C}^\e$ is the region where caustic effects are
433: relevant, and the factor $\e^{-1}$ is due to the integration in
434: time (recall that there is an $\e$ in front of the time derivative in
435: \eqref{eq:nls}).  The idea of this heuristic argument is that when the
436: nonlinear 
437: effects are negligible near $\mathcal C$ (in the sense that the
438: uniform norm of $u^\e-v^\e$ is small compared to that of $v^\e$ near
439: ${\mathcal C}^\e$), the above approximation should be valid. On the
440: other hand, it is expected that it ceases to be valid precisely when
441: nonlinear effects can no longer be neglected near the caustic:
442: $u^\e-v^\e$ is of the same order of magnitude as $v^\e$ in
443: $L^\infty({\mathcal C}^\e)$, or even larger. 
444: 
445: Practically, assume that in the linear case, $v^\e$ has an
446: amplitude $\e^{-\ell}$ in a boundary layer of size $\e^{k}$; then the
447: above quantity is
448: \begin{equation*}
449:   \e^{-1}\int_{{\mathcal C}^\e}\e^\alpha |v^\e|^{2\si} \sim
450:   \e^{-1}\e^\alpha | \e^{-\ell}|^{2\si}\e^{k}.
451: \end{equation*}
452: The value $\alpha $ is then critical when the above cumulated effects
453: are not negligible:
454: \begin{equation*}
455:   \alpha_c = 1+2\ell \si -k.
456: \end{equation*}
457: When $\alpha>\alpha_c$, the nonlinear effects are expected to be
458: negligible near the caustic: resuming the terminology of \cite{HK87},
459: we speak of ``linear caustic''. The case $\alpha = \alpha_c$ is called
460: ``nonlinear caustic''.
461: To conclude this paragraph, we examine this approach in the case of
462: our two examples.
463: In the case of a focal point, we have $k=1$ and $\ell=n/2$. This leads us
464: to the value:
465: \begin{equation*}
466:   \alpha_c(\text{focal point}) = n\si.
467: \end{equation*}
468: In the case of the cusp in dimension one, we have $k=2/3$ and
469: $\ell=1/3$ (which can be viewed thanks to the Airy function and its
470: asymptotic expansion, see e.g. \cite{Du,Hormander,HK87} or
471: \cite{Ludwig}), which yields:  
472: \begin{equation*}
473:   \alpha_c(\text{cusp in 1D}) = \frac{2\si +1}{3}.
474: \end{equation*}
475: One aspect of the numerical experiments presented below is to test
476: this notion of criticality in those two examples. 
477: 
478: 
479: \subsection{Justification for a focal point, and more heuristics}
480: \label{sec:justif}
481: In this paragraph, we assume $\phi_0(x)= -|x|^2/2$. A complete
482: justification of the above discussion is available \cite{CaIUMJ}:\\
483: \smallbreak
484: \begin{center}
485: \begin{tabular}[c]{l|c|c}
486:  &$\alpha >n\si$  & $\alpha =n\si$  \\
487: \hline
488: $\alpha >1$ &Linear caustic, & Nonlinear caustic,\\
489: &linear propagation   & linear propagation \\
490: \hline
491: $\alpha =1$&Linear caustic, &Nonlinear caustic,\\
492: &nonlinear propagation   & nonlinear propagation \\
493: \end{tabular}
494: \end{center}
495: \smallbreak
496: 
497: Consider $t\in [0,2]$, which includes the caustic crossing. The above
498: tables means: 
499: \begin{itemize}
500: \item If $\alpha >\max (1,n\si)$, then $u^\e$ can be approximated
501:   by $v^\e$ for $t\in [0,2]$. 
502: \item If $\alpha =1>n\si$, then the nonlinearity is negligible
503:   near the focal point, but not away from it. 
504: \item If $\alpha =n\si>1$, then nonlinear effects are relevant
505:   near the focal point, and only near the focal point.
506: \item If $\alpha =n\si=1$, then the nonlinearity is never
507:   negligible. 
508: \end{itemize}
509: %% ajout 2
510: We give some precisions in some cases of interest for the
511: numerics presented below. \\
512: 
513: In the one-dimensional case $n=1$, the following pointwise estimate is
514: proved in \cite{CaIUMJ} when $\alpha >\max (1,\si)$ or $\alpha =\si>1$:
515: \begin{equation}\label{eq:estponctuelle}
516:   |u^\e(t,x)|\le \frac{C}{\sqrt{|t-1|+\e}}\cdot
517: \end{equation}
518: Setting $w^\e=u^\e-v^\e$, we see that
519: \begin{equation*}
520:   i\e \d_t w^\e +\frac{\e^2}{2}\d_x^2 w^\e = \e^\alpha
521:   |u^\e|^{2\si}u^\e \quad ;\quad w^\e_{\mid t=0}=0.
522: \end{equation*}
523: The usual energy estimate yields, for $t\ge 0$:
524: \begin{equation*}
525:   \|w^\e(t)\|_{L^2}\le \e^{-1}\int_0^t \e^\alpha
526:   \big\| |u^\e(\tau)|^{2\si}u^\e(\tau)\big\|_{L^2}d\tau.
527: \end{equation*}
528: Using \eqref{eq:estponctuelle}, we infer, for $t\in [0,2]$:
529: \begin{equation*}
530:   \|w^\e(t)\|_{L^2}\lesssim
531:   \e^{\alpha-1}\(\int_{\{|\tau-1|>\e\}\cap \{\tau \in
532:   [0,2]\}}\frac{d\tau}{|\tau 
533:   -1|^\si}+\int_{|\tau-1|\le \e}\frac{d\tau}{\e^\si} \)\lesssim
534:   \e^{\alpha -\si}. 
535: \end{equation*}
536: Using the operator $\e\d_x$ and $x/\e+i(t-1)\d_x$, and
537: Gagliardo--Nirenberg inequalities as in \cite{CaIUMJ}, we find:
538: \begin{lemma}\label{lem3}
539:   Let $n=1$, $\alpha >\max (1,\si)$, $f\in {\mathcal S}(\R^n;\C)$, and
540:   $\phi_0(x)=-|x|^2/2$. Then we have, for the solutions of
541:   \eqref{eq:schrodlibre} and \eqref{eq:nls}:
542:   \begin{equation*}
543:     \sup_{0\le t\le 2} \big\| u^\e(t)-v^\e(t)\big\|_{L^2}\le C
544:     \e^{\alpha -\si}\quad ;\quad \big\| u^\e(t)-v^\e(t)\big\|_{L^\infty}\le C
545:     \frac{\e^{\alpha -\si} }{\sqrt{|t-1|+\e}}, \quad t\in [0,2].
546:   \end{equation*}
547: In particular, Lemma~\ref{lem3} implies, at time $t=2$:
548: \begin{equation*}
549:     u^\e(2,x)=e^{-in\frac{\pi}{2}}e^{i|x|^2/(2\e)}f\(-x\)+
550:     O\(\e^{\min(1,\alpha -\si)}\)\quad
551:     \text{in }L^2\cap  L^\infty(\R).
552:   \end{equation*}
553: \end{lemma}
554: (The above result is true also when $\alpha=\si>1$, but becomes far
555: less interesting.)
556: %% fin ajout 2
557: We now explain the critical case $\alpha =n\si>1$. The nonlinear effects near
558: the focal point are described in terms of the scattering operator
559: associated to the nonlinear Schr\"odinger equation. We rapidly present
560: this operator $S$ in Section~\ref{sec:scatt}. We have then:
561: \begin{equation*}
562:     u^\e(t,x)\Eq \e 0 
563: \left\{
564:   \begin{aligned}
565:     \frac{e^{i|x|^2/(2\e (t-1))}}{(1-t)^{n/2}}f\(\frac{x}{1-t}\)&
566:     \ \text{ if }t<1,\\
567: e^{-in\frac{\pi}{2}}\frac{e^{i|x|^2/(2\e
568:     (t-1))}}{(t-1)^{n/2}}Zf\(\frac{x}{1-t}\)& 
569:     \ \text{ if }t>1,
570:   \end{aligned}
571: \right.
572:   \end{equation*}
573: where $Z= {\mathcal F}\circ S\circ {\mathcal F}^{-1}$ is the conjugate
574: of $S$ by the Fourier transform (see 
575: \cite{CaIUMJ}),
576: \begin{equation*}
577:   {\mathcal F} \varphi(\xi)=\frac{1}{(2i\pi)^{n/2}}\int e^{-ix\cdot
578:   \xi}\varphi(x)dx. 
579: \end{equation*}
580: (Since $S$ is a nonlinear operator, the normalization of the Fourier
581: transform is an important detail.)
582: \smallbreak 
583: 
584: To conclude this paragraph, we give a few hints of what happens or may
585: happen when the propagation is linear, and the caustic is
586: super-critical, that is $1<\alpha <n\si$. First, the conservation
587: of mass and energy seem to rule out the possibility of a concentration
588: of the form
589: \begin{equation}\label{eq:concentration}
590:   u^\e(1,x)\sim \frac{1}{\e^{n/2}}\varphi \(\frac{x}{\e}\),  
591: \end{equation}
592: for some function $\varphi$ independent of $\e$. The above
593: relation holds for $v^\e$ and $u^\e$ when $\alpha >\max(1,n\si)$,
594: with $\varphi ={\mathcal F}f$ (for $v^\e$, this is obvious from
595: \eqref{eq:intosc}). When $\alpha =n\si>1$ (linear propagation and
596: nonlinear focal point), the above relation still holds, with a
597: different profile $\varphi$ (see \cite{CaIUMJ}). Now we see that the
598: energy is bounded as $\e\to 0$:
599: \begin{equation*}
600:   E^\e(t)= E^\e(0)\Eq \e 0 \frac{1}{2}\|x f\|_{L^2}^2.
601: \end{equation*}
602: Plugging a concentrating profile as in \eqref{eq:concentration} in the
603: second term of the energy would yield, thanks to the conservation of mass:
604: \begin{equation*}
605:   \e^\alpha \|u^\e(1)\|_{L^{2\si+2}}^{2\si+2}\thickapprox \e^{\alpha
606:   -n\si}\Tend \e 0 +\infty,
607: \end{equation*}
608: which is impossible since the energy is the sum of two positive
609: terms. This suggests two possible effects: near $t=1$,  the
610: amplification of the solution (in terms of powers of $\e$) may be
611: weaker than in 
612: the linear case; on the other hand,  nonlinear effects near the 
613: caustic should affect the \emph{phase} of the solution $u^\e$ in a
614: rather strong way, causing the appearance of new frequencies. A partial
615: justification of the last assertion may be found in
616: \cite{CaCascade}. 
617: \smallbreak
618: 
619: 
620: Extrapolating this argument, we expect that in the supercritical case
621: for a cusp (with $n=1$: $\frac{2\si+1}{3}>\alpha>1$), new
622: frequencies appear. If as in 
623: the linear case,
624: three phases are necessary to describe the solution
625: past the caustic, then the nonlinear interaction of these phases might
626: reveal the presence of new frequencies, even on the modulus of $u^\e$
627: (see Sect.~\ref{sec:cusp} for numerical 
628: tests that seem to confirm this heuristics). 
629: 
630: \subsection{The scattering operator for NLS}
631: \label{sec:scatt}
632: To explain what the operator $S$ mentioned in the previous section is,
633: consider the nonlinear Schr\"odinger equation 
634: \begin{align}
635:   i\d_t \psi +\frac{1}{2}\Delta \psi & = |\psi|^{2\si}\psi\quad ;\quad
636:   (t,x)\in \R\times \R^n,\label{eq:NLS} \\
637:   U(-t)\psi(t)_{\mid t=t_0}&= \psi_-,\label{ci:NLS}
638: \end{align}
639: where $U(t)= e^{i\frac{t}{2}\Delta}$ is the propagator of the linear
640: equation. To construct the scattering operator, we first want to give a
641: meaning to \eqref{eq:NLS}--\eqref{ci:NLS} when $t_0=-\infty$. This
642: means that the nonlinear effects are asymptotically negligible as $t\to
643: -\infty$: for instance, we expect at least
644: \begin{equation*}
645:   \| U(-t)\psi(t) -\psi_-\|_{L^2}= \| \psi(t) -U(t)\psi_-\|_{L^2}\Tend
646:   t {-\infty} 0.
647: \end{equation*}
648: This gives a rigorous meaning to the relation $\psi(t,x)\sim
649: U(t)\psi_-(x)$ which aims at saying that  as time goes to $-\infty$,
650: the nonlinear dynamics associated to \eqref{eq:NLS} can be compared to
651: the free dynamics given by $e^{i\frac{t}{2}\Delta}$. 
652: 
653: To define a scattering operator, we want to be able to say that as
654: $t\to +\infty$ as well, the nonlinear effect are asymptotically
655: negligible. That is, there exists $\psi_+\in L^2(\R^n)$ such that
656: \begin{equation*}
657:   \| U(-t)\psi(t) -\psi_+\|_{L^2}= \| \psi(t) -U(t)\psi_+\|_{L^2}\Tend
658:   t {+\infty} 0.
659: \end{equation*}
660: The scattering operator is then defined as $S:\psi_-\mapsto
661: \psi_+$. Since our numerical experiments concern the one-dimensional
662: case, we recall the existence of the scattering operator in this
663: setting, and refer to
664: \cite{CazCourant,CW92,GOV94,GV79Scatt,NakanishiOzawa} for some
665: extensions to the multidimensional framework: define 
666: \begin{equation*}
667:   \Sigma := H^1\cap {\mathcal F}(H^1)=\{ f\in L^2(\R^n)\ ;\
668:   \|f\|_\Sigma := \|f\|_{L^2}+\|x f\|_{L^2} + \|\nabla
669:   f\|_{L^2}<\infty\}. 
670: \end{equation*}
671: \begin{proposition}[Scattering theory]\label{prop:scattering}
672: Let $n=1$, and assume $\si \geq \frac{1+\sqrt{17}}{4} (>1)$.
673: \begin{itemize}
674: \item For every  $\psi_- \in \Sigma $,
675:  there exists a unique $\varphi\in \Sigma$ 
676:  such that the maximal solution $\psi\in C(\R,\Sigma)$ to
677:  \eqref{eq:NLS} satisfies $\psi_{\mid t=0}=\varphi$ and 
678: $$\left\| U(-t)\psi(t)-\psi_-\right\|_\Sigma \Tend
679: t {-\infty} 0.$$
680: \item For every $\varphi\in \Sigma$, there exists a unique
681:  $\psi_+\in \Sigma$  
682:  such that the maximal solution $\psi\in C(\R,\Sigma)$ to
683:  \eqref{eq:NLS} with $\psi_{\mid t=0}=\varphi$ satisfies
684: $$\left\| U(-t)\psi(t)-\psi_+\right\|_\Sigma \Tend
685: t {+\infty} 0.$$
686: \end{itemize}
687: The scattering operator is $S:\psi_-\mapsto \psi_+$. When $\si >1$,
688: the above conclusions remain, in a neighborhood of the origin. When
689: $\si>1$, we also have:
690: \begin{itemize}
691: \item For every $\psi_-\in\Sigma$, there exist a unique solution
692:  $\psi\in C(\R,H^1)$ to 
693:  \eqref{eq:NLS} and a unique $\psi_+\in
694:   H^1(\R)$  such that:
695:  \begin{equation*}
696:    \left\| U(-t)\psi(t)-\psi_-\right\|_{\Sigma} \Tend
697: t {-\infty} 0\quad ;\quad \left\| U(-t)\psi(t)-\psi_+\right\|_{L^2} \Tend
698: t {+\infty} 0.
699:  \end{equation*}
700: \end{itemize}
701: \end{proposition}
702: 
703: When $\si \le 1$, the above conclusions are false: if $\si =1$ for
704: instance, and if $\psi_-\in L^2$ with
705: $U_0(-t)\psi(t)-\psi_- \to 0$ in $L^2$ as $t\to -\infty$, then
706: $\psi=\psi_-=0$. One cannot compare the nonlinear dynamics with the
707: free dynamics (see \cite{Barab,Strauss74,Strauss81,Ginibre}). \\
708: Note that even though the scattering is proven to exist, very few of its
709: features are known. We refer to \cite{CazCourant} for some algebraic
710: properties. At least, this operator is not trivial: near
711: the origin, it is a non-trivial perturbation of the identity (see
712: \cite{CaWigner}). 
713: 
714: \section{Numerical approximation of semi-classical Schr\"odinger
715:   equations}\label{sec:numgen} 
716: \vspace*{-0.5pt}
717: \noindent
718: 
719: Hereafter we restrict our discussion to the one-dimensional case, that
720: is to say $n=1$ in all the preceding considerations. 
721: 
722: \subsection{Rigorous results for general time-splitting schemes}
723: 
724: It is interesting to notice that in the numerical literature, the
725: nonlinear equation \eqref{eq:nls} is treated exactly the same way the
726: linear one \eqref{eq:schrodlibre} would be in the presence of a
727: potential $V$ on its right-hand side. The strategy is called {\it
728:   time-splitting}, in its first or second order version (Lie or Strang
729: splitting, see e.g. \cite{BBD}) where one alternates every time step
730: $\DT>0$ between the solving of the Laplace operator and the handling
731: of the (nonlinear) differential equation. According to
732: Section \ref{sec:scatt}, $U$ will still stand for the free propagator,
733: whereas we shall use $V$ as the ODE solver; Lie time-splitting
734: algorithms generate the following type of approximation for
735: \eqref{eq:nls}, 
736: $$u^\e(n\DT,.) \simeq u^\e_\DT(n\DT,.):=[V(\DT) \circ U(\DT)]^n u^\e(t=0,.),$$
737: and Strang splittings, 
738: $$u^\e(n\DT,.) \simeq \tilde u^\e_\DT(n\DT,.):=V(\DT/2) \circ U(\DT)
739: [V(\DT) \circ U(\DT)]^{n-1} V(\DT/2) u^\e(t=0,.).$$ 
740: Many references exist; let us quote only \cite{BBD,MPP,BJM1,BJM2}.
741: 
742: On the contrary, few rigorous convergence results are available, hence
743: we shall mainly recall the results from \cite{BBD} which quantify
744: accurately the {\it splitting errors} assuming each time-step is
745: performed exactly\footnote{but we shall see in the sequel that this is
746:   far from being the case!}. Under this assumption, there holds: 
747: \begin{proposition}(\cite{BBD}, Theorem 4.1) 
748:   For any $T>0$ and $u^\e(t=0,.) \in H^2$, there
749:   exists a constant $C$ depending on the initial data for
750:   \eqref{eq:nls} and $h_0$, such that for $\DT \in [0,h_0]$ and $n\DT
751:   <T$,  
752: $$\|u^\e(n\DT,.) -u^\e_\DT(n\DT,.)\|_{L^2}\leq C h_0.$$
753: If moreover $u^\e(t=0,.) \in H^4$, then there holds under the same assumptions:
754: $$\|u^\e(n\DT,.) -\tilde u^\e_\DT(n\DT,.)\|_{L^2}\leq C h^2_0.$$
755: \end{proposition}
756: 
757: This result is concerned with splitting errors only and relies on the
758: knowledge of the exact solution operators $U$ and $V$. In order to
759: stick to this framework in the context of smooth solutions, it is
760: rather natural to approximate $U$ by means of a Fourier scheme taking
761: advantage of optimized FFT routines, as proposed in the paper
762: \cite{BJM2}. Moreover, this will guarantee that the $L^2$ norm (Mass)
763: of the numerical solution will be conserved up to round-off
764: errors. Unfortunately, the Hamiltonian $E^\e(t)$ is generally not
765: preserved; a method conserving both quantities exists (see the
766: so--called MCN algorithm, page 253 of \cite{DSS}) but it wouldn't be
767: efficient in the semiclassical regime because of the results in
768: \cite{MPP}. 
769: 
770: \subsection{Specific issues with finite-difference discretizations}
771: 
772: This is the main purpose of the paper \cite{MPP} to illustrate the
773: (surprising) fact that in semi-classical regime, usual
774: finite-difference schemes for \eqref{eq:schrodlibre} can deliver very
775: wrong approximations without any particular sign of instability in
776: case very restrictive meshing constraints turn out to be
777: bypassed. This can be quite easily checked through the location of
778: caustics, for instance. The analysis of those standard schemes has
779: been carried out by means of Wigner measures, so the conclusions hold
780: essentially for the quadratic observables coming out of the wave
781: function itself. 
782: 
783: \subsection{The case of FFT-based schemes}
784: 
785: This class of schemes became popular after the publication of
786: \cite{BJM1,BJM2}, mainly because treating the differential part of
787: (\ref{eq:nls}) by means of a discrete Fourier transform looked very
788: much like being the best possible compromise in terms of meshing
789: constraints. Indeed, in the linear case where (\ref{eq:schrodlibre})
790: is supplemented with a potential $V(x)$ on its right-hand side, it was
791: shown that the time-step $\DT$ could be chosen independent of $\e$
792: whereas the space discretization has to satisfy $\DX=\O(\e)$. This was
793: already much better when compared to finite-differences; moreover, the
794: method is naturally $L^2$-conservative. In \cite{BJM2}, these authors
795: extended their ``Fourier framework" to the weakly nonlinear
796: Schr\"odinger equations of the form (\ref{eq:nls}).  
797: %
798: %\begin{figure}[ht]
799: %\centerline{\epsfig{file=pb-FFT-cos(50.2.PI.t)+cos(350.2.PI.t)-real.eps,width=0.45\linewidth} 
800: %\epsfig{file=pb-FFT-cos(50.2.PI.t)+cos(350.2.PI.t)-imag.eps,width=0.45\linewidth}} 
801: %\caption{Typical inaccuracies for the FFT routine with the function 
802: %$\cos(50.2\pi x)+\cos(350.2\pi x)$: Real (left) and imaginary (right) parts
803: %of the numerical Fourier transform.} \label{pb-fft}
804: %\end{figure}
805: %
806: However, and despite the fact we do believe these 
807: ``FFT time-split schemes'' realize the best numerical strategy in
808: terms of gridding, we 
809: shall point out some shortcomings of the method in the next section. 
810: 
811: \subsection{A rigorous framework for FFT-based schemes}
812: \label{sec:rig_error}
813: 
814: We present here a preliminary result about truncation errors 
815: in Lebesgue spaces for Fourier schemes; its proof follows directly
816: from the Strichartz estimates on the torus due to
817: J.~Bourgain \cite{bou} (see also \cite{bou_lect}),  and from the study
818: of FFT by M.~Taylor 
819: \cite[pp.~250--254]{taylor_PDE}. Its derivation is 
820: not obvious though as it applies directly to widely-used schemes like
821: the one recalled in the forthcoming section. We restrict our attention to  
822: the 1D free Schr\"odinger equation \eqref{eq:schrodlibre} with $\e=1$,
823: and periodic boundary  
824: conditions: $x \in {\mathbb T}:={\mathbb R}/2\pi{\mathbb Z}$.
825: 
826: Hence we start from 
827: $$
828: i\d_t \psi + \frac 1 2 \d_x^2 \psi=0, \qquad \psi(t=0,\cdot)=\zeta=
829: \sum_{j \in \Z} \hat \zeta_j e^{ijx},\quad 
830: x \in [0,2\pi]. 
831: $$
832: We have explicitly:
833: $$
834: \psi(t,x)=\sum_{j \in \Z} \hat \zeta_j e^{ij(x-jt/2)}.
835: $$
836: In order to investigate the behavior of the FFT-scheme involving a
837: finite even number $N \in 2\N$ of modes, we introduce the Discrete
838: Fourier Transform of a continuous function $f$ on $[0,2\pi]$ as follows:
839: $$
840: f^\#_k \stackrel{\text{def}}{=}\frac 1 N \sum_{j=1}^N f(2j\pi
841: /N)e^{-i 2\pi jk/N}. 
842: $$
843: From \cite[p.~252]{taylor_PDE}, we recall:
844: %
845: \begin{lemma}\label{lem_tay}If the continuous function $f$ has
846: a convergent Fourier series, then:
847: $$\forall k, \qquad f^\#_k = \sum_{j \in \Z} \hat f_{k +jN}.$$
848: \end{lemma}
849: %
850: Now, what the Fourier numerical scheme really computes is:
851: $$
852: \forall k \in \left\{-\frac N 2, \frac N 2\right \},
853: \qquad \hat \psi^\#_k(t)\stackrel{\text{def}}{=}e^{-i k^2 t/2}\hat
854: \zeta^\#_k. 
855: $$
856: Thus the corresponding numerical solution $\psi^{Num}$ is built:
857: $$
858: \psi^{Num}(t,x)=\sum_{k=-\frac N 2}^{\frac N 2}e^{i k(x-
859: kt/2)}\sum_{j \in \Z}\hat \zeta_{k +jN}. 
860: $$
861: The question is therefore to study the discrepancy between $\psi$ and
862: $\psi^{Num}$: 
863: \begin{align*}
864: \psi^{Num}(t,x)-\psi(t,x)&= \sum_{k=-\frac N 2}^{\frac N 2}e^{i
865: k(x- kt/2)}\sum_{j \in \Z}\hat \zeta_{k +jN} 
866: -\sum_{k \in \Z} \hat \zeta_k e^{ik(x-kt/2)}\\
867: &= \sum_{j \in \Z}\left(\sum_{k=-\frac N 2}^{\frac N 2}e^{i k(x-
868:   kt/2)}\hat \zeta_{k +jN} \right)- 
869: \sum_{k \in \Z} \hat \zeta_k e^{ik(x-kt/2)}\\
870: &=\sum_{j \in \Z^\star}\left(\sum_{k=-\frac N 2}^{\frac N 2}e^{i
871:   k(x- kt/2)}\hat \zeta_{k +jN} \right)+ 
872: \sum_{|k| > \frac N 2} \hat \zeta_k e^{ik(x-kt/2)}\\
873: &\stackrel{\text{def}}{=} I + II.
874: \end{align*}
875: The second term $II$ in this last equality can be interpreted as a
876: usual truncation error; it means that 
877: some high frequencies in the solution are lost when discretizing the
878: problem on a finite grid. From \cite[p.~253]{taylor_PDE}, we know that
879: the modulus of this term can be controlled as follows: 
880: $$
881: \sum_{|k| > \frac N 2} |\hat \zeta_k| \leq \frac 1 {N^m} \| \zeta \|_{C^{1+m}},
882: $$
883: which is satisfactory provided the initial data $\zeta$ is smooth,
884: e.g. $C^2(\R)$. However, the first term $I$ is far more delicate and
885: reveals the propagation of the errors coming from the DFT/FFT
886: itself. It can be controlled thanks to the  periodic Strichartz
887: estimates proved by J.~Bourgain. From \cite[pp.~16--17]{bou_lect}, we
888: recall: 
889: %
890: \begin{lemma}\label{lem_bou}
891: Given any complex-valued sequence $(a_k)_k$, 
892: the following estimates hold for the corresponding 
893: functions in $[0,2\pi] \times [0,2\pi]$:
894: \begin{align*}
895: &\Big\|\sum_k a_k e^{i(kx-k^2t)}\Big\|^2_{L^4({\mathbb T}\times
896:   {\mathbb T}) } \le C \sum_k |a_k|^2,\\  
897: &\Big\|\sum_{|k| \leq N} a_k e^{i(kx-k^2t)}\Big\|^2_{L^6({\mathbb
898:     T}\times {\mathbb T})} \le C e^{2\log 
899: N/\log \log N}\sum_k |a_k|^2. 
900: \end{align*}
901: \end{lemma}
902: %
903: The second estimate looks more attractive as it ``sees" the finite
904: number of modes. We therefore  deduce that the first term can be
905: controlled in $L^6$ by means of: 
906: $$
907: C \exp\left(\frac{\log N/2}{\log \log N/2}\right)
908: \left (\sum_{j \in \Z^\star, |k| \leq \frac N 2} |\hat
909:   \zeta_{k+jN}|^2\right)^{\frac 1 2}. 
910: $$
911: For instance, if $\zeta$ is a finite superposition of Fourier modes,
912: then it is clear that this 
913: term cancels for $N$ large enough as $j \not = 0$ in the summation;
914: obviously, the second term $II$ vanishes too. 
915: 
916: %*** On peut enoncer une proposition la-dessus en disant a partir de
917: %    quel N, une donnee initiale composee d'une serie de Fourier finie
918: %    (disons M termes) est resolue exactement. *** 
919: 
920: The general case isn't completely clear yet.
921: 
922: \section{Experiments on the focal point}\label{sec:foc}
923: \vspace*{-0.5pt}
924: \noindent
925: 
926: This section aims at visualizing the asymptotics previously recalled;  
927: namely we shall compare numerical approximations of \eqref{eq:nls} and
928: \eqref{eq:schrodlibre} 
929: in 1D ($n=1$) for various values of the parameters $\alpha$ and
930: $\sigma$. The initial wave function 
931: is rather simple:
932: $$
933: v^\e(t=0,x)=u^\e(t=0,x)=\exp \left(-\Big(2+\frac i {2\e}\Big)(x-\pi)^2\right), \qquad x\in [0,2\pi].
934: $$
935: Numerical results have been obtained through the time-splitting FFT
936: schemes recalled in the 
937: previous section; we used 1024 modes and fixed $\e=1/150$. It is
938: convenient to observe results 
939: in $t=2$ since $|v^\e(t=2,.)|=|v^\e(t=0,.)|$.
940: 
941: \subsection{Subcritical case}
942: \label{sec:foc_free}
943: 
944: This case corresponds to $\sigma=2$ and $\alpha=2.5$; we expect to
945: observe a decay of the absolute 
946: errors between $v^\e$ and $u^\e$ for values $\e \ll 1$. This is indeed
947: the case, but Fig.~\ref{pf1} 
948: shows even a bit more, namely it compares pointwise the following
949: quantities: (recall Lemma \ref{lem1}) 
950: $$\left\{\begin{array}{l}
951: \Re\left(v^\e(t=0,x)\exp(-i(x-\pi)^2/2\e)\right), \\
952: \Im\left(v^\e(t=2,x)\exp(i(x-\pi)^2/2\e)\right), \\
953: \Im\left(u^\e(t=2,x)\exp(i(x-\pi)^2/2\e)\right).
954: \end{array}\right.
955: $$
956: %
957: \begin{figure}[ht]
958: \centerline{\epsfig{file=Pt-Focal/al25-si4-erreur.eps,width=0.45\linewidth}
959: \epsfig{file=Pt-Focal/al25-si4-modules.eps,width=0.45\linewidth}}
960: \caption{Absolute errors on the wave functions (left) and on the
961:   modulus (right) at  
962: $T=2$ with $\sigma=2$ and $\alpha=2.5$. (Subcritical case)} \label{pf1}
963: \end{figure}
964: %
965: On the left in Fig.~\ref{pf1}, we obviously observe that the absolute
966: errors are slightly bigger when 
967: considering the solution of the nonlinear equation \eqref{eq:nls}, $u^\e$. 
968: However, even for the free solution $v^\e$, one sees that the error
969: doesn't vanish despite the fact 
970: no time-splitting algorithm is needed. As the way of discretizing the solution
971: reveals itself important, we include
972: here the corresponding {\sc Scilab} routine for the free equation:
973: 
974: {\tt clear;deff('[y]=phase(x)',['y=-0.5*(x-$\pi$)$.^2$';])
975: 
976: deff('[y]=position(x)',['y=exp(-2*(x-$\pi$)$.^2$)'])
977: 
978: deff('[y]=Az(x)',['y=position(x).*exp(i*phase(x)./epsilon)'])
979: 
980: NMAX=2$^{10}$;n=-(NMAX)/2:(NMAX/2)-1;epsilon=1.0/150;
981: 
982: XSTART=0;XSTOP=2*$\pi$;DX=(XSTOP-XSTART)/NMAX;XSTOP=XSTOP-DX;
983: 
984: a=XSTART:DX:XSTOP,initialdata=Az(a);
985: 
986: vepsilon=fftshift(fft(initialdata,-1));
987: 
988: vepsilon=exp(-i*epsilon*($n.^2$)).*vepsilon;
989: 
990: vepsilon=fft(fftshift(vepsilon),1);
991: %
992: %plot2d(a,[real(initialdata.*exp(-i*phase(a)/epsilon))'
993: %
994: %$\qquad$ imag(vepsilon.*exp(i*phase(a)/epsilon))'])
995: }\\
996: %
997: Clearly, its outcome is in agreement with Lemma
998: \ref{lem1} since $\e$ is already quite small. The Maslov index is visible, 
999: up to an error around $10^{-3}$ for 1024 Fourier modes.
1000: 
1001: \subsection{Critical case}
1002: 
1003: We now put $\sigma=\alpha=2$ and the outcome is displayed on
1004: Fig.~\ref{pf2}; we still compare the same quantities. Absolute errors
1005: on wave functions (left side) are much bigger for $u^\e$ in this
1006: case. 
1007: %
1008: \begin{figure}[ht]
1009: \centerline{\epsfig{file=Pt-Focal/al2-si4-erreur.eps,width=0.45\linewidth}
1010: \epsfig{file=Pt-Focal/al2-si4-modules.eps,width=0.45\linewidth}}
1011: \caption{Same as Fig.~\ref{pf1}, but $\sigma=2$ and $\alpha=2$. (Critical
1012: case)} \label{pf2} 
1013: \end{figure}
1014: %
1015: In particular, no new frequencies appear in the numerical
1016: solutions. The nonlinear effect boils down to a small change on the
1017: modulus of $u^\e$. 
1018: 
1019: \subsection{Supercritical case}
1020: 
1021: We close this first series of tests by considering  $\sigma=2$ and
1022: $\alpha=1.5$ as shown in Fig.~\ref{pf3}. Of course, as no pointwise
1023: convergence is expected in this case, absolute errors are even bigger
1024: for both wave functions (left side) and moduli (right
1025: side).  
1026: %
1027: \begin{figure}[ht]
1028: \centerline{\epsfig{file=Pt-Focal/al15-si4-erreur.eps,width=0.45\linewidth}
1029: \epsfig{file=Pt-Focal/al15-si4-modules.eps,width=0.45\linewidth}}
1030: \caption{Same as Fig.~\ref{pf1}, but $\sigma=2$ and
1031: $\alpha=1.5$. (Supercritical case)} \label{pf3} 
1032: \end{figure}
1033: %
1034: %What is interesting also is to observe that new oscillations have
1035: %appeared inside the nonlinear solution's modulus: it seems that a
1036: %frequency doubling manifests itself in $u^\e$. This is consistent with
1037: %the heuristics and the rigorous results in \cite{CaCascade}. 
1038: Of course, the size of 
1039: the error on the modulus is much bigger too, and one should be
1040: extremely careful about the credit to give to the numerical
1041: simulations in the supercritical case. Indeed, this is a regime where
1042: a small error can be amplified at leading order (see \cite{CaCascade,CaARMA}). 
1043: 
1044: \section{Visualization of the scattering operator}\label{sec:scattnum}
1045: \vspace*{-0.5pt}
1046: \noindent
1047: 
1048: We aim now at illustrating the results on scattering theory through
1049: numerical computations still achieved through time-splitting FFT
1050: schemes. The algorithm we used for the approximation of
1051: the scattering operator is based on a nonlinear time-splitting routine
1052: flanked by two free evolution steps (implemented the way recalled in
1053: the previous section): 
1054: $$
1055: S \psi \simeq U(-T)\circ U_{NL}(2T) \circ U(-T) \psi, \qquad T \gg 1,
1056: $$
1057: with $U$, $U_{NL}$ standing for the solution operators of equations
1058: \eqref{eq:schrodlibre} and \eqref{eq:nls} in 1-D with $\e =1$
1059: respectively. We used $T=55$ in the computations hereafter. 
1060:  
1061: As no small parameter $\e$ is present in the problem, one may think
1062: that no major obstacle exists in carrying out this program; this isn't
1063: correct as the free evolutions can (and do!) dramatically increase the
1064: size of the computational domain for large $T$. It is interesting to
1065: notice that, in case one wants to use FFT-based schemes, both the
1066: direct computation for small $\e$ and the scattering operator
1067: approximation lead to a ``large computational domain difficulty": in
1068: the Fourier space for the first case, in the usual space for the
1069: second. 
1070: 
1071: A way to understand the scattering operator is to
1072: visualize the average effects of the nonlinearities appearing in
1073: equations of the form 
1074: \begin{equation}
1075:   \label{eq:nls2}
1076:   i \d_t u_\lambda +\frac{1}{2}\partial_x^2 u_\lambda
1077:   =\lambda |u_\lambda|^{2\si}u_\lambda, \qquad u_\lambda(t=0,x)=\exp(-5 x^2),
1078: \end{equation}
1079: for various values of $\sigma \geq 1$ and $\lambda$. Intuitively, as
1080:  $\sigma$ increases, the nonlinearity becomes shorter
1081:  range. Similarly, as $\lambda$ increases, the nonlinearity becomes
1082:  stronger, and it should take a larger amount of time before we can
1083:  consider it has become negligible. In all the tests we performed, it
1084:  was somehow surprising to observe how fast the algorithm converges:
1085:  one does not have to consider ``very large'' values of $T$ so that  
1086: \begin{equation*}
1087: U(-T)\circ U_{NL}(2T) \circ U(-T) \psi
1088: \end{equation*}
1089: becomes stable and visually independent of $T$. 
1090: 
1091: \subsection{Quintic nonlinearity ($\sigma=2$)}
1092: 
1093: The parameter $\lambda$ controls in some sense the strength of the
1094: nonlinearity\footnote{through the stiffness of the associated
1095:   differential equation.} inside \eqref{eq:nls2}, as can be seen on
1096: Fig.~\ref{run1}. This figure displays the position density of the
1097: initial data, the scattered solution for $T=55$ and a ``mixed state"
1098: $\tilde u_\lambda(t=0)=U(-T)U_{NL}(T)u_\lambda(t=0)$. As our
1099: time-splitting/FFT algorithm preserves only the $L^2$ norm, but not
1100: the Hamiltonian, we first restricted ourselves to moderate values of
1101: $\lambda \geq 0$ (defocusing case).  
1102: %
1103: \begin{figure}[ht]
1104: \centerline{\epsfig{file=Scattering/Puissance_4/P4-C1.eps,width=0.45\linewidth}  
1105: \epsfig{file=Scattering/Puissance_4/P4-C5.eps,width=0.45\linewidth}}
1106: \caption{Initial data, numerical solution at $T=0$, scattering state
1107:   for \eqref{eq:nls2} with 
1108: $\sigma=2$ and $\lambda=1$ (left), $\lambda=5$ (right)} \label{run1}
1109: \end{figure}
1110: %
1111: However, as a numerical experiment, we wanted to display the outcome
1112: of our scheme for the stronger case $\lambda=25$ on Fig.~\ref{run2}:
1113: notice the change of shape in the scattered solution. Moreover, on
1114: this figure, we also tried to show what happens for $\lambda=-1$, that
1115: is to say for the focusing case despite there may be finite time
1116: blow-up (but there is scattering for small data). We checked that 
1117: the energy associated to this data is (and remains) 
1118: positive, a case where the virial identity , \cite{CazCourant},
1119: does not imply blow-up. The computational domain for these runs 
1120: was $[-100\pi, 100\pi]$ and $2^{13}-1$ Fourier modes were used. 
1121: %
1122: \begin{figure}[ht]
1123: \centerline{\epsfig{file=Scattering/Puissance_4/P4-C25.eps,width=0.45\linewidth} 
1124: \epsfig{file=Scattering/Puissance_4/P4-C-1.eps,width=0.45\linewidth}}
1125: \caption{Same as Fig.~\ref{run1}, but $\lambda=25$ (left),
1126:   $\lambda=-1$ (right)} \label{run2} 
1127: \end{figure}
1128: %
1129: \subsection{Power 3 ($\sigma=1.5$)}
1130: 
1131: Now let's observe the effects of lowering the $\sigma$ value while
1132: keeping other parameters equal, see Fig.~\ref{run3}. It is interesting
1133: to see that the change of shape appearing for $\lambda=25$ is stronger
1134: than in the preceding case. On the contrary, the increase of the
1135: numerical solution's support is slightly less important. 
1136: %
1137: \begin{figure}[ht]
1138: \centerline{\epsfig{file=Scattering/Puissance_3/P3-C5.eps,width=0.45\linewidth}  
1139: \epsfig{file=Scattering/Puissance_3/P3-C25.eps,width=0.45\linewidth}}
1140: \caption{Same as Fig.~\ref{run1}, but $\sigma=1.5$ and $\lambda=5$
1141:   (left), $\lambda=25$ (right)} \label{run3} 
1142: \end{figure}
1143: %
1144: This hints that increasing the $\sigma$ value tends to expand the
1145: support of the scattered solution whereas increasing the $\lambda \geq
1146: 0$ value (defocusing case) leads to an oscillatory behavior. However,
1147: we stress that since the energy, 
1148: $$
1149: E(t):=\frac 1 2 \|\d_x u^\lambda (t) \|_{L^2}^2 + 
1150: \frac \lambda {\sigma + 1}\| u^\lambda(t)\|_{L^{2\si +2}}^{2\si +2}=E(0),
1151: $$
1152: of the numerical solution
1153: changes more with a bigger $\lambda$ (its mass being always kept
1154: constant), these oscillations might be spurious. We actually don't
1155: know how this fact can be decided; our profiles have been checked to
1156: be stable on a finer grid. 
1157: 
1158: \subsection{Power 6 ($\sigma=3$)}
1159: 
1160: In order to get some numerical evidence about the dependence of the
1161: scattered solution on $\sigma$, we display on Fig.~\ref{run4} the
1162: outcome for $\sigma=3$. It is quite clear that the scattered solutions
1163: for both values of $\lambda$ are less peaked. Their support is bigger
1164: and the oscillations for $\lambda=25$ are weaker, their frequency
1165: remained the same though. 
1166: %
1167: \begin{figure}[ht]
1168: \centerline{\epsfig{file=Scattering/Puissance_6/P6-C5.eps,width=0.45\linewidth}
1169: \epsfig{file=Scattering/Puissance_6/P6-C25.eps,width=0.45\linewidth}}
1170: \caption{Same as Fig.~\ref{run1}, but $\sigma=3$ and $\lambda=5$
1171:   (left), $\lambda=25$ (right)} \label{run4} 
1172: \end{figure}
1173: %
1174: This agrees with the behavior we sketched in the preceding subsection
1175: as $\sigma$ and $\lambda$ vary. 
1176: 
1177: \section{Experiments on a cusp caustic}\label{sec:cusp}
1178: \vspace*{-0.5pt}
1179: \noindent
1180: 
1181: Let us now go back to comparing the quadratic observables generated by
1182: numerical approximation of equations \eqref{eq:nls} and
1183: \eqref{eq:schrodlibre} endowed with a small parameter $\e$ in 1-D. In
1184: this section we fixed $\e = 1/150$. Figure \ref{run-cusp} displays the
1185: position density of the initial data for both equations, i.e. 
1186: $$
1187: u^\e(t=0,x)=v^\e(t=0,x)=\exp\Big(-2(x-\pi)^2-i\cos(x)/\e \Big), \qquad
1188: x \in [0,2\pi], 
1189: $$
1190: together with the position density of the numerical approximations of
1191: \eqref{eq:nls}, \eqref{eq:schrodlibre} in $T=3.5$. The point here is
1192: to investigate what happens for the case of such a self-interfering
1193: Gaussian pulse, since no scattering theory is known for this
1194: problem. What we would like to check is whether the theoretical
1195: results on the focus point recalled and visualized in the preceding
1196: sections can be thought of as a guideline for this more complex case
1197: involving a non-trivial caustic. 
1198: %
1199: \begin{figure}[p]
1200: \centerline{\epsfig{file=Cusp/gaussienne/al4-si4.eps,angle=270,width=0.7\linewidth}} 
1201: \centerline{\epsfig{file=Cusp/gaussienne/al3-si4.eps,angle=270,width=0.7\linewidth}}
1202: \centerline{\epsfig{file=Cusp/gaussienne/al2-si4.eps,angle=270,width=0.7\linewidth}}
1203: \caption{Position densities in the cusp caustic: $\alpha=4,3,2$ (top to bottom)} \label{run-cusp}
1204: \end{figure}
1205: %
1206: We shall observe position densities for the unique value of $\sigma=4$
1207: as a similar behavior has been seen to hold for different
1208: nonlinearities with convenient values of $\alpha$. $4095$ Fourier
1209: modes have been used in order to produce these results. 
1210: 
1211: \subsection{Subcritical picture: $\alpha=4$}
1212: 
1213: This case could be referred to as subcritical since it is noticeable on
1214: the top of Fig.~\ref{run-cusp} that the free and the nonlinear
1215: numerical solutions do agree for this reasonably small value of
1216: $\e$. In particular, the frequencies of oscillations are
1217: identical. This is very similar compared to the behavior investigated
1218: in \cite{CaIUMJ}. 
1219: 
1220: \subsection{Critical picture: $\alpha=3$}
1221: 
1222: The parameter $\alpha$ is now in a ``critical range" as we observe
1223: that both solutions differ much more, but the frequency of the
1224: oscillations looks like being still the same in both cases. In order
1225: to establish this fact, we display on the left of Fig.\ref{fft-cusp}
1226: the FFT of the position densities: a peak at the same frequency is
1227: clearly noticeable.  
1228: %
1229: \begin{figure}[ht]
1230: \centerline{\epsfig{file=Cusp/gaussienne/fft-al3-si4.eps,width=0.45\linewidth}
1231: \epsfig{file=Cusp/gaussienne/fft-al2-si4.eps,width=0.45\linewidth}}
1232: \caption{Fourier transforms of the position densities with $\alpha=3$
1233:   (left), $\alpha=2$ (right)} \label{fft-cusp} 
1234: \end{figure}
1235: %
1236: The nonlinear effect manifests itself through a change of order zero
1237: in the moduli, as we already observed on the right side of
1238: Fig.~\ref{pf2}; notice also the similarity with the scattering state
1239: shown on Fig.~\ref{run2} (right). This does agree with the $\alpha_c$ 
1240: value derived in Section \ref{sec:heur} 
1241: 
1242: \subsection{Supercritical picture: $\alpha=2$}
1243: 
1244: In this last case, there is no similarity no more between the
1245: approximate solutions of \eqref{eq:nls}, \eqref{eq:schrodlibre}, as
1246: seen on both Fig.~\ref{run-cusp} and \ref{fft-cusp}. Especially, the
1247: right side of Fig.~\ref{fft-cusp} reveals that new frequencies show up
1248: inside the position density of the nonlinear solution. We have
1249: therefore a change of order zero in the moduli and in the
1250: frequency. This is of course reminiscent of Fig.~\ref{pf3} in which a
1251: frequency doubling seems to show up in the supercritical regime. 
1252: 
1253: \section{Conclusion}\label{sec:concl}
1254: \vspace*{-0.5pt}
1255: \noindent
1256: 
1257: We have presented the semi-classical limit for the nonlinear
1258: Schr\"odinger equation in the presence of a caustic. When the caustic
1259: is reduced to a point, the numerical experiments are in good agreement
1260: with the analytical results as far as the notion of criticality is
1261: concerned. However in the critical case, described by a nonlinear
1262: scattering operator, the leading order nonlinear effects are rather
1263: hard to visualize in the semi-classical limit. This is why we
1264: simulated the scattering operator in a separate way. 
1265: 
1266: Our numerical
1267: tests give encouraging evidence of new phenomena concerning the
1268: phase of the wave in the supercritical case when a focal point is
1269: formed (appearance of new frequencies). In the presence of a cusp caustic, 
1270: the numerical experiments are in
1271: good agreement with the heuristic arguments that we presented here, for
1272: which no rigorous justification is available so far. 
1273: 
1274: 
1275: 
1276: \begin{thebibliography}{10}
1277: \bibitem{BJM1}
1278: {W.~Bao, S.~Jin, and P.~A. Markowich}, {\em On time-splitting spectral
1279:   approximations for the {S}chr{\"o}dinger equation in the
1280:   semiclassical regime}, 
1281:   J. Comput. Phys., \textbf{175} (2002), no.~2, 487--524.
1282: 
1283: \bibitem{BJM2}
1284: {W.~Bao, S.~Jin, and P.~A. Markowich}, {\em Numerical study of
1285:   time-splitting spectral discretizations of nonlinear {S}chr{\"o}dinger
1286:   equations in the semiclassical regimes}, SIAM J. Sci. Comput.,
1287: \textbf{25} (2003), no.~1, 27--64.
1288: 
1289: \bibitem{Barab}
1290: {J.~E. Barab}, {\em Nonexistence of asymptotically free solutions for
1291:   nonlinear {S}chr{\"o}dinger equation}, J. Math. Phys., \textbf{25}
1292:  (1984), 3270--3273.
1293: 
1294: \bibitem{BBD}
1295: {C.~Besse, B.~Bid{\'e}garay, and S.~Descombes}, {\em Order estimates in
1296:   time of splitting methods for the nonlinear {S}chr{\"o}dinger equation}, SIAM
1297:   J. Numer. Anal., \textbf{40} (2002), no.~1, 26--40.
1298: 
1299: \bibitem{bou}
1300: {J. Bourgain}, {\em Fourier restriction phenomena for certain lattice subsets
1301: and applications to nonlinear evolution equations. I. Schr\"odinger
1302: equations}, Geom. and Funct. Anal., \textbf{3} (1993), 107--156.
1303: 
1304: \bibitem{bou_lect}
1305: {J. Bourgain}, {\em Nonlinear Schr\"odinger equations}, in:
1306: Hyperbolic equations and frequency interactions, eds. L. Cafarelli and
1307: W. E. IAS/Park City vol. 5. 
1308: 
1309: \bibitem{BurqZworski}
1310: {N.~Burq and M.~Zworski}, {\em Instability for the semiclassical
1311:   non-linear {S}chr\"odinger equation}, Comm. Math. Phys. \textbf{260}
1312: (2005), no.~1, 45--58.
1313: 
1314: \bibitem{CaIUMJ}
1315: {R.~Carles}, {\em Geometric optics with caustic crossing for some nonlinear
1316:   {S}chr{\"o}dinger equations}, Indiana Univ. Math. J., 49 (2000),
1317: pp.~475--551. 
1318: 
1319: \bibitem{CaWigner}
1320: {R.~Carles}, {\em Remarques sur les
1321:   mesures de {W}igner}, C. {R}. {A}cad. {S}ci. {P}aris, t. 332,
1322: {S}{\'e}rie {I}, 
1323:   332 (2001), pp.~981--984.
1324: 
1325: \bibitem{CaIHP}
1326: {R.~Carles}, {\em Semi-classical
1327:   {S}chr{\"o}dinger equations with harmonic potential and nonlinear
1328:   perturbation}, Ann. Inst. H. Poincar{\'e} Anal. Non Lin{\'e}aire,
1329: \textbf{20} (2003), no.~3, 501--542.
1330: 
1331: \bibitem{CaCascade}
1332: {R.~Carles}, {\em Cascade of phase
1333:   shifts for nonlinear {S}chr{\"o}dinger equations}, J. Hyperbolic
1334: Differ. Equ.,   (2007).
1335: \newblock To appear.
1336: 
1337: \bibitem{CaARMA}
1338: {R.~Carles}, {\em Geometric optics and
1339:   instability for semi-classical {S}chr{\"o}dinger equations}, Arch. Ration.
1340:   Mech. Anal. \textbf{183} (2007), no.~3, 525--553.
1341: 
1342: 
1343: \bibitem{CaBKW}
1344: {R.~Carles}, {\em {WKB} analysis for
1345:   nonlinear {S}chr{\"o}dinger equations with potential},
1346: Comm. Math. Phys. \textbf{269} (2007), no.~1, 195--221.
1347: 
1348: \bibitem{CazCourant}
1349: {T.~Cazenave}, {\em Semilinear {S}chr{\"o}dinger equations}, vol.~10 of
1350:   Courant Lecture Notes in Mathematics, New York University Courant Institute
1351:   of Mathematical Sciences, New York, 2003.
1352: 
1353: \bibitem{CW92}
1354: {T.~Cazenave and F.~Weissler}, {\em Rapidly decaying solutions of the
1355:   nonlinear {S}chr{\"o}dinger equation}, Comm. Math. Phys. \textbf{147} (1992),
1356:   75--100.
1357: 
1358: \bibitem{DGPS}
1359: {F.~Dalfovo, S.~Giorgini, L.~P. Pitaevskii, and S.~Stringari}, {\em Theory
1360:   of {B}ose-{E}instein condensation in trapped gases},
1361: Rev. Mod. Phys., \textbf{71} 
1362:   (1999), no.~3,  463--512.
1363: 
1364: \bibitem{Du}
1365: {J.~J. Duistermaat}, {\em Oscillatory integrals, {L}agrange immersions and
1366:   unfolding of singularities}, Comm. Pure Appl. Math. \textbf{27}
1367: (1974), 207--281. 
1368: 
1369: \bibitem{DSS}
1370: {A.~Dur{\'a}n and J.~M. Sanz-Serna}, {\em The numerical integration of
1371:   relative equilibrium solutions. {T}he nonlinear {S}chr{\"o}dinger equation},
1372:   IMA J. Numer. Anal. \textbf{20} (2000), 235--261.
1373: 
1374: \bibitem{Ginibre}
1375: {J.~Ginibre}, {\em An introduction to nonlinear {S}chr{\"o}dinger equations},
1376:   in: {Nonlinear waves (Sapporo, 1995)}, eds. R.~Agemi, Y.~Giga,
1377:   and T.~Ozawa,
1378:   GAKUTO International Series, Math. Sciences and Appl., Gakk\={o}tosho, Tokyo,
1379:   1997, pp.~85--133.
1380: 
1381: \bibitem{GOV94}
1382: {J.~Ginibre, T.~Ozawa, and G.~Velo}, {\em On the existence of the wave
1383:   operators for a class of nonlinear {S}chr{\"o}dinger equations}, {A}nn. {IHP}
1384:   ({P}hysique {T}h{\'e}orique) \textbf{60} (1994), 211--239.
1385: 
1386: \bibitem{GV79Scatt}
1387: {J.~Ginibre and G.~Velo}, {\em On a class of nonlinear {S}chr{\"o}dinger
1388:   equations. {II} {S}cattering theory, general case},
1389: J. Funct. Anal. \textbf{32} (1979), 33--71.
1390: 
1391: \bibitem{GosseWKB}
1392: {L.~Gosse}, {\em A case study on the reliability of multiphase {WKB}
1393:   approximation for the one-dimensional {S}chr{\"o}dinger equation},
1394: in: {Numerical 
1395:   methods for hyperbolic and kinetic problems}, vol.~7 of IRMA Lect. Math.
1396:   Theor. Phys., Eur. Math. Soc., Z{\"u}rich, 2005, pp.~131--141.
1397: 
1398: \bibitem{GosseJinLi}
1399: {L.~Gosse, S.~Jin, and X.~Li}, {\em Two moment systems for computing
1400:   multiphase semiclassical limits of the {S}chr{\"o}dinger equation}, Math.
1401:   Models Methods Appl. Sci. \textbf{13} (2003), no.~12, 1689--1723.
1402: 
1403: \bibitem{Hormander}
1404: {L.~H{{\"o}}rmander}, {\em The analysis of linear partial differential
1405:   operators}, Springer-Verlag, Berlin, 1994.
1406: 
1407: \bibitem{HK87}
1408: {J.~Hunter and J.~Keller}, {\em Caustics of nonlinear waves}, Wave
1409: motion \textbf{9} (1987), 429--443.
1410: 
1411: \bibitem{JMRTAMS95}
1412: {J.-L. Joly, G.~M{\'e}tivier, and J.~Rauch}, {\em Focusing at a point and
1413:   absorption of nonlinear oscillations}, Trans. Amer. Math. Soc.
1414: \textbf{347} (1995), no.~10, 3921--3969.
1415: 
1416: \bibitem{JMRMemoir}
1417: {J.-L. Joly, G.~M{\'e}tivier, and J.~Rauch}, {\em Caustics for
1418:   dissipative semilinear oscillations},
1419: Mem. Amer. Math. Soc. \textbf{144} (2000), no.~685, 
1420:   pp.~viii+72.
1421: 
1422: \bibitem{koko}
1423: {G.~T. Kossioris}, {\em Formation of singularities for viscosity solutions
1424:   of {H}amilton-{J}acobi equations in one space variable}, Comm. Partial
1425:   Differential Equations  \textbf{18} (1993), 747--770.
1426: 
1427: \bibitem{Ludwig}
1428: {D. Ludwig}, {\em Uniform asymptotic expansions at a caustic},
1429: Comm. Pure Appl. Math. \textbf{19} (1966), 215--250.
1430: 
1431: \bibitem{MPP}
1432: {P.~A. Markowich, P.~Pietra, and C.~Pohl}, {\em Numerical approximation of
1433:   quadratic observables of {S}chr{\"o}dinger-type equations in the
1434:   semi-classical limit}, Numer. Math.  \textbf{81} (1999), 595--630.
1435: 
1436: \bibitem{NakanishiOzawa}
1437: {K.~Nakanishi and T.~Ozawa}, {\em Remarks on scattering for nonlinear
1438:   {S}chr{\"o}dinger equations}, NoDEA Nonlinear Differential Equations
1439: Appl. \textbf{9} (2002), no.~1, 45--68.
1440: 
1441: \bibitem{PiSt}
1442: {L.~Pitaevskii and S.~Stringari}, {\em Bose-{E}instein condensation},
1443:   vol.~116 of International Series of Monographs on Physics, The Clarendon
1444:   Press Oxford University Press, Oxford, 2003.
1445: 
1446: \bibitem{Strauss74}
1447: {W.~A. Strauss}, {\em Nonlinear scattering theory}, in: {Scattering
1448:   theory in 
1449:   mathematical physics}, eds. J.~Lavita and J.~P. Marchand (Reidel, 1974).
1450: 
1451: \bibitem{Strauss81}
1452: {W.~A. Strauss}, {\em Nonlinear scattering
1453:   theory at low energy}, J. Funct. Anal.  \textbf{41} (1981), 110--133.
1454:   
1455: \bibitem{taylor_PDE}
1456: {M. Taylor}, {\em Partial Differential Equations, Vol. I},
1457: Springer, New York, 1996.  
1458: 
1459: 
1460: \end{thebibliography}
1461: 
1462: 
1463: 
1464: \end{document}
1465: 
1466: